You are on page 1of 707

DURABILITY AND SUSTAINABILITY OF FIBER

REINFORCED POLYMER (FRP) COMPOSITES


FOR CONSTRUCTION AND REHABILITATION
Proceedings of the Fourth
International Conference (CDCC-11)
Qubec (Qubec) Canada July 20-22, 2011

DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES


EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION
Comptes rendus de la quatrime
Confrence Internationale (CDCC-11)
Qubec (Qubec) Canada 20-22 juillet 2011

Editors diteurs

Brahim Benmokrane

Universit de Sherbrooke
Sherbrooke (Qubec) Canada

Ehab El-Salakawy

University of Manitoba
Winnipeg (Manitoba) Canada

Ehab Ahmed

Universit de Sherbrooke
Sherbrooke (Qubec) Canada

DOCUMENTATION

The organizing committee and the sponsors of the CDCC-11 are not
responsible for the statements or opinions expressed in this publication.
The papers contained herein are published in the exact form submitted by the
authors. Any statements of view expressed in the papers are those of the
authors. Mention of trade names or commercial products does not constitute
endorsement or recommendation for use.
Le comit dorganisation ainsi que les commanditaires de la CDCC-11
dclinent toute responsabilit quant aux affirmations et points de vue exprims
dans le prsent ouvrage.
Les manuscrits contenus dans ce volume sont prsents dans la forme tablie
par les auteurs. Toutes les opinions avances dans les manuscrits
appartiennent leurs auteurs. Lutilisation des marques commerciales et des
noms de produits ne constitue en aucun cas un endossement ou une
recommandation.

ISBN: 978-2-7622-0196-3

While supplies last, additional copies of the proceedings may be obtained


from / On peut se procurer des exemplaires additionnels des comptes rendus
(Jusqu puisement des stocks) en sadressant :
CDCC-11
Department of Civil Engineering
Faculty of Engineering
Universit de Sherbrooke
Sherbrooke (Qubec)
CANADA J1K 2R1

FOREWORD
Applications of fibre reinforced polymer (FRP) composites in civil structures
have increased significantly in recent years. Since most of these structures are
designed with service lives of more than 50 years, the long term durability of
FRP materials in these applications is a critical issue.
The present conference is the fourth of its kind focusing on the durability,
sustainability and field applications of fibre reinforced polymers (FRP) for
construction and rehabilitation of structures (bridges, buildings, marine
structures, etc.). The first conference was held in Sherbrooke, Qubec, Canada
on August 1998, followed by one in Montral, Qubec, Canada on May 2002.
The third one was held in Qubec City, Qubec, Canada on May 2007. The
objective of this series of internationals conferences is to provide a forum to
academics, researchers, engineers, structure owners, FRP manufacturers, and
delegates from public and industrial institutions to present and exchange views
on present and future research on long-term durability and sustainability of
FRP composite materials in bridges and other infrastructure systems exposed
to harsh environmental conditions.
A total of 75 peer-reviewed papers are included in the proceedings. They have
been authored by various experts in the field from 19 different countries
around the world, including Algeria, Australia, Canada, China, Egypt, France,
Germany, Hong Kong, Italy, Iran, Japan, KSA, Netherland, Portugal,
Singapore, South Korea, Switzerland, United Kingdom and USA. On behalf
of the organizing committee, I am very thankful to the authors for their
valuable contributions to the conference. I am also very grateful to the paper
reviewers for their time and diligence.
We are thankful to the Department of Civil Engineering, Faculty of
Engineering, the Universit de Sherbrooke for their tremendous organizational
support without which this conference would not be possible. Financial and/or
logistic support from the conference sponsors: The ISIS Canada Network
Association, the Natural Science and Engineering Research Council of Canada
(NSERC), the Ministry of Transportation of Quebec (MTQ), the Research
Centre on Applied Polymers and Composites (CREPEC), the Research Centre
on Concrete Infrastructures (CRIB), and the American Composites
Manufacturers Association (ACMA) are gratefully acknowledged. We also
thank our corporate co-sponsors for their active participation.

We have been fortunate to receive help from a host of people in organizing


this conference. Professors Charles E. Bakis, Pennsylvania State University,
Nemkumar Banthia, University of British Columbia and John Myers, Missouri
University of Science and Technology deserve special mention. We are very
grateful to Mrs. Nathalie Valle, the Conference Secretary and Drs. Ahmed
Farghaly, Robert Mathieu, Patrice Cousin, Hamdy Mohamed, Brahim
Tighiouart and Mr. Christian Dulude for their assistance with the various tasks
to ensure the success of this conference.
We extend our heartiest welcome to the conference participants coming from
many near and far parts of the world.
The Editors,
Dr. Brahim Benmokrane, P. Eng.
NSERC & Canada Research Chair
Professor
Department of Civil Engineering
Faculty of Engineering
University of Sherbrooke

Dr. Ehab El-Salakawy, P. Eng.


Canada Research Chair Professor
Department of Civil Engineering
Faculty of Engineering
University of Manitoba

Dr. Ehab Ahmed, P. Eng.


Department of Civil Engineering
Faculty of Engineering
University of Sherbrooke
June 10, 2011

AVANT-PROPOS
Lutilisation des matriaux composites en polymres renforcs de fibres (PRF)
comme matriaux de construction dans le domaine du gnie civil a augment
de faon significative au cours des dernires annes. Compte tenu que les
structures de gnie civil sont conues pour des dures de service de plus de 50
ans, la durabilit long terme des matriaux composites de PRF constitue un
facteur majeur pour ce type dapplications.
La prsente confrence est la quatrime du genre qui traite de thmes se
rapportant principalement la durabilit, au cycle de vie et les applications sur
le terrain des matriaux composites en PRF pour la construction et la
rhabilitation de structures. Les ditions prcdentes ont eu lieu Sherbrooke,
Qubec, Canada en 1998, Montral, Qubec, Canada en 2002 et Qubec,
Qubec, Canada en 2007. Lobjectif de cette srie de confrences
internationales est de procurer un forum aux universitaires, chercheurs,
ingnieurs, propritaires douvrages, manufacturiers de produits de PRF et
membres dinstitutions publiques ou industrielles pour prsenter et changer
leurs points de vue sur les recherches prsentes et futures sur la durabilit
long terme et le cycle de vie des matriaux composites de PRF pour les ponts
routiers et dautres structures exposes des conditions environnementales
svres.
Soixante-quinze (75) articles valus sont inclus dans le compte-rendu de la
confrence. Ces articles ont t rdigs par plusieurs spcialistes dans le
domaine provenant de 19 pays dont lAlgrie, lAustralie, le Canada, la Chine,
lgypte, la France, lAllemagne, la Hong Kong, lItalie, lIran, le Japon,
lArabie Saoudite, la Hollande, le Portugal, Singapore, la Core du Sud, la
Suisse, lAngleterre, et les tats-Unis dAmrique. Au nom du comit
dorganisation, je remercie vivement les auteurs pour leurs contributions trs
importantes cette confrence. Je remercie aussi toutes les personnes qui ont
valu les articles pour leur temps et assiduit.
Nous remercions le Dpartement de gnie civil de la Facult de gnie de
lUniversit de Sherbrooke pour son support exceptionnel sans lequel cette
confrence naurait pas t possible. Les supports financiers et/ou logistiques
des commanditaires de cette confrence : ISIS Canada Network Association, le
Conseil de recherche en sciences naturelles et en gnie du Canada (CRSNG),
le ministre des transports du Qubec (MTQ), le Centre de recherche en
plasturgie et composites (CREPEC), le Centre de recherche sur les

infrastructures en bton (CRIB) et lAmerican Composites Manufacturers


Association (ACMA) sont grandement apprcis. Nous tenons aussi
remercier les co-commanditaires pour leur participation active.
Lorganisation de cette confrence a t rendue possible grce laide et la
participation de plusieurs personnes. Les professeurs Charles E. Bakis, de
Pennsylvania State University, Nemkumar Banthia, de l'University of British
Columbia et John Myers, de Missouri University of Science and Technology
mritent dtre particulirement mentionns. Nous sommes aussi
reconnaissants envers Mme Nathalie Valle, Secrtaire de la confrence et
MM. Ahmed Farghaly, Mathieu Robert, Patrice Cousin, Hamdy Mohamed,
Brahim Tighiouart et Christian Dulude pour leur assiduit dans lexcution de
leurs tches respectives qui garantissent le succs de cette confrence.
Nous souhaitons la bienvenue la plus cordiale toutes les participantes et
tous les participants qui viennent de pays proches et lointains assister cette
confrence.
Les diteurs,
Brahim Benmokrane, P. Eng.
Titulaire de Chaires du CRSNG
et du Canada
Dpartement de gnie civil
Facult de gnie
Universit de Sherbrooke

Ehab El-Salakawy, P. Eng.


Titulaire de Chaire du Canada
Dpartement de gnie civil
Facult de gnie
Universit de Manitoba

Ehab Ahmed, P. Eng.


Dpartement de gnie civil
Facult de gnie
Universit de Sherbrooke
10 juin 2011

ORGANIZING COMMITTEE / COMIT DORGANISATION

Honorary Chair of the Conference / Prsident honorifique de la confrence


Daniel Bouchard, Ministre des Transports du Qubec
Conference Chair / Prsident de la confrence
Brahim Benmokrane, Universit de Sherbrooke
Editorial Chair / Directeur scientifique
Ehab El-Salakawy, University of Manitoba
Co-Chairs / Co-prsidents
Charles E. Bakis, Pennsylvania State University
Nemkumar Banthia, University of British Columbia
John Myers, Missouri University of Science and Technology
Program Co-ordinator / Coordinateur du programme
Ahmed Farghaly, Universit de Sherbrooke
Information System Assistant / Coordonnateur des systmes d'information
Ehab Ahmed, Universit de Sherbrooke
Administrative Assistant / Assistante administrative
Nathalie Valle, Universit de Sherbrooke

NATIONAL ORGANIZING COMMITTEE /


COMIT ORGANISATEUR NATIONAL
Nemkumar Banthia, University of British Columbia
Brahim Benmokrane, University of Sherbrooke
Omar Chaallal, cole de technologie suprieure
Grard Desgagn, Ministry of Transportation of Quebec
Mario Desroches, Sika Canada
Bernard Drouin, Pultrall Inc.
Raafat El-Hacha, University of Calgary
Ehab El-Salakawy, University of Manitoba
Darrell Evans, P.E.I. Department of Transportation
Garth Fallis, Vector Construction Group
Mark Green, Queens University
Pierre Labossire, University of Sherbrooke
David Lai, Ministry of Transportation of Ontario
Aftab Mufti, University of Manitoba
Kenneth Neale, University of Sherbrooke
John Newhook, Dalhousie University
Ghani Razaqpur, McMaster University
Murat Saatcioglu, University of Ottawa
Shamim Sheikh, University of Toronto
Khaled Soudki, University of Waterloo
Allan Wiseman, Public Works Canada (PWGSC)

INTERNATIONAL SCIENTIFIC COMMITTEE /


COMIT SCIENTIFIQUE INTERNATIONAL
R. Al-Mahaidi, Australia

I. Mahfouz, Egypt

L. Bank, USA

U. Meier, Switzerland

N. Banthia, Canada

A. Mosallam, USA

J. Barros, Portugal

G. Monti, Italy

B. Benmokrane, Canada

A. Mufti, Canada

L. Bisby, UK

A. Nanni, USA

O. Chaallal, Canada

K. Neale, Canada

J.F. Chen, UK

D.J. Oehlers, Australia

O. Cosenza, Italy

R. Razaqpur, Canada

C. Dolan, USA

S. Rizkalla, USA

E. El-Salakawy, Canada

R. Sen, USA

M.A. Erki, Canada

C. Shield, USA

A. Fam, Canada

J. Sim, Korea

H. Fukuyama, Japan

S.T. Smith, China

H. GangaRao, USA

L. Taerwe, Belgium

P. Hamelin, France

B. Taljsten, Demark

T. Hamilton, USA

K.H. Tan, Singapore

K. A. Harries, USA

J.G. Teng, China

V.M. Karbhari, USA

R. Tepfers, Sweden

T. Keller, Switzerland

T. Triantafillou, Greece

K. Kobayashi, Japan

T. Uomoto, Japan

J.J. Lesko, USA

Z.S. Wu, Japan

SPONSORS / COMMANDITAlRES

Dpartement de gnie civil, Facult de


gnie, Universit de Sherbrooke
ISIS Canada Network Association

Natural Science and Engineering Research Council


of Canada (NSERC) / le Conseil de recherche en
sciences naturelles et en gnie du Canada (CRSNG)
Ministre des transports du Qubec (MTQ)

The International Institute for FRP in Construction

Research Center on Concrete Infrastructures


Centre de recherche sur les infrastructures en bton

The Canadian Society for Civil Engineering


Socit canadienne de gnie civil

Center for Applied Research on Polymers and Composites


Centre de recherche en plasturgie et composites

American Composites Manufacturers Association

CO-SPONSORS / CO-COMMANDITAlRES
Avensys Inc.
BP composites Ltd
Canadian Association for Composite Structures and Materials
(CACSMA)
Carbonfibreplus
FiReP Rebar Canada Inc.
Hughes Brothers Inc.
Hydro-Qubec
Industrial Materials Institute of NRC (IMI)
ITF Labs
MagmaTech Ltd
Ministre des transports du Qubec (MTQ)
Osmos Canada Inc.
Public Works and Governmental Services of Canada (PWGSC)
Pultrall Inc.
Roctest Ltd.
Schck Canada Inc.
Sika Canada Inc.
Vector Construction Inc.

TABLE OF CONTENTS / TABLE DES MATIRES

Alternative Environmental Knockdown Factors (CE) Derived From


Extensive Experimental Data .....................................................................................

K.A. Harries and B.M. Shahrooz


A Field Study on Durability of Bond Strength between CFRP Wraps and
Concrete.........................................................................................................................

11

S.A. Sheikh, A.D. Caspary and M. Stafford


Bond Behavior of Externally Bonded Fiber Reinforced Polymer
Laminates Subjected to In-Situ Service Loading and Environmental
Conditioning ..................................................................................................................

19

J.J. Myers and N. Muncy


Durability of GFRP Rods in Field Demonstration Projects across Canada ...........

27

A.A. Mufti, J. Newhook, B. Benmokrane, G. Tadros and H.M. Vogel


Effect of Temperature Change on the Behavior of FRP-to-Concrete
Bonded Joints ................................................................................................................

37

W.Y. Gao, J.G. Teng and J.G. Dai


Long-term Bond Behavior of GFRP Rebars in Severe Environments ....................

47

B. Juette, A. Weber and C. Witt


Effect of Various Severe Environmental Exposures on CFRP and SFRP
Confined Circular Concrete Columns ........................................................................

55

M.A. Mashrik and R. El-Hacha


Temperature Effect on Curing Time Effects on the Creep Response of
Fiber Reinforced Polymer Composites Bonded to Concrete ....................................

63

Y. Jeong, A. Jaipuriar, M.M. Lopez and C.E. Bakis


Bond Behaviour of GFRP Bars to Concrete Subjected to Fatigue
Loading and Freeze-Thaw Cycles ...............................................................................

71

J.R. Alves, A.A. El-Ragaby and E. El-Salakawy


Durability Characteristics of New GFRP Dowels for Concrete Pavement .............
M. Montaigu, M. Robert and B. Benmokrane

81

Evaluation of the Durability of CFRP Sheets after Climatic Exposure ..................

91

I. Nishizaki, P. Labossiere, K.W. Neale, M. Demers and T. Tomiyama


Durability of Externally Bonded FRP Composites: Comparison of
Accelerated Test Methods with Real Time Exposure .............................................. 101
H.R. Hamilton, C.W. Dolan and J.E. Tanner
Tensile Capacity of Stressed CFRP Strand Exposed to Extreme
Aggressive Groundwater Environments ................................................................... 109
M. Sentry, R. Al-Mahaidi, A. Bouazza and L. Carrigan
The Effect of Long-Term Loading on the FRP-Concrete Bonding
Strength ........................................................................................................................ 119
C. Mazzotti and M. Savoia
Design Optimization of Composite Structures Considering Bonded Joins
Durability ...................................................................................................................... 129
P. Hamelin, A. Si Larbi and E. Ferrier
Durability of Glass FRP Bars Subjected to Extreme Environmental and
Loading Conditions ...................................................................................................... 139
M. Robert, A. Fam and B. Benmokrane
Fire Testing of FRP Strengthened Reinforced Concrete Columns.......................... 151
N. Bnichou, D. Cree, E.U. Chowdhury, M.F. Green and L.A. Bisby
An Investigation into the Sustainability of FRP Reinforcing Bars .......................... 159
M.C.K. Pearson, T. Donchev and M. Limbachiya
Methods for Evaluating Long-Term Performance of FRP-Structural
Systems Applied to Concrete Bridges ......................................................................... 167
K. Crawford
Creep Behaviour of GFRP Cables, Short and Long Term Test ............................. 175
E. Ferrier, A. Si-Larbi and P. Hamelin
Durability of GFRP bars under Low Temperature Environment and
High Temperature Alkali Exposure .......................................................................... 185
S. A. Sheikh, D.T.C. Johnson and A.D. Caspary
Durability of FRP-Concrete Interface under Fatigue Loading ............................... 195
C. Carloni1, K.V. Subramaniam, M. Savoia and C. Mazzotti

ii

Phenomenological and Experimental Study of Pultruded UD FRP


Composites Static and Deferred Rupture under Combined Flexural,
Compressive and Torsional Loadings ........................................................................ 203
N. Kotelnikova-Weiler and J.-F. Caron
Fatigue Performance CFRP Post-Tensioned Tendons ............................................. 213
A. El Refai, J. West and K. Soudki
Transverse Shear of GFRP Rods: Test Method Development and
Potential for Durability Assessment ........................................................................... 221
T.R. Gentry
Fatigue Life Behavior of Adhesively Bonded Pultruded FRP Joints under
Tensile and Compressive Loading .............................................................................. 229
R. Sarfaraz, A.P. Vassilopoulos and T. Keller
Fatigue Performance of Composite Sandwich Panels with and without
Internal Ribs ................................................................................................................. 237
H. Mathieson and A. Fam
HEMP Fiber Composites for Structural Retrofit ...................................................... 245
D. Asprone, M. Durante and A. Prota
Efficiency of Different FRP-Based Flexural Strengthening Techniques in
Beams Submitted to Fatigue Loading......................................................................... 253
J. Sena-Cruz, J. Barros, M. Coelho and L. Silva
A Reactive Multiphase Multicomponent Approach for Predicting the
Performance of FRP Composites in Concrete ........................................................... 263
M. Boulfiza
Validation of Digital Image Correlation Technique on CFRP and SFRP
wrapped Cylinders Subjected to Freeze-Thaw Environmental Exposure .............. 275
K. Abdelrahman and R. El-Hacha
Experimental Study of the Influence of Matrix System on Pultruded
GFRP Composites' Creep Behaviour under Flexural Loading .............................. 285
S. Barboura, N. Kotelnikova-Weiler, J.-F. Caron and O. Baverel
CFRP Confinement of Non-Seismic Columns under Explosive Loading ............... 299
A. Lloyd, M. Saatcioglu and T.K. Tikka

iii

Performance of Concrete Bridge Barriers Reinforced with Conditioned


GFRP Bars in Alkaline Solution ................................................................................. 309
E.A. Ahmed and B. Benmokrane
Thermal Endurance and Fatigue Performance of Prestressed Concrete
Prism Reinforcement in Civil Infrastructure ............................................................ 319
H.M. Vogel and D. Svecova
FRP Externally Bonded Systems for a Sustainable Strengthening of
Masonry Structures ...................................................................................................... 327
F. Ceroni, B. Ferracuti, M. Pecce and M. Savoia
Bridge Deck-Guardrail Anchorage Detailing for Sustainable
Construction .................................................................................................................. 337
K. Sennah, N. Nikravan, J. Louie, A. Hassan, N. Al-Bayati, M. El-Sayed and
M. Sayed-Ahmed
More than 10 Years Successful Field Applications of FRP bars in
Canada .......................................................................................................................... 345
B. Drouin, G. Latour and H.M. Mohamed
Mechanical Recovery of Epoxy Adhesives Subsequent to Exceeding Glass
Transition Temperature .............................................................................................. 357
O. Moussa, A. Vassilopoulous, J. De Castro and T. Keller
Use of Silane Adhesion Promoter to Enhance FRP-to-Steel Bond
Performance .................................................................................................................. 365
J. Sizemore, J. Aidoo, K. A. Harries and J. Monnell
Structural Health Monitoring of Corroded Steel Reinforcement in FRP
Repaired Beams Using Multiplexed Fibre Bragg Grating Sensors ........................ 375
R. Al-Hammoud, K. Soudki and T. Topper
Short Term and Long Term Properties of Newly Developped Bent GFRP
Reinforcing Bars ........................................................................................................... 383
A. Weber and C. Witt
Sustainability Problems Associated with the Heat and Fire Protection of
Fiber-Reinforced Polymers (FRP) ............................................................................. 391
H. Ibrahim and I. Mahfouz

iv

Effects of Aging on Glass Transition and Creep Behavior of Epoxy used


in FRP Strengthening ................................................................................................... 401
A. Jaipuriar, J. Flood, Y. Jeong, C.E. Bakis and M.M. Lopez
More than Twenty Years of Field Applications of Composites with
Durability Testing and Field Observations ............................................................... 409
E.R. Fyfe, R. Watson and M. McCullagh
Vehicle Crash Testing of a GFRP-Reinforced PL-3 Concrete Bridge
Barrier ........................................................................................................................... 417
K. Sennah, B. Juette, A. Weber and C. Witt
Experimental Setup and Thermal Insulation Assessment of Sandwich
Concrete Walls with GFRP Connectors for Sustainable Construction................... 425
G. Woltman, M. Hanna, D. G. Tomlinson and A. Fam
FRP Composites for Strengthening Bridges in Victoria ........................................... 433
Y.C. Koay
Ageing Effects on the Debonding Behavior of FRP Reinforcements:
Experiments and Damage Modeling ........................................................................... 443
K. Benzarti, F. Freddi, M. Quiertant and S. Chataigner
Numerical Simulation of Thermal Deformations in FRP-Reinforced
Concrete Members in Hot Region............................................................................... 453
A. Zaidi, R. Masmoudi and M. Bouhicha
Effects of the Inter-phase on the Properties of Glass Fibre Reinforced
Polymer (GFRP) Composites ..................................................................................... 461
A.S.M. Kamal, M. Boulfiza and S. Panigrahi
Sustainable Structures Using Hybrid-Composite Beams ........................................ 469
J. Hillman
Considerations for Design of FRP Retrofitted Reinforced Concrete Wall
Panels Subjected to Blast Loading .............................................................................. 477
E. Jacques and M. Saatcioglu
Mechanical Properties and Durability of Sustainable FRP Composites ................ 485
P.V. Vijay, H.V.S. GangaRao and M.P.K. Reddy

Durability Study of Basalt Fiber Reinforced Epoxy Wet Layups Used in


Rehabilitation ................................................................................................................ 493
G. Xian, C. Wang, B. Xiao and H. Li
Comparative Durability Analysis of CFRP Strengthened RC Highway
Bridges .......................................................................................................................... 501
O. Ali, D. Bigaud and E. Ferrier
Axial Behaviour of Concrete Filled FRP Tube Columns after 300-FreezeThaw Cycles .................................................................................................................. 513
H. El-Zefzafy, H.M. Mohamed and R. Masmoudi
CFRP-Strengthened Beams under Repeated Impact Loading ............................... 523
A. Parvin and T.A. Mohammed
Bond Stress-Slip Model for FRP Rebars in Concrete ............................................... 531
S. Quayyum and A. Rteil
A Case Study on the Use of Advanced Fiber Wrap Composites for
Reinforced Concrete Repair of Port Terminal Wharf Caissons .............................. 539
J.C. Percival and T.T. Jimenez
Probabilistic Development of a Life Cycle Inventory (LCI) Dataset for
Pultruded Fiber Reinforced Polymer (FRP) Composites ......................................... 547
S.M. Ali, M.D. Lepech and J.P. Basbagill
Study on Moisture Fickian Diffusion Process of a Pultruded FRP
Composite Material ..................................................................................................... 555
X. Jiang, H. Kolstein, and F.S.K. Bijlaard
Durability of FRP-Strengthened RC Members under Various Loadings ............. 563
A. Parvin and A. Kulikowski
Fatigue Performance of Hybrid SFRP-GFRP-UHPC Beams ................................. 575
D. Chen and R. El-Hacha
SDOF Analysis of Protective Hardening Design for Reinforced Concrete
Columns using a Fiber Reinforced Polymer Wrap ................................................... 585
J. Quek and M.C. Ow
Effect of Mechanical Load on the Thermal Behavior of Concrete Slab
Reinforced with Fibre Reinforced Polymer (FRP) Bars .......................................... 593
H. Bellakehal, R. Masmoudi, A. Zaidi and M. Bouhicha
vi

Proposed Shear Model for Reinforced Concrete Beams Wrapped with


FRP: A Genetic Algorithm Approach ........................................................................ 603
C. Marshall, A. Rteil and M.S. Alam
Long-Term Durability of FRP Reinforcements Damaged by High
Temperature ................................................................................................................ 611
D.Y. Moon, H. Oh and J. Sim
Flexural Behavior of Simple and Fixed-end Beams Strengthened with
FRP bars in NSM Method .......................................................................................... 619
M.K. Sharbatdar, M. Mohamadian and S.M. Jaberi
Fatigue Performance RC Beams Strengthened with Prestressed NSMStrips and Rods ............................................................................................................ 627
F. Oudah and R. El-Hacha
Concrete Demolition of Bridge Barriers and Deck Slabs Reinforced with
GFRP Bars ................................................................................................................... 637
E.A. Ahmed, C. Dulude, S. Goulet and B. Benmokrane
Effect of Freeze-Thaw Cycles on the Compression Strength of Concrete
Filled FRP Tube Cylinders .......................................................................................... 649
H. El-Zefzafy, H.M. Mohamed and R. Masmoudi
Effects of Harsh Laboratory and Field Environmental Conditions on the
Durability of Concrete Wrapped GFRP Bars .......................................................... 659
T.H. Almusallam, Y.A. Al-Salloum, S. Alsayed, S.E. El-Gamal and M. Aqel
Durability of Concrete-Filled FRP Tubes Subjected to Salt Solution ..................... 667
M. Robert, A. Fam and B. Benmokrane
Freeze-Thaw Cycling Effect on the Mechanical Properties of FRPFilament-Wound Tubes .............................................................................................. 677
H. El-Zefzafy, H.M. Mohamed and R. Masmoudi

vii

viii

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

ALTERNATIVE ENVIRONMENTAL KNOCKDOWN FACTORS


(CE) DERIVED FROM EXTENSIVE EXPERIMENTAL DATA
K.A. Harries 1 and B.M. Shahrooz 2
1
2

Associate Professor, Dept of Civil and Env. Engrg, Univ. of Pittsburgh, Pittsburgh PA, USA
Professor, School of Advanced Structures, Univ. of Cincinnati, Cincinnati OH, USA

ABSTRACT
An extensive study investigating the performance of fiber reinforced polymer (FRP)
materials and FRP bond to concrete encompassing 64 permutations of FRP material, test
method, and environmental conditioning was conducted. Based on the findings of this
study, alternative environmental exposure factors - the so-called CE knockdown factors are proposed. Factors associated with material properties remain dependent on the FRP
material with values of 0.90 and 0.80 proposed for CFRP and GFRP, respectively. It is also
proposed that these factors be applied to both FRP strength and modulus values only. Strain
capacity will, therefore, remain unchanged. In addition, factors to be applied to bond
capacity are proposed; these are dependent on expected material quality and are proposed
as 0.90 for preformed CFRP and 0.50 for fabric systems regardless of material.
1. INTRODUCTION
When designing with FRP materials, it is important to distinguish between material
resistance factors, load reduction factors, and environmental reduction factors. This paper
discusses only the latter so-called knockdown factors. These factors are given the notation
CE in ACI 440.2R [1]; this notation is adopted here. The ACI-prescribed values of CE are
given in Table 1. CE factors are not intended to capture effects of sustained such as creep
and fatigue; these effects are typically addressed using FRP strain limits, which are beyond
the scope of the present study.
Conventionally, the CE factor is applied to FRP strength (i.e., CEFu) and strain capacity
(CEu) but not modulus [1]. The results of the present study call into question this
convention, and demonstrate that strength and modulus are affected while strain capacity is
not significantly affected by environmental conditioning. Additionally, the CE factor is
typically only applied to material properties and not when determining bond capacity. This
study shows that environmental exposure has a greater effect on bond behaviour than on
FRP strength or modulus. Finally, this study also indicates a difference in performance
between preformed and hand laid-up FRP with the preformed FRP demonstrating superior
1

durability. This difference between manufactured and hand lay-up materials reflects issues
of quality control and should not be surprising. Nonetheless, existing guidelines [1] do not
acknowledge this difference, effectively penalizing the use of preformed materials.
Table 1. CE values prescribed by ACI 440.2R-08 [1].
Exposure
CFRP Plate CFRP Fabric
Interior Exposure
0.95
Exterior Exposure
0.85
Aggressive Environment Exposure
0.85

GFRP Fabric
0.75
0.65
0.50

The objective of this work is to establish a basis for recommending environmental exposure
factors suitable for design of externally bonded FRP materials. This goal is achieved based
on the results of a database having 64 permutations of FRP material, test method, and
environmental conditioning [2,3] described in the following sections.
2. EXPERIMENTAL PROGRAM
An extensive experimental program investigating the behaviour of three FRP systems
subjected to nine different environmental conditioning protocols was conducted [2,3]. The
effect of environmental conditioning was assessed using four different standard test
methods. Table 2 summarizes the test matrix for this program.
2.1 Environmental Conditioning Considered
Eight different environmental conditioning protocols, in addition to control specimens,
were considered. The ambient control conditions averaged 22oC and 70% RH and provide
the baseline against which results from conditioned tests are compared. The following
environmental conditioning protocols were considered in this study; additional details are
found in references [2] and [3].
Water Exposure: Test specimens were exposed to conditions of 100% humidity at 38oC for
durations of 1000, 3000, and 10,000 hours in accordance with ASTM D2247.
Salt Water Exposure: Test specimens were immersed in substitute ocean water, prepared
according to ASTM D1141, at 22oC for durations of 1000, 3000, and 10,000 hours.
Alkaline Exposure: Test specimens were immersed in a saturated solution of calcium
carbonate (CaCO3) at 22oC for durations of 1000, 3000, and 10,000 hours. During
exposure, the solution was maintained at a pH of 9.5.
Dry Heat Exposure: Test specimens were exposed at 60oC in a forced-draft circulation-air
furnace in accordance with ASTM D3045 for durations of 1000 and 3000 hours.
Diesel Fuel Exposure: Test specimens were immersed in diesel fuel at 22oC for 4 hours, in
accordance with ASTM C581.

Exposure
Baseline
Water
Salt Water
Alkaline
Dry Heat
Diesel
Weathering
Freeze-Heat

Table 2. Durability test program [2,3].


Duration
Description
laboratory
1000, 3000 &
100% RH @ 38oC
10000h
1000, 3000 &
Immersion in salt water @ 22oC
10000h
1000, 3000 &
Immersion in pH 9.5 CaCO3
10000h
solution @ 22oC
1000 & 3000h
60oC
4h
Immersion in diesel fuel @ 220C
1000 cycles
2h of UV @ 63oC
2h @ 100% RH
(4000h)
20 cycles
9h @ -18oC
15h @ 100% RH @ 38oC
(480h)

Baseline
Freeze-Thaw

laboratory
360 cycles
(1583h)

70min @ -18oC
70 min @ 4.5oC and UV

Tests Conducted
3 test methods:
ASTM D3039
tension,
ASTM D2344
SBS & bond test
4 specimens of:
CFRP plate,
CFRP fabric &
GFRP fabric
2 beam flexure
tests of: unretrofit
control, CFRP
plate, CFRP fabric
& GFRP fabric

Weathering Exposure: Specimens were exposed to UV light from a carbon arc source at a
temperature of 63C for 2 hours (ASTM G23) followed by exposure to conditions of 100%
humidity at 22oC for 2 hours. This procedure was repeated for 1000 cycles.
Freeze-Heat Exposure: Test specimens were initially placed in conditions of 100%
humidity at 38oC for 500 hours to allow for moisture absorption prior to freezing. Upon
removal from the humidity chamber, specimens were air dried for 48 hours and then placed
in a freezer at -18oC for 9 hours. Following this step, the specimens were placed back into
the humidity chamber at 100% humidity and 38oC for 15 hours. This freeze-thaw cycling
was repeated a total of 20 cycles.
Freeze-Thaw Exposure: Only reinforced concrete flexural beam specimens having bonded
FRP were subjected to this exposure. Specimens were subject to freeze-thaw conditioning
based on the ASTM C666 protocol. The test protocol extended for 360 cycles, each
consisting of the following four steps: 1) 70 minutes at -18C at 30% RH; 2) 20-minute
ramp up to 4.5C (resulting in 90% RH); 3) 70 minutes at 4.5 C at 50% RH with UV lights
on; and 4) UV lights off and 80-minute ramp down to -18C (resulting in 40% RH).
2.2 Test Methods
Four test methods were used to assess the effects of environmental conditioning. The test
methods and specimen preparation are described as follows; additional details are found in
references [2] and [3].

Tension Test: Tension capacity parallel to the fiber was determined using standard ASTM
D3039 coupons and test method. Baseline and conditioned specimens were cut using a
water-cooled diamond blade into coupons 254 mm long by 25.4 mm wide having 50 mm
long fiberglass gripping tabs at each end. Tension tests on the baseline specimens also
provided fundamental material properties reported in Table 3.
Short Beam Shear Test: The short beam shear (SBS) test provides interlaminar shear
strength (ILSS). This property is not applicable to single ply fabric specimens; therefore,
SBS tests were only conducted on the preformed CFRP plate material. The test was carried
out according to the method of ASTM D2344. Specimens were cut using a water-cooled
diamond blade into coupons 19 mm long by 6.4 mm wide. The specimens were loaded in
three-point flexure over a span of 12.7 mm.
Bond to Concrete: The bond test specimen consisted of two 51 mm concrete cubes spaced
25 mm apart bonded together using 19 mm wide by 127 mm long FRP strips on opposing
faces. The bonded length of FRP on both cubes was 38 mm. The average 56-day
compressive strength for all 192 cubes was 30 MPa. Prior to FRP application, the concrete
was air-dried for 96 hours and a wire brush was used to remove all laitance from the faces
to be bonded. The CFRP plate was bonded using a compatible structural epoxy adhesive.
The fabric systems were bonded during the layup process using the saturating resin.
Custom-made U-shaped collars were used to grip opposing cubes in a universal test
machine and tension was applied in displacement control at a rate of 2.2 mm/min.
Beam Flexure: Sixteen concrete beams 154 mm deep by 203 mm wide by 2440 mm long
reinforced internally with two #3 (9.5 mm) bars, top and bottom, and U shaped W2.9 (4.9
mm) deformed wire stirrups at 152 mm centers were used. The concrete compressive
strength determined at the age of testing was 38 MPa. Beam surfaces to receive FRP were
sandblasted to remove laitance. The CFRP plate was bonded using a compatible structural
epoxy adhesive. The fabric systems were bonded during the layup process using the
saturating resin. A single 2130 mm long layer of FRP was applied to each beam: for the
CFRP plate, a 102 mm wide plate was bonded to the beam soffit; the CFRP fabric repair
was 152 mm wide and the GFRP fabric was 305 mm wide resulting in it extending 51 mm
up each side of the beam. The FRP details are not comparable in terms of strengthening
effect (nor were they intended to be); the objective of this study was only to draw
comparisons between similar beams subjected to different environmental conditioning.
Beams were tested in three-point flexure over a span of 2288 mm.
2.3 FRP Material Systems
Three commercially available FRP systems were included in this study: 1) a preformed
unidirectional CFRP plate; 2) a unidirectional carbon fiber fabric; and 3) a unidirectional
glass fiber fabric. The carbon and glass fabrics were hand laid-up into their CFRP and
GFRP forms using a compatible two-part saturating resin. When bonded to concrete, a
layer of resin was used to prime the concrete surface prior to adding the saturated fabric
sheets. The CFRP plate was bonded to the concrete using a compatible two-part structural
epoxy. Experimentally determined [3] material properties of all three systems are given in
Table 3. Throughout this study, FRP modulus and strength are given in terms of force per
4

unit width of FRP (N/mm). This approach is consistent with that now promulgated by
ASTM D7565 and normalizes for FRP thickness which varies for a hand laid-up fabric
system. In essence, the values reported are E x t and Fu x t, respectively.
Table 3. Material properties of FRP system components.
ASTM
CFRP
CFRP
GFRP
property
units
test
plate
fabric
fabric
material thickness
mm
1.14
varies with layup
D3039 N/mm
tensile strength, Fu x t
2550
650
330
D3039 kN/mm
modulus of elasticity, E x t
142
58
18
D3039
%
1.88
1.06
1.87
ultimate strain, u
o
-6 o
CTE @ 20 C
E831
10 / C
1.48
3.94
13.77
o
glass transition temperature, Tg D4065
C
150
52
45
3. RESULTS OF EXPERIMENTAL PROGRAM
A summary of material properties and performance of all specimens subject to all
environmental conditioning is provided in Table 4. Reported values are normalized by
those obtained from the baseline tests. Detailed results are presented in [3].
Table 4. Average measured properties normalized to those obtained for baseline exposure.

Freeze-Thawe

Freeze-Heat

Weathering

Diesel

Heat 3000h

Heat 1000h

Alkaline 10000h

Alkaline 3000h

Alkaline 1000h

Salt Water 10000h

Salt Water 3000h

Salt Water 1000h

Water 10000h

Water 3000h

Water 1000h

Exposure

GFRP fabric CFRP fabric

CFRP Plate

E x ta 0.96 0.97 1.06 1.01 1.00 1.03 1.02 0.96 1.03 1.01 1.05 1.02 0.99 0.99
a
Fu x t 1.05 1.09 1.15 1.02 1.07 1.12 1.02 1.04 1.13 1.06 1.06 1.03 1.02 1.03
ua
1.02 1.04 1.09 0.84 1.02 1.07 0.97 0.95 1.02 1.01 1.00 0.96 1.02 0.97
FSBSb
1.01 1.02 0.98 1.05 1.03 0.98 1.01 1.00 1.01 1.04 1.07 1.03 0.94 0.97
bondc 1.07 1.14 1.18 1.17 1.30 0.94 1.16 1.18 0.98 0.91 0.98 1.00
1.15
flexd
0.94
E x ta 1.12 0.98 1.05 1.10 0.98 1.14 1.09 1.02 0.90 0.98 1.10 1.00 1.03 1.12
Fu x ta 1.26 0.92 1.08 1.22 0.97 0.97 1.22 0.97 0.97 1.17 1.18 0.92 1.20 1.14
ua
1.14 1.01 1.13 1.22 0.99 0.84 1.17 1.06 1.09 1.06 1.16 1.13 1.30 1.19
bondc 0.91 1.26 1.16 1.29 1.29 1.13 1.13 1.40 1.44 0.61 0.60 1.00
1.28
flexd
1.02
E x ta 1.00 1.00 1.00 1.06 0.94 0.94 1.00 1.00 1.00 1.06 1.00 1.00 0.94 1.00
Fu x ta 0.91 0.82 0.97 1.06 0.88 0.94 0.85 0.85 0.91 0.88 0.94 0.94 0.88 1.00
ua
0.94 0.83 1.01 0.96 0.89 1.03 0.81 0.96 1.07 0.79 0.87 1.01 0.90 0.96
bondc 1.09 0.87 1.01 1.10 1.11 0.87 0.90 0.83 0.52 1.05 1.01 1.06
1.15
flexd
0.97
lowest observed values are presented in bold text; all values are average of four tests except freeze-thaw (2 tests)
a
obtained from D3039 tension test; b obtained from D2343 short beam shear test; c obtained from bond test
d
obtained from flexural beam test; e unretrofit control beams = 0.92

3.1 CFRP Plate


The preformed CFRP plate material was relatively unaffected by the environmental
conditioning conducted. As a performed product, the CFRP plate system has better
consistency and quality and has cured since its manufacture. Additionally, a resin-rich outer
layer protects the underlying carbon tows from exposure. Of note is the consistent
improvement in modulus and strength with increased exposure to water at 38oC. This
observation is likely indicative of the elevated temperature resulting in additional post-cure.
Due to the relatively low CTE of the CFRP plate, few potentially damage-causing residual
stresses are likely to develop due to elevated post-cure temperature; therefore, no
detrimental effects are observed. These results, and the relatively inert behaviour when
exposed to salt water, demonstrate that the resin-rich outer layer of the CFRP plate inhibits
water absorption and is resistant to combined exposure (weathering). In no event does the
modulus or strength of the CFRP plate fall below 95% of the baseline value.
3.2 CFRP Fabric
The hand-laid up CFRP fabric also performed relatively well under the environmental
conditioning conducted. The greater variation compared to the CFRP plate can be partially
attributed to variation introduced in a hand lay-up process. Some increase in system
performance is evident in the conditioning methods that involved elevated temperature; this
increase may be attributed the post-cure effects of elevated temperature. However, there is
also evidence of water absorption that negatively affects the system performance. For
instance, the tension properties of the CFRP fabric are improved following 1000h exposure
to 38oC water (attributed to post-cure); however, the performance degrades at longer
exposures indicating eventual absorption. In this case, the elevated post-cure temperature
may result in some microcracking associated with temperature-induced residual stress
allowing absorption to take place. The additional detrimental effects of salt are also
apparent in the behaviour of the CFRP fabric. Nonetheless, the tensile modulus and
strength do not fall below 90% of the baseline values under any of the exposures
considered.
3.3 GFRP Fabric
The GFRP fabric was less durable than the CFRP products. No post-cure resulting from
elevated temperature was evident, and the detrimental effects of absorption are readily
apparent. As may be expected, the GFRP was particularly affected by exposure to an
alkaline environment. Glass fiber in GFRP applications is known to degrade in an alkaline
environment; for this reason the integrity of the surrounding resin matrix is critical to good
performance. A relatively high CTE can lead to matrix cracking allowing absorption of
concrete pore water with its incumbent alkalinity. The GFRP also indicated deterioration
under both elevated temperature environments. This decline may be an indication that these
environments exposed the specimens to temperatures greater than their glass transition
temperature, Tg. If this is the case, deterioration is expected to be due to the phase change in
the polymer matrix. This effect was not seen in the CFRP fabric specimens that did have a
marginally greater value of Tg.

3.4 Glass Transition Temperature and Water Absorption


Single specimens of all three FRP materials from each conditioning regime were tested to
assess any shift in Tg [3]. In no case was a statistically significant change in Tg recorded.
However, in those specimens where absorption was suspected, distinctive peaks at 0o and
100oC were noted, corresponding to the freezing and boiling points (phase changes) of the
absorbed water. This confirmed the hypothesis that absorption was present in both FRP
fabrics although not in the CFRP plate material.
3.5 Bond to Concrete
Bond to concrete is an interface property and not itself affected by the deterioration of the
FRP material. Certainly, the FRP may deteriorate to a degree where failure occurs prior to
bond strength being developed but this decline does not imply bond failure per se. The
majority of the reported deterioration of bond capacity associated with environmental
conditioning was not bond related but simply reflects the deterioration and failure of the
FRP material. The results of the GFRP bond specimens subjected to alkaline environment,
for instance, exhibit the same degree of deterioration as the comparable tension tests. Some
adhesive failures were noted at the FRP or concrete interfaces of the bond line. These
failures may be attributed to environmental induced degradation of the adhesive layer.
Deteriorated bond capacity of the CFRP plate system subjected to heating conditions
(60oC) may be attributed to the elevated temperature approaching the Tg value of the
adhesive epoxy used. While this value was not determined in the present study, a typical
value would be on the order of 50oC.
3.6 In situ Performance Beam Flexure Tests
The degree of strengthening provided varied considerably. The CFRP plate, CFRP fabric
and GFRP fabric-retrofit beams had capacities 218%, 157% and 149% of their companion
unretrofit beam capacities, respectively. The effect of 360 freeze-thaw cycles on the
unretrofit concrete beam was to reduce its capacity approximately 8%. This reduction is
consistent with results expected from good quality concrete and indicates little damage to
the concrete resulting from the freeze-thaw conditioning. The reductions associated with
freeze-thaw exposure of the CFRP plate and GFRP fabric retrofit beams are only 6% and
4%, respectively. The beams having CFRP fabric subjected to freeze-thaw exposure had no
apparent strength loss. The reduced degree of environmental-induced degradation
associated with the presence of FRP may result from the FRP actually protecting the
concrete from freeze-thaw damage. With the FRP in place, a large region of soffit concrete
is protected from water ingress and, therefore, the effects of freeze-thaw cycles are
diminished.
Primary failure of the retrofit beams was associated with loss of the FRP strengthening
effect. An evaluation of failure modes in this regard can reveal information on the FRP
systems resistance to the combined freeze-thaw and UV exposure provided. The CFRP
plate system exhibited typical intermediate crack-induced (IC) debonding [4] in all cases.
This observation is consistent with the expected mode of flexural failure and was
7

apparently unaffected by the freeze-thaw conditioning. The behaviour of the CFRP fabricreinforced beams changed from one controlled by adhesive failure of the bond between
FRP and concrete to IC debonding for the conditioned beams. This shift is counterintuitive
and may reflect poor quality lay-up or surface preparation of the unconditioned specimens.
Nonetheless, the fact that the conditioned beams performed admirably is an indication that
little freeze-thaw induced deterioration occurred.
The GFRP fabric retrofit beams had such a small effective increase in reinforcement
attributable to the GFRP that the fabric ruptured prior to debonding. In all cases, the FRP
rupture resembled a brooming type of failure typical of tension specimens. Such a failure
is indicative of a good quality FRP having a sound matrix and good fiber continuity [5].
While the failures of the GFRP fabric were similar, the extent of the rupture (brooming)
area was reduced for the conditioned beam specimens indicating that there may be some
degradation of the GFRP system present.
4. RECOMMENDATIONS
Based on the extensive experimental program described, the following environmental
exposure factors for design of externally bonded FRP materials are proposed (Table 5).
Environmental exposure factors associated with material properties are dependent on the
FRP material with values of 0.90 and 0.80 proposed for CFRP and GFRP, respectively. It is
proposed that these factors be applied to both FRP strength (i.e., CEFu) and modulus (CEE)
values. Since the linearity of FRP behaviour is unaffected, strain capacity remains
unchanged since the combinatorial effects of reducing both the stress and modulus equally
result in unchanged strain (i.e., u = CEFu/CEE). This application to strength and modulus,
but not strain, is supported by the present study.
In addition, exposure factors to be applied to bond capacity are proposed (Table 5); these
factors are dependent on expected material quality and are proposed as 0.90 for preformed
CFRP and 0.50 for fabric systems regardless of material. Based on the conditioning
performed in this study, it is deemed that these proposed factors are appropriate for exterior
exposure. The factors can likely be relaxed for interior exposure while environmentspecific factors need to be developed for specified aggressive environments.
Table 5. Reduction factors for environmental exposure.
CFRP
CFRP
GFRP
Plate
Fabric
Fabric
Lowest values observed in this study
0.96
0.90
0.94
Tension Modulus, E x t
1.00
0.92
0.82
Tension Strength, Fu x t
Bond capacity
0.91
0.60
0.52
Recommendations
FRP Material Properties
0.90
0.80
Bond Capacity (to concrete substrate)
0.90
0.50

5. REFERENCES
1.
2.
3.
4.
5.

ACI Committee 440. 2008. Guide for the Design and Construction of Externally
Bonded FRP Systems for Strengthening Concrete Structures. ACI 440.2R-08.
American Concrete Institute, Farmington Hills, MI, 76 p.
Cromwell, J.R., Shahrooz, B.M. and Harries, K.A. 2011. Environmental Durability of
Externally Bonded FRP Materials Intended for Repair of Concrete Structures. ASCE
Journal of Composites for Construction (in review).
Pack, J.R. 2003. Environmental Durability Evaluation of Externally Bonded
Composites. MSCE thesis, Department of Civil and Environmental Engineering,
University of Cincinnati, 289 p.
Teng, J.G. Chen, G.F., Smith, S.T. and Lam, L. 2002. FRP-Strengthened RC
Structures. John Wiley & Sons, New York, 245 p.
Agarwal, B.D. and Broutman, L.J. 1990. Analysis and Performance of Fiber
Composites, 2nd edition. John Wiley & Sons, New York, 449 p.

10

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

A FIELD STUDY ON DURABILITY OF BOND STRENGTH


BETWEEN CFRP WRAPS AND CONCRETE
S.A. Sheikh1, A.D. Caspary1 and M. Stafford2
1
2

Department of Civil Engineering, University of Toronto, Toronto, Canada


R. J. Watson, Inc., East Elmhurst, NY, U.S.A.

ABSTRACT
The use of composite materials, in particular fibre-reinforced polymers (FRP), for
infrastructure strengthening, retrofit and rehabilitation applications has been steadily
increasing in the field. The advantages of FRP as a retrofit material include among others
ease of application, high specific stiffness and strength ratios, and resistance to
electromagnetic corrosion.
The failure modes of externally wrapped FRP include tensile rupture, as well as delamination and/or peeling of FRP material due to shear distortions or cracking. The delamination and peeling modes are directly related to bond properties between the FRP and
concrete.
An investigation to determine the in-situ bond strength between FRP wraps and concrete
exposed to actual field conditions was performed by the University of Toronto. The paper
summarizes the direct tension pull-off tests performed on the two CFRP retrofitted
structures, one a highway bridge repaired in 1996 and the other a 14-storey residential
building retrofitted in 1999. Both structures are located in Toronto. Only the columns in the
bridge were repaired while in the building structures, the beams, columns and walls were
rehabilitated. The repair of both the structures was carried out using external wet-lay CFRP
wraps. The results indicated a variation of bond strength in the range of 2.26 MPa to 3.92
MPa. In a number of field specimens the failure was not achieved due to the limited
capacity of the equipment used. The failure modes included separation of pull-off plates
from the FRP wraps, bond failure at the interface between FRP and concrete, and tensile
failure of concrete.
1. INTRODUCTION
The use of composite materials, in particular fibre-reinforced polymers (FRP), for
infrastructure strengthening, retrofit and rehabilitation applications has been a widely
studied topic in recent years. The advantages of FRP as a retrofit material include high

11

specific stiffness and strength ratios, outstanding fatigue behaviour, resistance to


electromagnetic corrosion and ease of application.
Two of the most common failure modes in concrete structures externally wrapped with
FRP are material tensile rupture and delamination and/or peeling of FRP wrap. The
delamination and peeling modes are directly related to bond characteristics between the
FRP and concrete. Performance of retrofitted structures thus significantly depends on the
bond properties between the FRP sheets and the concrete. As such, the bond strength and
the methods used to determine it are critical to the understanding and evaluation of FRP
retrofitting techniques, as well as the FRP-concrete behaviour and failure mechanism.
While laboratory studies can provide useful information, they can rarely replicate the actual
conditions experienced by a structure in the field. Thus, there exists a critical need for data
from field applications and tests to determine reliable estimates of the bond strength
between concrete and FRP.
Two structures upgraded with FRP in 1990s were used to evaluate the FRP-concrete bond
properties after years of service in the field. The Leslie Street Bridge, built in the sixties, is
located in Toronto, Ontario. The highway has 7 westbound and 6 eastbound lanes at this
location. The columns are about 920 mm to 1010 mm in diameter. The specified concrete
strength was 20.7 MPa (3000 psi). The Amadiyya Abode of Peace, a 14-storey residential
building, is also located in Toronto, Ontario. The exterior west and east faces of the
building each contain 14 beam-column supports. The columns are approximately 830 mm
in diameter, and the beams are approximately 600 mm wide and 1200 mm deep.
2. REPAIR WITH FRP
After decades of overuse and neglect, our aging infrastructure is failing, particularly in
areas that are exposed to years of harsh weathering elements, resulting in corrosion of the
reinforcing steel embedded within concrete structures. Figures 1 and 2, respectively, show
examples of severely corroded bridge columns and exterior building columns before and
after the FRP retrofit (Sheikh et al. 2002, Sheikh 2007). The concrete cover is completely
lost in bridge columns while in the building structure, the damage due to steel corrosion is
still limited with concrete severely cracked. Neither of these conditions is uncommon.
Engineers have been searching for alternative materials to repair and prevent these types of
structural issues and in the past 15 years, fibre reinforced polymers (FRPs) have emerged as
innovative materials that can be useful in many infrastructure rehabilitation/ strengthening
projects.
One highly successful method for infrastructure rehabilitation involves wrapping concrete
columns and beams with FRPs. The confinement provided by FRP significantly increases
the strength and deformation capacity of columns. In the case of beams, the transverse fibre
wrap significantly increases the shear strength. This technique is now widely recognized as
an effective and efficient strengthening, repair, and rehabilitation method. Twenty eight
bents consisting of beam and columns were repaired on the west and east exterior faces of
the apartment building. The current conditions of the repaired columns in Figures 1 and 2,
after exposure to extreme environmental conditions in and around Toronto for up to 15
years, show remarkable performance of FRP repair. The repaired structures showed
12

excellent resistance to freeze-thaw cycles, UV radiation, low temperatures, temperature


cycling, chemicals (alkalis and chlorides), ice, rain, sustained load, impact and fatigue.

Before Repair in 1996

2010

Fig. 1. Severely corroded bridge columns before and after repair

Before Repair

Repair in 1999

2010

Fig. 2. Corroded and damaged building beams and columns and FRP repair
Although, FRP repair has been used for all types of strcutural components, its use in
columns has been uniquely advantageous. Closed FRP wrap provides shear and
confinement reinforcement but the bond between FRP and concrete is not critical in most
such applications. Test data from columns in the field retrofitted with FRP, however,
provide valuable information about the durability and longevity of FRP repair as far as the
bond between concrete and FRP is concerned.
3. PULL-OFF TESTS FOR FRP-CONCRETE BOND
As the first step in the preparation for pull-off tests, the protective paint and barriers are
carefully removed from the surface of FRP which has previously been bonded to the
concrete substrate. The FRP surface is then cleaned in accordance with the requirements of
the manufacturers of the epoxy used for adhering the metal disc to the FRP surface. FRP
areas equal to the size of the disc is cut through to the concrete layer with the help of a core
drill before the discs are attached. After the epoxy has sufficiently cured, a testing apparatus

13

(Figure 3) is used to pull off the discs and measure the value of tensile force required to debond the metal disc and all adhered material from the concrete surface. The disc diameter
of the commercially available system varies between 20 mm and 50 mm. The epoxy is
specified to have a tensile capacity of at least 5.0 MPa. The current test series at the bridge
was part of a program to evaluate FRP-concrete bond at four locations across Canada for
which it was decided to use 50 mm diameter discs for bridge tests (Banthia et al. 2010).
The larger disc was considered to provide a more representative value for bond strength.
For the building, two disc sizes (50 mm and 20 mm) were used to study the effect of disc
size on bond strength. The ideal failure mode is described as cohesive failure within the
concrete at a stress greater than 1.4 MPa, according to available codes and standards (ACI,
2004; ASTM, 1993; CSA, 2004).
Observations from this test include the force required to de-bond the specimen; the location
of the plane of failure and a qualitative assessment of the bond. The location of failure
plane is generally denoted as E (Adhesive epoxy failure premature failure), E/F
(Adhesive epoxy failure combined with cohesive FRP failure within the laminate
premature failure), F (Adhesive failure of FRP concrete interface, or total cohesive failure
within the FRP laminate), F/C % (FRP-Concrete interface failure given as the percentage of
test area retaining concrete/aggregate), and C (Cohesive failure within the concrete). See
Figure 3 for test set-up and location of potential failure planes.

Fig. 3. Pull-Off Test Set-up and Potential Failure Locations


In most applications, load is generally transferred between FRP and concrete through a
combination of shear and tension stresses, particularly in flexural members (Wan et al.,
2002) while the interface between FRP and concrete is subjected to direct tensile stress in
this test. Although, this test may not simulate stress conditions for flexural members, it
provides a useful indication for structural performance. This test is simple and a valuable
tool for investigating qualitative and relative values for different parameters.
3.1 Ahmadiyya Abode of Peace Building
One Beam and one Column (north and south Face) on the south end of the west side of the
structure were selected for tests. For the column tests, the locations of the discs were
selected randomly as the FRP was used to provide confinement. Since the beams were
retrofitted for shear, the discs were specifically placed near the neutral-axis (mid-height),
where shear forces would be maximum. Before the discs were adhered to the members, the
FRP surface was thoroughly cleaned and FRP areas equal to the size of the disc were cut
through to the concrete layer with the help of a core drill. Nine 50mm-disc and eight
20mm-disc tests were installed.

14

3.2 Leslie Street Bridge on Highway 401


Tests were carried out on two columns which were wrapped with CFRP to halt corrosion
due to excessive use of de-icing salts during the winter season. The retrofitted columns
divide the north bound lanes of Leslie Street from the south bound lanes and receive
extensive exposure to ice and salt as a result of heavy traffic (Figure 4). The locations for
discs were selected to represent this exposure. The core marking on the column were
made on two grids; one facing the traffic with a 30 degrees angle to the axis of the
roadandtheotherontheoppositesideofthecolumn. The FRP surface was prepared in
the same manner as described above. A total of 16 tests were completed.Figure 4 shows a
spot at the lower end of the column where FRP has been damaged. This was caused by a
direct hit by a vehicle. Obviously, the test locations were selected at a distance away from
this damaged area. Figure 5 shows two different variations of the commercially available
testing apparatus.

Fig. 4. Leslie Street Bridge, Toronto

Fig. 5. Testing Apparatus and Procedure

15

4. RESULTS
4.1 Observations and Conditions Status
The Leslie Street Bridge pull-off tests were performed in the second week of August in
2009. The temperature ranged between 12oC and 30oC and the relative humidity was 52%
to 99%. FRP surface at most locations on the columns appeared intact. The pull-off tests on
the apartment building beams and columns were carried out in the second week of
November 2009. The temperature during that time in Toronto varied between 2oC and 12oC
and the relative humidity varied between 69% and 82%. In this structrue, the FRP appeared
in excellent condition and much better than the FRP on bridge columns.
4.2 Pull-off Test Values
The maximum pull-off load value recorded during each test was used to calculate the pulloff bond strength by diving it by the disc area of 1964 mm2. Tables 1 and 2 show average
pull-off strength and stadard deviation for all the tests conducted at the Ahmadiyya
Building and Leslie Bridge, respectively. The capacity of the pull-off tester using 50 mm
discs was 3.5 MPa. In several tests, the discs could not be pulled off and the maximum
strength recorded was 3.5 MPa. The minimum pull-off strength observed in bridge columns
was 2.0 MPa. Since only 50 mm discs were used at the bridge, the pull-off strength for the
FRP is thus underestimated due to equipment limitation. Although the same limitation was
experienced for the tests on beams and columns in the building, the average strength from
50 mm and 20 mm discs are very similar. The maximum pull-off strength with 20 mm discs
was 3.92 MPa while the minimum recorded was 1.5 MPa. Figure 6 shows a typical failures
that can be classified as F/C (FRP-Concrete interface failure given as the percentage of test
area retaining concrete/aggregate) and C (Cohesive failure within the concrete). The failure
modes observed at the building include 5 F, 4 C, 4 F/C, and 2 E (premature). Two tests
were terminated due to the capcity limit of the tester. At the bridge the failure modes are as
follows: 4 F, 4 C and 2 F/C. Six tests were terminated due to the capcity limit of the tester.
It should be noted that the FRP repair of the bridge has endured the environmental exposure
for about 15 years under service conditions. The age of repair in the case of the building
structure is 12 years. The average bond strength ranging from 2.6 MPa and 3.1 MPa after
12 to 15 years of service in extreme enviornmental conditions indicates good performance
of FRP repair. Much lower pull-off strength has been reported in some bridges (Banthia
2010) but it was observed that factors such as placement characteristics, surface finish,
infiltration of water behind the FRP layer and other workmanship issues played important
roles in determining the long-term performance of the repair.
Table 1. Average Pull-Off Strength Values at Leslie Street Bridge
Average Pull-Off Strength Standard Deviation (SD)
MPa
MPa
Column 1
3.06
0.63
Column 2
3.14
0.57
All Tests
3.10
0.58

16

Table 2. Average Pull-Off Strength Values at Ahmadiyya Building


Average Pull-Off Strength
Standard Deviation (SD)
MPa
MPa
Column (50mm)
2.78
0.95
Beam (50mm)
2.74
1.26
Column/Beam (50mm)
2.76
1.03
Column (20mm)

2.45

1.00

Beam (20mm)
Column/Beam (20mm)
All Tests

2.72
2.63
2.71

1.27
1.09
1.02

Fig. 6. Typical De-bonded Specimen

5. CONCLUDING REMARKS
Pull-off tests were carried out on an apartment building columns and beams that were
repaired 12 years ago by wrapping them with CFRP and on bridge columns that were
repaired with CFRP 15 years ago. The average bond strength observed in the building
structure was 2.71 MPa for both columns and beams with a standard deviation of 1.02 MPa.
The average bond strength observed in the columns of the bridge was 3.10 MPa and the
standard deviation was 0.58 MPa. A minimum strength of 1.4 MPa for field tests is
considered to be adequate.
After 12 to 15 years of service in the Toronto area which is characterized by typical
Canadian weather and environmental conditions including extreme temperature variations,
freeze-thaw cycles and exposure to deicing salt, the bond between FRP and concrete was
found to be adequate.

17

6. ACKNOWLEDGMENTS
The authors would like to thank the Ministry of Transportation of Ontario for allowing
access to the bridge site and arranging traffic control and the management of Ahmadiyya
Abode of Peace for access to the bulding and providing amenities needed for field work.
Research support from ISIS Canada is also gratefully acknowledged.
7. REFERENCES
1.
2.
3.
4.
5.
6.
7.

ACI Committee 440. 2004. Guide Test Methods for Fiber-Reinforced Polymers (FRPs)
for Reinforcing or Strengthening Concrete Structures. ACI 440.3R-04, Farmington
Hills, MI, 40 p.
ASTM International Committee D-01. 1993. Standard Test Method for Pull-Off
Strength of Coatings Using Portable Adhesion Testers. ASTM D-4541, Pennsylvania.
Banthia, N., Demers, M., Mufti, A. and Sheikh, S. A. 2010. Durability of FRPConcrete Bond in FRP-Strengthened Bridges. ACI Concrete International, 32(8): 4551.
Canadian Standards Association (CSA). 2002. Design and Construction of Building
Components with Fibre-Reinforced Polymers. CSA S806-02, Mississauga, ON, 177 p.
Sheikh, S. 2007. Field and Laboratory Performance of Bridge Columns Repaired with
Wrapped GFRP Sheets. Canadian Journal of Civil Engineering, 34: 403-413.
Sheikh, S.A., DeRose, D. and Murdukhi, J. 2002. Retrofitting of Concrete Structures
with Fibre Reinforced Polymers. ACI Structural Journal, 99(4): 451-459.
Wan, B., Petrou, M.F., Harries, K.A., Sutton, M.A. and Yang, B. 2002. Experimental
Investigation of Bond between FRP and Concrete. The Third International Conference
on Composites in Infrastructure, San Fransisco, CA, June.

18

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

BOND BEHAVIOR OF EXTERNALLY BONDED FIBER


REINFORCED POLYMER LAMINATES SUBJECTED TO IN-SITU
SERVICE LOADING AND ENVIRONMENTAL CONDITIONING
J.J. Myers and N. Muncy
Missouri University of Science and Technology, Rolla, Missouri, USA

ABSTRACT
The use of fiber reinforced polymers (FRP) in repairing and strengthening bridges has been
an active area of research and implementation in recent years. In particular, adhering FRP
to the tension face of reinforced concrete (RC) beams has provided an increase in load
carrying capacity and extended the service life of the structures. The key to the efficient
performance of FRP strengthened concrete members is the bond between FRP sheets and
the concrete to which they are bonded. This study examines the bond durability of FRP
laminates installed on bridge structures in the State of Missouri, USA after several years of
exposure to service conditions including both mechanical and environmental loading.
Since 1998, more than twenty-five (25) bridges have been built and/or repaired using FRP
composite materials in the State of Missouri. One project was selected for this bond
behaviour study investigation presented herein. The paper discusses both the pull-off
strength and failure mode(s) of the bond tests conducted on FRP laminates under field
conditions in both actively stressed and largely unstressed locations. In terms of
environmental behaviour, the investigation examines FRP laminates with and without a
protective ultraviolet (UV) coating to investigate any differential level of degradation due
to UV in locations of direct and intense UV exposure. The study also compares the bond
performance of laminates in locations that are subjected to repeated moisture including wet
and dry cycles/freezing and thawing cycles. Failure modes are reported as well as bond
strength levels.
1. INTRODUCTION
Building and maintaining infrastructure in the United States of America is essential to the
economic and well-being of Americans. In the future, increasing amounts of infrastructure
such as bridges will become deficient and require structural attention [1]. In 2003, the
Missouri Department of Transportation retrofitted the Dallas County Bridge P-0962 with
Fiber Reinforced Polymers (FRP). Carbon FRP laminate sheets were bonded externally to
the concrete surface using a high strength epoxy to help increase the shear and flexural

19

capacity of the bridge [2]. As the FRP strengthening system is exposed to various
weathering effects including ultraviolet light, freeze/thaw, and water exposure, the bond
strength of the FRP-concrete interface over time was of particular interest. A premature
debonding failure occurring between the FRP-concrete interfaces will greatly reduce the
capabilities of the FRP strengthening system.
2. LITERATURE REVIEW
FRP for repair and strengthening of reinforced concrete (RC) systems has been an active
area of research in recent years. FRP has been used to strengthen bridges and provide
several advantages over other strengthening procedures. Some advantages include minimal
traffic disturbance, short construction times, and lightweight materials allowing for easier
construction [2]. In the August 2010 issue of American Concrete Institute (ACI) Concrete
International Magazine, an analysis of FRP performance was investigated on four inservice bridges in Canada. The results from the carbon fiber reinforced polymers indicated
low bond strengths and variability among each test [3]. While the environmental conditions
in Canada may be more severe than those experienced in Missouri, the results from the
Dallas County Bridge will reflect the various environmental effects on FRP experienced in
Missouri.
3. METHODOLOGY
The DYNA Z Pull-Off Tester with digital Manometer (Figure 1) was used to test the
strength of the FRP-concrete interfaces. This was performed by applying discs to the
exposed surface of the FRP using Loctite Professional Heavy Duty Epoxy. The discs were
secured and left for approximately 24 hours to ensure the epoxy reached the proper
strength. Then, a diamond bit was used to core around the disc and into the concrete to a
target depth of 6.4-mm (0.25-in). The pull-off tester was then used to pull the discs off in
direct tension (Figure 1) and a peak value was recorded by the pull-off tester.

Fig.1. DYNA Z Pull-Off Tester being placed over the disc to record the pull-off strength

20

4. RESULTS
Each disc was applied at a specific location to determine the failure mode due to different
environmental conditions. Figure 2 identifies the disc numbers and Figure 3 shows the
failure modes. Table 1 shows the disc locations and the results of each pull-off test. The
concrete temperature for all pull-off tests was approximately 41 F.

Fig. 2. Disc numbers corresponding to the disc location

Fig. 3. Failure modes of the discs


Discs 1, 2, and 3 were applied to panel 1 of the abutment, which did not have any
sandblasting performed (i.e. no surface preparation) on the concrete surface prior to the
application of the FRP. As shown in Figure 3, discs 1 and 2 pulled concrete indicating that
the failure occurred in the concrete rather than at the interface between the FRP and

21

concrete surface. Disc 3 was partially applied to a delamination zone previously induced for
delamination growth studies so the disc 3 results were discarded.
Discs 4, 5, and 6 were applied to panel 3 of the abutment, which had a moderately light
sandblast concrete surface preparation prior to application of the FRP. As shown in Figure
3, all three discs pulled concrete indicating that the failure occurred in the concrete yielding
the most promising failure mode and not between the FRP and concrete surface.
Discs 7, 8, and 9 were applied to panel 6 of the abutment, which had a heavy sandblast
performed on the concrete surface prior to the application of the FRP. As shown in Figure
3, all three discs pulled concrete again indicating that the failure occurred in the concrete
yielding the most promising failure mode and not between the FRP and concrete surface.
Discs 10, 11, and 12 were applied to the lower part of a column where direct exposure to
moisture in the form of water (both liquid and frozen) occurred. As shown in Figure 3, all
three discs pulled concrete but exposed some FRP causing the failure mode to occur
between the FRP-concrete interfaces.
Table 1. Pull-Off Test Results
Standard Deviation
Location
Disc Load (lbs) Stress (psi)
(psi)
Panel 1
1
2683
847
338
Panel 1
2
1170
370
Panel 1
3
1822
575
Panel 3
4
2607
823
141
Panel 3
5
2095
662
Panel 3
6
1717
542
Panel 6
7
943
298
82
Panel 6
8
1356
428
Panel 6
9
1426
450
Column
10
355
112
72
Column
11
792
250
Column
12
687
217
UV Protected
13
920
291
85
UV Protected
14
1187
375
UV Protected
15
646
204
UV Unprotected
16
1892
598
5
UV Unprotected
17
1915
605
UV Unprotected
18
2037
643
Slab
19
472
149
26
Slab
20
587
185
Conversion: 1 psi = 0.00689 MPa

22

Average
(psi)
608

676

392

193

290

601
167

Discs 13, 14, and 15 were applied to the pier bent that had an ultraviolet (UV) protection
applied over the FRP. The discs were pulled and the sub-surface was found to be saturated.
Figure 3 shows that the failure occurred in both the concrete and at the FRP-concrete
interface.
Discs 16, 17, and 18 were applied to the pier bent that did not have UV protection applied
over the FRP. Figure 3 shows that the failure mode was primarily between the FRPconcrete interface. Disc 18 was not cored correctly so the results for disc 18 were
discarded.
Discs 19 and 20 (Not Pictured) were applied underneath the slab of the bridge deck where
the FRP was subjected to the most frequent stress variations. The discs were pulled and the
failure mode occurred in the concrete. The concrete was extremely saturated causing the
discs to be removed under lower loads. This suggests that moisture was trapped between
the outermost layer of concrete and FRP.
It may be noted that two locations fell below the 1.38 MPa (200 psi) minimum bond
strength limit presented in ACI 440 design guideline, namely the column base and
underside of the slab. Both of these locations had evidence of moisture.
The overall bond strength results are generally higher than those recorded on the previous
study undertaken in Canada by Banthia et al. [3]. These variations could be due to the
varied exposure conditions including lower temperatures experienced in Canada and/or
variations in the strengthening system or its application. Figure 4 shows the mean
temperatures between Kansas City, Missouri and Quebec Canada as a point of reference
[4,5]. The average monthly temperature is 2.8 C to 8.3 C (5 F to 15 F) lower in Canada
depending on the month. The colder weather could create longer exposure to freezing
conditions resulting in deterioration or freeze/thaw damage of the FRP.

Temperature(F)

MeanTemperatures
90
80
70
60
50
40
30
20
10
0

KansasCity
Quebec

Jan Feb Mar Apr May Jun Jul Aug Sept Oct Nov Dec
Month

C = (F - 32) * 5/9
Fig. 4. Mean temperatures of Kansas City, Missouri and Quebec, Canada
23

5. CONCLUSION
The results showed that the failure mode occurred within the concrete based on the pull-off
tests for discs 1-9 (excluding disc 3). The results from this work indicated that roughness
did not have a long-term effect on the failure mode, but the roughness did play a role on
FRP-concrete interface bond strength. The highest level of roughness preparation did not
produce the optimal long-term bond strength.
When the FRP was exposed to water either constantly or intermittently such as at the base
of a column (even under low levels of stress), the failure mode occurred between the FRPconcrete interface. Discs 10, 11, and 12 clearly show that some concrete and FRP are
exposed causing the FRP-concrete mode of failure. This location also recorded the lowest
bond strength level. These results suggest that the bond strength of the FRP system is not
nearly as effective over time when exposed to moisture on a regular basis.
When direct daylight was exposed to the UV protected FRP, the failure mode occurred in
the concrete. These discs pulled low compared to the rest of the results due to the concrete
being saturated at the test region. Although the average strength was around 2.00 MPa (290
psi), more pull-off tests should be performed on the UV protected FRP when excessive
moisture is not present in the concrete. When the FRP was not UV protected, the failure
mode occurred between the FRP-concrete interfaces. Although the UV unprotected FRP
could withstand a greater load, the failure mode indicates that the UV protection had a more
desirable failure mode compared to the UV unprotected laminate area.
The discs placed on the underside of the slab recorded the lowest bond strength. The
concrete appeared to be saturated at the time of testing. The failure mode was in the
concrete indicating a desirable failure mode. However, the average bond strength was only
1.15 MPa (167 psi). More pull-off tests should be performed when the concrete is not
saturated to see if the moisture had an impact on the test results.
Based upon the pull-off tests, there is variability between the three tests in each location.
The standard deviation ranged from a high of 2.33 MPa (338 psi) to a low of 0.34 MPa (5
psi) throughout testing highlighting the variability in the test method. Additional bond
testing of more than ten other bridges in Missouri are planned during this calendar year to
increase the data pool on the long-term bond performance of externally bonded FRP
systems.
6. REFERENCES
1.
2.

Missouri S&T. 2010a. Preservation of Missouri Transportation Infrastructure: NonDestructive Testing of FRP Laminates. Missouri S&T, University Transportation
Center Report, http://utc.mst.edu/research/r096.html, December.
Missouri S&T. 2010b. Preservation of Missouri Transportation Infrastructure:
Validation of FRP Composite Technology through Field Testing-In-situ load testing on
Bridges P-962, T-530, X-495, X596, and Y-298. Missouri S&T, University
Transportation Center, http://utc.mst.edu/research/r095.html, December.

24

3.
4.

5.

Banthia, N., Demers, M., Mufti, A. and Sheikh, S. A. 2010. Durability of FRPConcrete Bond in FRP-Strengthened Bridges. ACI Concrete International, 32(8): 4551.
Canada's National Climate Archive. 2011. Monthly Observation Data - Canada's
National Climate Archive - Archives Climatiques Nationales Du Canada |
Meteorological Service of Canada - Service Mtorologique Du Canada.
(http://www.climate.weatheroffice.gc.ca/), January.
National Weather Service Climate. 2011. NOAA's National Weather Service.
http://www.weather.gov/climate/, January.

25

26

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

DURABILITY OF GFRP RODS IN FIELD DEMONSTRATION


PROJECTS ACROSS CANADA
A.A. Mufti1, J.Newhook2, B.Benmokrane3, G.Tadros4 and H.M. Vogel5
1

President, ISIS Canada Research Network, Winnipeg, MB, Canada


Professor, Dalhousie University, Dept. of Civil & Resource Engrg., Halifax, NS, Canada
3
NSERC/Canada Research Chair Prof., Dept. of Civil Eng., Universit de Sherbrooke, QC, Canada
4
Consultant, ISIS Canada Research Network, Winnipeg, MB, Canada
5
Ph.D. Student, University of Manitoba, Dept. Of Civil Engrg., Winnipeg, MB, Canada
2

ABSTRACT
In 2004, the ISIS Canada Research Network initiated a project in which the performance of
Glass Fiber Reinforced Polymer (GFRP) used in demonstration structures across Canada
was studied after 5 to 8 years of exposure to concrete. The study was designed around
apprehensions from the engineering research community to use GFRP as primary
reinforcement in concrete structures due to studies that have indicated that glass fibers may
deteriorate in the relatively high pH environment of concrete. The 2004 study indicated that
the apprehension of the research community was not valid as GFRP was in an excellent
condition. More recently, in 2009, core specimens of GFRP reinforcement were collected
from three of the field demonstration projects studied in the earlier rounds of tests.
Analytical methods including Scanning Electron Microscopy (SEM), Energy Dispersive Xray (EDX), Optical Microscopy (OM) and Differential Scanning Calorimetry (DSC) were
used to assess the condition of GFRP after 10 to 13 years of exposure to concrete in the
selected structures. These analyses were performed at the University of Manitoba by the
ISIS Canada Resource Center and at the University of Sherbrooke with the intent of
providing further information on the durability of GFRP in actual concrete field
environments as opposed to simulated laboratory alkaline environments.
1. INTRODUCTION
The choice of GFRP as a preferred material for concrete reinforcement in North America
relates to its economical competitiveness, its resistance to corrosion, its electromagnetic
permeability, its high strength to weight ratio as well as its excellent fatigue performance
(Memon and Mufti 2004, Sheikh and Homam 2004). However, research performed by
Mufti et al. (2005) emphasizes that numerous and, to some extent, contradictory statements
concerning the durability of GFRP in alkaline concrete environments have been published
in technical literature. Some of these publications suggest that exposure of reinforcement to
simulated concrete pore water and temperatures elevated to 80oC can cause a decrease in

27

the mechanical properties of the material (Porter and Barnes 1998, Bank and Russell 1995,
Sen et al. 2002, Bank et al. 1998). Other experiments have suggested little deterioration of
GFRP reinforcement after 12 months of exposure to alkaline solutions maintained between
20 and 38oC. Accordingly, research performed by Sheard et al. (1997) for the
EUROCONCRETE project indicates that GFRP reinforcement can resist alkaline
environments and that it is suitable for use in a concrete environment.
Results in this paper are intended to provide qualitative and quantitative information on the
performance of GFRP rebar extracted from actual structures across Canada. The samples
represent 10 to 13 years of exposure to concrete environments in the field and are not
bound by the limitations of standard tests performed in laboratory simulated environments.
2. BACKGROUND AND METHODOLOGY
2.1 Field Structures Selected for the Durability Study
From the five structures selected in the initial rounds of tests in 2004, three were
investigated for the durability analysis presented in this paper. The first structure, shown in
Fig. 1(a), is Halls Harbor Wharf. It is located in Nova Scotia on the Bay of Fundy shore
and comprises 45MPa concrete with steel-free precast concrete panels containing ISOROD
GFRP reinforcement and concrete pile cap beams reinforced with a hybrid GFRP-steel bar
system. The structure was 10 years old at the time of extracting the cores. It is typically
subjected to temperatures ranging between -35 and 35oC with wet and dry cycles as well as
splash and tidal salt water. It is also common for the structure to experience daily freezethaw cycles.
The second structure, the Joffre Bridge, is shown in Fig. 1(b). It is located in Sherbrooke,
Qubec, and spans the St-Franois River. The traffic barriers as well as the sidewalks were
designed with 45MPa concrete and GFRP C-bar reinforcement. The structure was 12 years
old at the time of core extraction and is subjected to the same environmental conditions as
Halls Harbour Wharf with the exception of splash and tidal water. It can also be overlaid
by substantial amounts of de-icing salts during the winter months.
The third and final structure is the Crowchild Trail Bridge located in Calgary, Alberta. The
structure is shown in Fig. 1(c) and contains ribbed-deformed GFRP C-bar reinforcement in
the barrier walls and the deck slab. Unlike Halls Harbour Wharf and Joffre Bridge, the
concrete used in the Crowchild Trail Bridge was designed to a compressive strength of
35MPa. The bridge was 13 years old at the time of core extraction and operates under a
thermal range of -15 to 23oC. It is also subjected to freeze-thaw cycles and can be regularly
overlaid with de-icing salt.

28

(a) Halls Harbour Wharf

(b) Joffre Bridge

(c) Crowchild Trail Bridge

Fig. 1. Field Projects Considered for Durability Assessment


2.2 Analytical Methods and Sample Preparation
Experienced contractors were hired to extract at least 5 cores of GFRP reinforced concrete
at various locations in each of the three structures considered. The samples had diameters
ranging from 75mm to 100mm and lengths varying between 75mm and 200mm.
The first analytical method applied to the samples was Optical Microscopy (OM). The
method was used to investigate the condition of the interface between the reinforcement
and surrounding concrete. The cores were carefully cut perpendicular to the axis of three
reinforcing bar segments using a high speed diamond saw. The cutting procedure ensured
that the surface roughness of the segments was minimal to ease sanding/polishing. The
sanding/polishing procedure for OM consisted of three progressively finer sanding steps
and one final polishing step with a 1 alumina powder solution.
The second analytical method considered consisted of performing a Differential Scanning
Calorimetry (DSC) analysis to establish the glass transition temperature Tg of the polymeric
matrix containing the glass fibers in the reinforcing bars. The tests were performed in
Winnipeg, Manitoba, by the Composite Innovation Centre (CIC) using a TA Instrument
Q2000 differential scanning calorimeter having an accuracy of 0.1oC. The values for Tg
will be compared to the values obtained during the first round of testing in order to detect
any changes in the structure and stability of the matrix. It has been documented (Mufti et al.
2005) that the value of Tg can be reduced by the presence of moisture or alkalis in the
polymer, which can cause cracking/hydrolysis of the matrix and fiber-matrix delamination.
The Infrared spectroscopy (FTIR) technique was used as the third method of analysis to
determine the presence of chemical species which are formed during the hydrolysis
reaction. The technique can reveal the presence or not of chemical degradation in the resin
matrix. The chemical durability of the resin matrix is governed to a large extent by the
chemical nature of the structure of the polymeric chain. All resin types have ester bonds in
their structure and this is the weakest link of the polymer. A possible degradation
mechanism of resin is alkali hydrolysis of the ester linkages. Due to the alkaline
environment in concrete, alkali hydrolysis is expected to some extent. During the
hydrolysis reaction, the OH- induces ester linkage attack and the resin chain is broken.
Consequently, the structure of the resin is disrupted and the material properties are
changed. The resin material degrades and eventually will not be able to transfer stresses to

29

the glass fibers and to protect the glass fibers against alkaline attack. Changes in the amount
of hydroxyl groups present in the composite material provide insight into the hydrolysis
reaction. The relative amount of hydroxyl groups in the specimens were measured by
determining the ratio between the maximum of the band corresponding to the hydroxyl
groups at 3430 cm-1 and the band corresponding to the carbon-hydrogen groups at 2900 cm1
in the FTIR spectra. The C-H content is considered constant. Since the vinyl ester resins
naturally contain hydroxyl groups, all the spectra present a strong absorption band in this
region. The FTIR analysis was performed at University of Sherbrooke on GFRP samples
from Joffre Bridge and Halls Harbour Wharf.
Scanning Electron Microscope and Energy Dispersive X-Ray (SEM/EDX) analysis was
also performed on samples of composite reinforcement from each of the three structures
considered in this paper. The SEM portion of the analysis was used to investigate whether
the fibers have been damaged by chemical degradation or other means. Conversely, the
EDX analysis was performed to trace chemicals that are likely to cause a degradation of the
fibers over the service life. The presence of sodium or potassium in the EDX scans is
generally an indication of alkali migration from the concrete pore solution toward the glass
fibers. The samples were cut using a low speed diamond saw to minimize surface damage
and finished with one sanding step and three progressively finer polishing steps using a
0.3 alumina powder solution.
3. ANALYSIS RESULTS AND DISCUSSION
3.1 Optical Microscopy
Selected scans from the OM are shown in Fig. 2 for each of the structures analyzed. None
of the scans taken intermittently along the length of the samples extracted from the cores
revealed degradation of the interface separating the reinforcement from the concrete after
10 to 13 years of exposure to alkalinity, freeze-thaw as well as wet and dry cycles, de-icing
salts, salt water and/or thermal loading. There are no visible gaps between the bars and the
surrounding concrete and cracks, slits or fissures are absent in both materials. The result is
encouraging and indicates that the bond has been preserved between the materials. The
composite reinforcement, the surrounding concrete and the interface are clearly identified
in the figure.
ISOROD
Interface

C-Bar GFRP

C-Bar GFRP
Interface

Interface
Concrete
Concrete

(a) Halls Harbour Wharf

Concrete

(b) Joffre Bridge

(c) Crowchild Trail Bridge

Fig. 2. Interface Scans from the OM Analysis

30

3.2 Differential Scanning Calorimetry


The glass transition temperature values for the GFRP reinforcement removed from the field
structures are presented in Table 1 along with those obtained from the first round of testing
performed by Mufti et al. (2005). The values for the most recent set of tests performed in
the final stages of 2009 represent an average obtained from a total of three samples that
were cut from each bar specimen given to the CIC and University of Sherbrooke for DSC
testing.
A total of two bar sizes were extracted from the concrete cores of Halls Harbour Wharf
and Joffre Bridge, namely 16mm and 9mm diameter. One bar size was extracted from
Crowchild Trail Bridge, namely 16mm diameter. The glass transition temperature values
for the GFRP specimens removed from the demonstration structures are presented in Table
1. It should be noted that the 9mm diameter bars from Halls Harbour Wharf and 16mm
diameter bars from Crowchild Trail Bridge were not considered during the initial round of
tests in 2004. In the absence of these values, a full comparison with the current state of the
reinforcement is not possible. The results presented in Table 1 indicate that the structure of
the polymer matrix was not significantly disrupted by exposure to the environment at the
Halls Harbour Wharf and Joffre Bridge. Furthermore, it can be noted that the Tg value
obtained for GFRP samples (C-Bar 16 mm) from Crowchild Trail Bridge is similar to the
value obtained for the same GFRP bar (C-Bar 16 mm) used in Joffre Bridge.
Table 1. Glass Transition Temperature (Tg)
Structure
GFRP
First round tests
(2004)
Tg (oC)
Halls Harbor Wharf
Isorod (16 mm dia.)
123

Joffre Bridge

Recent 2009
Tests - Tg (oC)
123

Isorod (9 mm dia.)

N/A

118

C-Bar (16 mm dia.)

107

103

C-Bar (9 mm dia.)

127

121

N/A

103

Crowchild Trail Bridge C-Bar (16 mm dia.)


3.3 Infrared Spectroscopy

Typical results of the FTIR analysis for control and in-service GFRP samples from Joffre
Bridge (Joffre Bridge 9 and 15 mm) and Halls Harbour Wharf (Isorod 15 mm) are shown
in Fig. 3. No significant change in the spectra of the specimens from the demonstration
structures was observed. The contents in resin matrix or glass fiber in the small amounts of
powder required in the analysis vary from one sample to another. Such differences affect
the intensity and the shape of the absorption bands.
Table 2 presents the content ratio of OH / CH for the control specimen as well as for those
removed from the Joffre Bridge and Halls Harbour Wharf demonstration structures. The
31

results show very little change in hydroxyl content in the exposed specimens. In effect, the
OH/CH ratio decreases in the sample from the structures (exposed) indicating that no
hydrolysis reaction occurred in the specimens during exposure of the demonstration
structures to natural environmental conditions.

Fig. 3. FTIR spectra for GFRP bars for Joffre Bridge and Halls Harbor Wharf
Table 2. Ratio of OH/CH for control and exposed specimens
Specimen
OH/CH ratio
Reference (C-Bar 9 mm)
0.57
Joffre Bridge (C-Bar 16 mm)
0.52
Halls Harbour Wharf (Isorod 16 mm)
0.50
Joffre Bridge (C-Bar 9 mm)
0.45
3.3 Scanning Electron Microscope and Energy Dispersive X-Ray
A representative micrograph from the SEM analysis performed on the cross-section of
reinforcement used in Halls Harbour Wharf is shown in Fig. (a). A small sand grain from
the surface finish of the reinforcement can be seen at the top of the scan. The scan does not
reveal any evidence of deterioration in the fibers, the polymer or the fiber/matrix interface.
The result confirms that the GFRP reinforcement in the structure has not undergone alkali
attack in the concrete environment. The white interspersed deposits found in the scan are
residues from the polishing process.

32

(a) Halls Harbor Wharf

(b) Joffre Bridge

(c) Crowchild Trail Bridge

Fig. 4. SEM Micrographs of Reinforcement Cross-Section


Fig. 4(b) shows a micrograph obtained from the SEM analysis performed on reinforcement
extracted from Joffre Bridge. Results from this scan as well as the remaining scans from
Joffre Bridge continue to suggest the absence of deterioration in the fibers, the polymer as
well as the interface. Similar conclusions can be made from the scan shown in Fig. 4(c) for
the reinforcement removed from Crowchild Trail Bridge. All structures therefore contain
reinforcement in which glass fibers have not been exposed to alkali attack from the
surrounding concrete.
In order to confirm results from the SEM analysis, a series of EDX analyses were also
performed on the GFRP samples from each structure. Since silica glass is known to
dissolve in strong alkaline solutions (Mufti et al. 2005), the analyses were used to assess
consequent chemical changes in the glass fibers. Results are illustrated in Fig. 5 for each of
the structures considered.

Si

Ca

(a) Halls Harbor Wharf

Si

Si

Ca

Ca
A

(b) Joffre Bridge

(c) Crowchild Trail Bridge

Fig. 5. EDX Scans of Glass Fibers

33

As observed during the initial round of tests, the main elements detected on the surface of
the samples are silica (Si), calcium (Ca) and Aluminum (Al). There is no evidence of
sodium (Na) or potassium (K) in the scans, which suggests that alkali solution from the
concrete pores has not migrated within the reinforcement of the structures considered.
4. CONCLUSIONS
Several conclusions can be drawn from the results of the analyses presented in this paper.
These conclusions pertain to the durability of GFRP reinforcement that has been used in
Halls Harbor Wharf, Joffre Bridge and Crowchild Trail Bridge for a period of concrete
exposure ranging between 10 and 13 years.

Results from OM did not show any sign of gaps or debonding between the GFRP
reinforcing bars and the surrounding concrete in all of the three structures examined.
The results also suggest that there were no cracks, slits or fissures in the concrete,
indicating that the bond was preserved between the materials.

The results from DSC indicate that the structure of the polymer matrix of GFRP
reinforcement was not significantly disrupted by exposure to the environment.
Neither hydrolysis nor significant changes in the glass transition temperature of the
matrix took place after exposure to the combined effects of concrete alkaline
environment and the external natural environmental exposure for 10-13 years.

The FTIR spectroscopy analyses support the results from the DSC analyses and OM
examinations.

Micrographs from the SEM analyses did not reveal any signs of physical damage to
the fibers, polymer or fiber/matrix interface. The result brings confirmation to the fact
that the structures were not exposed to the detrimental effects of alkali attack from the
surrounding concrete.

Accordingly, results from the EDX analysis do not show traces of alkaline solution
within the cross-section of the reinforcement for any of the structures considered.

5. ACKNOWLEDGMENTS
The authors would like to thank Lisa Hibbert and Eugene Rothwell from the CIC for their
help in performing DSC testing on the samples as well as Dr. Ravinder Sidhu from the
Department of Geological Sciences at the University of Manitoba and Dr. Patrice Cousin
from the Department of Civil Engineering at the university of Sherbrooke for their help in
performing the SEM/EDX and FTIR analyses.

34

6. REFERENCES
1.
2.
3.

4.
5.
6.
7.

8.

Memon, A.H. And Mufti, A.A. 2004. Fatigue Behaviour of Second Generation SteelFree Concrete Bridge Deck Slab. Proceedings of the 2nd International Conference on
FRP Composites in Civil Engineering (CICE 2004), Adelaide, Australia, pp. 765-772.
Sheikh, S. And Homam, M. 2004. A Decade of Performance of FRP-Repaired
Concrete Structures. Proceedings of the ISIS-SHM Workshop, Winnipeg, Manitoba,
Canada, pp. 1866-1873.
Porter, M. L. and Barnes, B. A. 1998. Accelerated Durability of FRP Reinforcement
for Concrete Structures. Proceedings of the 1st International Conference on Durability
of Fiber Reinforced Polymer (FRP) Composite for Construction (CDCC), eds.
Benmokrane, B. & El-Salakawy, E., Sherbrooke, Qubec, Canada, pp. 191-201.
Bank, L. C. and G. Russell T. 1995. Accelerated Test Methods to Determine the LongTerm Behaviour of FRP Composite Structures: Environmental Effects. Journal of
Reinforced Plastics and Composites, 14: 559-587.
Sen, R., Mullins, G., and Salem, T. 2002. Durability of E-Glass/Vinylester
Reinforcement in Alkaline Solution. ACI Structural Journal. 99(3): 369-375.
Bank L.C., Gentry, T.R., Barkatt, A., Prian, L., Wang, F. and Mangla, S. R. 1998.
Accelerated Aging of Pultruded Glass/Vinylester Rods. Proceedings of the 2nd
International Conference on Fiber Composites in Infrastructure (ICCI), 2: 423437.
Sheard, P., Clarke, J.L., Dill, M., Hammerslely, G. and Richardson, D. 1997.
EUROCONCRETE - Taking Account of Durability for Design of FRP Reinforced
Concrete Structures. Non-Metallic (FRP) Reinforcement for Concrete Structures,
Proceeding of 3rd International Symposium, Sapporo, Japan, 2: 75-82.
Mufti, A., Onofrei, M., Benmokrane, B., Banthia, N., Boulfiza, M., Newhook, J.,
Bakht, B., Tadros, G., and Brett, P. 2005. Studies of Concrete Reinforced with GFRP
Specimens from Field Projects. ISIS Canada Research Network Technical Report,
Winnipeg, Manitoba, Canada, 65 p.

35

36

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

EFFECT OF TEMPERATURE CHANGE ON THE BEHAVIOR OF


FRP-TO-CONCRETE BONDED JOINTS
W.Y. Gao 1, J.G. Teng 2 and J.G. Dai 3
1

PhD student, Dept. of Civil & Struct. Eng., Hong Kong Polytechnic Univ., China
Chair Professor, Dept. of Civil & Struct. Eng., Hong Kong Polytechnic Univ., China
3
Assistant Professor, Dept. of Civil & Struct. Eng., Hong Kong Polytechnic Univ., China
2

ABSTRACT
Service temperature variations may significantly affect the bond behavior between
externally bonded fiber reinforced polymer (FRP) reinforcement and concrete. This paper
presents a closed-form analytical solution for the full-range debonding process of FRP-toconcrete bonded joints under combined mechanical and thermal loads. The solution is
based on a bilinear bond-slip model and leads to closed-form expressions. Numerical
results from the solution are presented to illustrate the effect of temperature change on the
debonding load (i.e. bond strength). It is shown that provided the material properties are not
affected by a temperature change, a significant but realistic temperature decrease from the
installation temperature can significantly reduce the bond strength due to the development
of significant thermal interfacial stresses. A useful function of the closed-form solution lies
in the interpretation of pull test results: the solution allows the effect of thermal stresses to
be isolated from the effect of property changes of the bondline in obtaining bond-slip
responses from pull tests.
1. INTRODUCTION
In reinforced concrete (RC) structures strengthened with externally bonded fiber reinforced
polymer (FRP) reinforcement, the bond behavior between FRP and concrete often controls
the load-carrying capacity of the strengthened structure [1]. As a result, many studies have
been conducted on the bond behavior of FRP-to-concrete interfaces (e.g. [2-6]). In
particular, the single-lap pull test (Fig. 1) (or a double-lap pull test which can be seen as
two single-lap tests being conducted simultaneously) has been widely used to study the
ultimate load (i.e. bond strength) of FRP-to-concrete bonded joints and the local bond-slip
behavior of the interface (e.g. [2-4, 7]).
FRP-strengthened RC structures in service are likely to experience significant temperature
variations (e.g. ambient temperature changes and exposure to fire) and such variations can
have a significant effect on the bond performance of FRP-to-concrete interfaces. To
understand the bond behavior of FRP-to-concrete interfaces at different temperatures, the
37

pull test has also been used (e.g. [8-12]). Results from such tests reflect directly the
combined effects of a number of factors including temperature-induced interfacial stresses
and temperature-induced property changes in the bondline (the adhesive and the adjacent
parts of the adherends) and the adherends. A key purpose of such pull tests is to determine
the bond-slip curve of the interface at a specific temperature change, for which the effect of
thermal interfacial stresses needs to be isolated from the effect of temperature-induced
bondline property changes, as only the latter should be included in a bond-slip model for
use in a theoretical model for FRP-strengthened RC structures subjected to temperature
variations. This issue has received little attention and indeed in some existing studies, these
thermal stresses were simply ignored in interpreting pull test results (e.g. [9,10]).
This paper presents an analytical solution for the full-range behavior of FRP-to-concrete
bonded joints at a specific temperature variation from the ambient temperature at
installation. The solution is an extension of the existing analytical solution of Yuan et al.
[6] and serves two important purposes: (a) to provide the first theoretical explanation of
bond strength variations observed in tests due to temperature variations; (b) to provide a
rigorous theoretical basis for the experimental determination of the bond-slip curve by
isolating the effect of thermal stresses from the effect of property changes of the bondline.

FRP
Adhesive
Concrete
d

Fig. 1. Schematic diagram of a single-lap pull test.


2. ANALYTICAL SOLUTION
2.1 Assumptions and Notation
Figure 1 shows the theoretical idealization of a single-lap bonded joint where both
adherends are assumed to experience only membrane deformation. Moreover, it is assumed
that the width and thickness of each of the three components (plate, adhesive layer and
concrete prism) are constant in the longitudinal direction. In such a simplified theoretical
model, the adhesive layer (i.e. the interface or the bondline whose deformation represents
the deformation of the actual adhesive layer and that of the adjacent parts of the two
adherends) is subjected to shear deformation, so that mode II interfacial fracture is the
failure mode. This theoretical model is a close approximation of the behavior of a real
bonded joint [6]. In addition, it is assumed that the interface is still within the linear elastic
range of behavior and the properties of the adherends are not affected by the imposition of
the thermal load; these two assumptions means that the degree of temperature change needs
38

to be appropriately limited. The applicability of the solution to bonded joints whose


adherends have experienced property degradations is discussed later in the paper.
In Fig. 1, ,
and are the thickness, width and bonded length of the FRP plate,
respectively, while
and
are the width and thickness of the concrete prism,
respectively. The elastic moduli of the plate and the concrete are
and , respectively.
For convenience of presentation, the left end of the plate ( =0) is referred to as the free end
and the right end ( = ) the loaded end hereafter.
2.2 Governing Equations
Based on the above assumptions, the horizontal equilibrium consideration of the FRP plate
and of the overall joint cross-section give the following equations:
0

(1)
0

(2)

where is the shear stress in the adhesive layer,


is the axial stress in the FRP plate and
is the axial stress in the concrete prism. The constitutive equations for the adhesive layer
and the two adherends are described by
(3)

(4)

(5)

where
and
are the longitudinal displacements of the FRP plate and of the concrete,
and
are the coefficients of thermal expansion of the FRP plate and of
respectively;
the concrete, respectively; and is the service temperature variation (thermal load).
The interfacial slip
that is

is defined as the relative displacement between the two adherends;


(6)

From Eqs. (1)-(6), the following second-order differential equation can be derived:

(7)

where

39

(8)

In Eqs. (7) and (8), is the local bond strength (i.e. the maximum shear stress on the bondis the interfacial fracture energy (the area underneath the interfacial
slip curve) and
bond-slip curve).
Substituting Eq. (6) into Eqs. (2), (4) and (5) yields

(9)

2.3 Bond-Slip Model


A bilinear bond-slip model has been widely used to study the bond behavior of FRP-toconcrete bonded joints (e.g. [6] ); such a bilinear model is also adopted in the present study.
As shown in Fig. 2, the bilinear bond-slip relationship consists of two segments: an initial
linear elastic segment where the bond shear stress increases linearly with the interfacial slip
until it reaches the peak stress (the corresponding slip is denoted by ) and a softening
segment where the shear stress reduces with the interfacial slip until it becomes zero (the
corresponding slip is denoted by ). This bond-slip model (Fig. 2) is described by the
following equation:
when0
when

(10)

0when

Fig. 2 Bi-linear bond-slip model.


2.4 Predictions for the Debonding Process
With the bond-slip model defined above and following the approach of Yuan et al. [6], the
governing equation (Eq. 7) can be solved to find the interfacial shear stress distribution
along the interface and the load-displacement response of the bonded joint under combined
thermal and mechanical loads.

40

For a bond slip model as shown in Fig. 2, the debonding process can be divided into three
stages [6]: (1) the elastic stage (Stage I), during which the load is small and the interfacial
shear stress stays below ; (2) the elastic-softening stage (Stage II), during which the shear
slip at the loaded end ( = ) has exceeded but is smaller than ; (3) the elasticsoftening-debonding stage (Stage III), during which the interfacial slip at the loaded end
and the shear stress there has reduced to zero; during
has exceeded the separation slip
this stage, interfacial debonding initiates at the loaded end and propagates along the
interface. Since the interfacial shear stress and the tensile stress in FRP can be easily
derived from the interfacial slip, only the expressions for the interfacial slip and the loaddisplacement relationship at the loaded end are presented below. Other details of the
analytical solution can be found elsewhere [13].
2.4.1. Elastic stage (Stage I)
During the elastic stage (i.e Stage I), the bond-slip curve is given by the first expression of
Eq. 10. Substituting Eq. 10 into Eq. 7 and making use of the boundary conditions ( = 0 at
= 0;

at = ), the expression for the interfacial slip is obtained:

where

sinh

(11)

The slip at the loaded end (i.e. the value of at = ) is also referred to as the displacement
of the bonded joint and is denoted by . Based on this definition, the load- displacement
relationship is given by:

tanh

sinh

(12)

The initial displacement due to the temperature change can be calculated as (i.e. = 0
and = in Eq. 11):

ctanh

sinh

(13)

At the end of the elastic stage, the interfacial shear stress reaches with a slip at
the loaded end. Substituting = into Eq. 12 leads to the load at the initiation of
interfacial softening (the beginning of the elastic-softening stage):
1

sinh

41

tanh

(14)

2.4.2. Elastic-softening stage (Stage II)


When the loaded end slip first exceeds , the free end slip is still less than , so the right
part of the interface is now in a softening state while the left part of the interface is still in
the linear elastic state. As the load further increases, the length of the softening zone
also increases. In the elastic zone (0
,0
), the interfacial slip has the
same form as Eq. 11:

cosh

In the softening zone (


in the FRP plate are given by

(15)

), the interfacial slip and the axial stress

tanh

sinh

sin

cos

where

(16)

(17)
at the loaded end ( = ) into Eq. 17 yields

tanh

cos

sin

and the displacement

tanh

sin

Substituting the boundary condition

cos

(18)
at the loaded end can be found from Eq. 16 (i.e. = ) to be
tanh

sin

cos
(19)

Eqs. 18-19 can be used to predict the load-displacement relationship for the elasticsoftening stage by varying the value of .
2.4.3. Elastic-softening-debonding stage (Stage III)
At the end of Stage II, the slip at the loaded end reaches , indicating the initiation of
debonding at the loaded end. The corresponding value of is its maximum possible value
(denoted by ) and is determined using the following equation which is obtained from Eq.
19:
42

tanh

sin

cos

(20)
For an infinite bond length, Eq. 20 reduces to
arctan

(21)

As the debonding crack propagates, the location of the peak shear stress
moves away
from the loaded end towards the free end, and as a result, the total length of the intact
interface reduces. Assuming that the length of the debonded interface is , the equations for
the interfacial slip (Eqs. 15 & 16) and the load (Eq. 18) are still valid if is replaced by ( ). Whereas, the displacement can be identified with the consideration of the thermalinduced slip in the debonded length, as shown in following

(22)

3. NUMERICAL RESULTS
Klamer [11, 12] conducted a series of pull tests on double-lap FRP-to-concrete bonded
joints (Fig. 3) at temperatures ranging from -20 oC to 100 oC. The test specimens (including
the installation of strain gauges on the FRP plate) were prepared at 20 oC (the reference
temperature) but the tests were conducted at a different temperature (-20 oC, 20 oC, 40 oC,
50 oC, 70 oC 80 oC or 100 oC). The thicknesses of the FRP pultruded plate and the adhesive
layer were 1.2 mm and 1.5 mm, respectively, while the bond length was 300 mm. Figure 4
presents a comparison between the predicted thermal strains in the CFRP plate and the
experimental values for one of the specimens which was subjected to a temperature
increase of 30 oC (i.e. = 30 oC) before the external load was applied (Klamer [11]
reported the thermal strains of the FRP plate only for this specimen). In making the
predictions, the following parameters were used as provided by Klamer [11]: = 100 mm,
= 75 mm, = 150 mm, =165 000 MPa, =26 800 MPa, = 10.210-6 /oC and =
0.310-6 /oC. From the experimental load-displacement curves of the two specimens tested
at the reference temperature (20 oC), the loads corresponding to the initiation of softening
and debonding were found to be 20 KN and 44.7 KN respectively [13], so the interfacial
parameters were identified as: =0.35 mm, =0.08 mm, =3.07 MPa, and
=0.54
N/mm [6]. The close agreement between the test results and the analytical predictions
demonstrates the validity of the present closed-form solution.
Figure 5 shows a comparison between the debonding loads of FRP-to-concrete bonded
joints obtained from Klamers tests [11] and the analytical solution. In this figure, the
debonding loads are normalized by the corresponding experimental or predicted value for
the reference temperature of 20 oC for a clearer comparison. It is seen that the experimental
debonding load initially increases as the temperature increases (or decreases as the
temperature reduces) but the trend reverses when the temperature is around the glass
transition temperature
of the bonding adhesive which was 62 oC [11]. The analytical
43

debonding load increases monotonically with the temperature. The differences between the
experimental debonding loads and the analytical predictions for temperatures above the
glass transition temperature are due to the omission of the effect of softening of the bonding
adhesive as in the present predictions the bilinear bond-slip law was identified from the test
results for the reference temperature of 20 oC. In addition, any softening of the adherends
was also ignored. The close agreement between the experimental results and the analytical
predictions for temperatures below
further verify the reliability of the closed-form
solution.
Figure 6 presents numerical results from the analytical solution to examine the effect of
temperature change on the normalized debonding load for FRP plates of different
thicknesses (i.e. different tensile stiffnesses), with = 0 oC being for the reference case.
It is clearly seen that for the same temperature rise, the increase in the debonding load is
larger when a thicker FRP plate is used. If a 2.4 mm thick FRP plate is used, a temperature
increase of 50 oC leads to about a 27% increase in the debonding load. This aspect needs to
be considered in engineering practice when an FRP-strengthened structure is subjected to
significant service temperature variations.
300

Analytical predictions
Test data

Strain in FRP (um/m)

250
200
150
100
50
0

Fig. 3. Test specimens of Klamer [11]

Normalized ultimate debonding load (KN)

Normalized ultimate debonding load

1.4
1.2
1
0.8
0.6
0.4
0.2
0

20

100
150
200
Distance from the left end (mm)

250

300

1.3

Analytical predictions
Test data

1.6

0
-20

50

Fig. 4. Thermal strain distribution in the FRP

2
1.8

40

60

80

1.25
1.2
1.15
1.1
1.05
1
-10

100

Temperature ( C)

Fig. 5. Normalized ultimate debonding load


vs. temperature

0.15 mm
0.3 mm
0.6 mm
1.2 mm
2.4 mm

10

20
30
o
Temperature ( C)

40

50

60

Fig. 6. Normalized ultimate debonding load


vs. temperature

44

4. DETERMINATION OF BOND-SLIP CURVES FROM PULL TESTS


It is now clear that the debonding load of an FRP-to-concrete bonded joint can be
significantly affected by a temperature variation from the installation temperature. This has
a significant implication when pull test results are used to derive bond-slip curves. In such a
derivation, the interfacial fracture energy
is directly related to the debonding load of the
bonded joint when temperature variations are not involved (e.g. [4,5 14]). When pull tests
are used to derive bond-slip laws at a temperature change, previous authors have followed
the same approach [9]. In a correct approach, the effect of thermal stresses on the
debonding load needs to be eliminated using the present solution for the establishment of
the bond-slip curve which may include the effect of softening of the bonding adhesive.
Readers are referred to Ref. [13] for detailed discussions.
5. CONCLUSIONS
This paper has presented a closed-form analytical solution for the full-range behavior of
FRP-to-concrete bonded joints under combined mechanical and thermal loads. A bilinear
local bond-slip relationship is employed in the solution, but the general characteristics
observed from the solution are applicable to interfaces with a similar bond-slip model. The
solution provides close-form expressions for the interfacial slip, interfacial shear stress,
stress in the FRP plate as well as the load-displacement relationship for the entire
debonding process. The predictions of the closed-form solution have been compared with
the limited test data available, demonstrating close agreement between them. It has been
shown that the debonding load increases as the temperature increases provided the material
properties of the bonded joint have not been affected by the temperature change; this
increase is more significant when a stiffener FRP plate is used. The present solution can be
used to design against debonding failure in FRP-strengthened steel/concrete structures
subjected to significant service temperature variations. In addition, the solution also
provides a useful tool by which the test results of FRP-to-concrete bonded joints with
temperature variations can be correctly interpreted to obtain bond-slip curves.
6. ACKNOWLEDGEMENTS
The authors are grateful for the financial support received from the Research Grants
Council of the Hong Kong SAR (Project No: PolyU 516509) and The Hong Kong
Polytechnic University provided through a PhD studentship to the first author.
7. REFERENCES
1.
2.
3.

Teng, J.G., Chen, J.F., Smith, S.T., and Lam, L. 2002. FRP-Strengthened RC
Structures. John Wiley & Sons Ltd., Chichester, 266 p.
Chen, J.F., and Teng, J.G. 2001.Anchorage strength models for FRP and steel plates
bonded to concrete. Journal of Structural Engineering, ASCE, 127(7): 784-791.
Yao, J., Teng, J.G., and Chen, J.F. 2005. Experimental study on FRP-to-concrete
bonded joints. Composites Part B: Engineering, 36(2): 99-113.

45

4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.

Dai, J.G., Ueda, T., and Sato, Y. 2005. Development of the nonlinear bond stress-slip
model of fiber reinforced plastics sheet-concrete interfaces with a simple method.
Journal of Composites for Construction, ASCE, 9(1): 52-62.
Dai, J.G., Ueda, T., and Sato, Y. 2006. Unified analytical approaches for determining
shear bond characteristics of FRP-concrete interfaces through pullout tests. Journal of
Advanced Concrete Technology, JCI, 4(1): 133-145.
Yuan, H., Teng, J.G., Seracino, R., Wu, Z.S., and Yao, J. 2004. Full-range behavior of
FRP-to-concrete bonded joints. Engineering Structures, 26(5): 553-565.
De Lorenzis, L., Miller, B., and Nanni, A. 2001. Bond of Fiber-reinforced Polymer
Laminates to Concrete. ACI Material Journal, 98(1): 256-264.
Blontrock, H. 2003. Analysis and Modeling of the Fire Resistance of Concrete
Elements with Externally Bonded FRP Reinforcement. PhD Thesis, Ghent University,
Ghent, Belgium.
Wu, Z.S., Iwashita, K., Yagashiro, S., Ishikawa, T., and Hamaguchi, Y. 2005.
Temperature effect on bonding and debonding behavior between FRP sheets and
concrete. Journal of the Society of Materials Science, 54(5): 474-480.
Leone, M., Matthys, S., and Aiello, M.A. 2009. Effect of elevated service temperature
on bond between FRP EBR systems and concrete. Composites Part B: Engineering,
40(1): 85-93.
Klamer, E. 2006. The Influence of Temperature on Concrete Structures Strengthened
with Externally Bonded CFRP. Research Report, Eindhoven University of Technology,
Eindhoven, Netherlands.
Klamer, E. 2009. Influence of Temperature on Concrete Beams Strengthened in
Flexure with CFRP, PhD Thesis, Eindhoven University of Technology, Eindhoven,
Netherlands.
Gao, W.Y., Teng, J.G., and Dai, J.G. 2011. Effect of temperature variation on the fullrange behavior of FRP-to-concrete bonded joints. (in preparation).
Lu, X.Z., Teng, J.G., Ye, L.P., and Jiang, J.J. 2005. Bond-slip models for FRP
sheets/plates bonded to concrete. Engineering Structures, 27(6): 920-937.

46

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

LONG-TERM BOND BEHAVIOUR OF GFRP REBAR IN SEVER


ENVIRONMENTS
B. Juette1, A. Weber2 and C. Witt3
1

Product Manager ComBAR, Schoeck Bauteile GmbH, Baden-Baden, Germany


Head of R & D, Division of GFR, Schoeck Bauteile GmbH, Baden-Baden, Germany
3
President, Schoeck-Canada Inc., Kitchener, ON
2

ABSTRACT
Increasingly glass fibre reinforced polymer (GFRP) rebars are being installed - instead of
conventional steel reinforcement - in concrete structures located in corrosive, that is moist
and warm, environments. Recent material developments have resulted in much higher
tensile strengths and moduli of elasticity compared to GFRP bars of earlier generations.
These material developments make newer generation GFRP bars suitable for long-term
applications under high sustained stresses. Under these conditions the long-term bond
properties of these bars are critical. For the structural design long-term design values of the
bond strength must be determined.
The bond behaviour and the bond creep behaviour of reinforcing bars in general are
discussed in the paper. The critical issues, especially those applying to GFRP bars, are
defined. The results of a test series on the bond durability and a test series on the bond
creep behaviour of a newest generation GFRP rebar are presented. Finally, a procedure for
the determination of long-term design values of the bond strength using the results of the
bond creep tests is explained in detail.
1. INTRODUCTION
Glass fibre reinforced polymer (GFRP) reinforcing bars were first developed as crack
reinforcement in North America in the 1980s. Massive problems encountered there due to
the corrosion of the steel reinforcement in bridge decks had led to the search for alternative
reinforcing materials to conventional steel rebar. As the prices for stainless steel are
extremely volatile and usually very high, GFRPs continue to be an economic reinforcement
solution for bridge decks and barrier walls.
The first materials available on the market had relatively low tensile strengths and moduli
of elasticity. The approach at the time was, therefore, to install comparatively large
amounts of bars with large diameters (high rebar ratio) in the bridge decks. Thereby the
stresses in the bars were kept relatively small while the widths of the cracks in the decks
47

were limited to acceptable sizes. As the corresponding bond stresses on the bars were
relatively low, bond behaviour was tested only in short-term tests. Little attention was paid
to long-term bond behaviour and bond creep behaviour of these bars. (ACI440; CSA S80602)
In the late 1990s the search for corrosion resistant reinforcing bars began in Europe, where
high prices for stainless steel rebar made its installation increasingly uneconomic. Here the
focus lay on load bearing reinforcement in relatively thin concrete elements, such as
balcony slabs, subjected to high loads.
As the stresses which needed to be permanently sustained by these bars were much higher,
durability and long-term behaviour became a central issue in the qualification tests of these
bars. A testing concept was developed to analyse the bond creep behaviour of these bars
and determine design values of the long-term bond stress which insure acceptable
performance of these bars for service life spans of as much as 100 years.
2. BOND CREEP
2.1 Bond Creep in Reinforced Concrete
Bond creep of steel reinforcement in concrete is defined as the time dependent increase of
the slip of a steel reinforcing bar embedded in concrete under permanent load (Franke
1976).
Bond creep occurs along the embedment length of reinforcing bars as the slip of the bars
increases with time. The stresses in the bar and the surrounding concrete change in that
region, as a result. Bond creep is one of the reasons for the time dependent increase of the
widths of cracks in reinforced concrete elements.
One of the causes of the increase in the relative displacement between the reinforcing bar
and the surrounding concrete lies in the visco-elastic and plastic compression of the
concrete corbels as local compressive stress peaks occur in the immediate vicinity of the
bar ribs. Radial micro cracking of the concrete is most likely more pronounced near the ribs
or the reinforcing bar, also contributing to an increase in the relative displacement between
the concrete and the reinforcing bar.
2.2 Bond Creep in Concrete Reinforced with GFRP Bars
GFRP bars are available from a large number of producers worldwide. A large number of
different materials (fibres and resins) are used in their production. Their geometry and
surface properties and therefore their properties in concrete vary greatly. Whereas bars,
especially those of earlier generations, are often laid, newer generation bars, conceived to
sustain higher permanent tensile stresses, are usually produced in a so called pultrusion
process. In this process the bars are pulled through the machinery at high tension. This
ensures a nearly perfect linear and parallel alignment of the fibres in the bars.

48

Bar surfaces also vary greatly. A number of producers sand coat the finished bars to insure
bond, others apply ribs to the core of the bars during the production process or after
hardening of the bars, others cut ribs into the hardened bars.
The bond between a reinforcing bar and the surrounding concrete is ensured via three
mechanisms: adhesion, shear (force transfer in to the concrete corbels parallel to the bar)
and friction between the bar and the concrete. All three mechanisms are strongly influenced
by surface geometry, texture and consistency. As GFRP bars contain linearly oriented
fibres, they are, unlike steel, not isotropic. As a result, their bond behaviour is additionally
controlled by the inter-laminar shear properties of the material itself (transfer of forces
along the interface between the ribs and the core cross section of the bar).
As a result of all these considerations the bond and bond creep behaviour of the various
GFRP bars available on the world market differ significantly. Detailed and strenuous shortand long-term bond and bond creep testing is required to determine safe design values of
the bond stress for each individual bar.
3. TESTING BOND CREEP BEHAVIOR OF FRP REBARS
3.1 Intent of Testing Series
In conjunction with the applications for general construction authority certifications of a
newest generation pultruded GFRP reinforcing bar in Germany and in The Netherlands
(DIBt 2008, KOMO, KIWA 2009) a testing scheme was developed to determine the
allowable design value of the long-term bond stress sustainable by these bars. The
behaviour of the bars after the concrete has cracked and their bond creep behaviour in
higher grade concretes were to be studied in an extensive testing series using this scheme.
3.2 Test Specimen
Eighteen (18) test specimen were prepared according to the RILEM testing
recommendation RC 6. GFRP rebars with a core / nominal diameter of 8, 16 and 25 mm
were tested. The concrete cubes had side lengths of 150mm and 250mm respectively. The
bars were located in the center of the concrete cubes. The bottom end of the bars extended
beyond the face of the cube to allow for the measurement of the slip at the unloaded bar
end. The embedment length of the bars was four or five times the bar diameter. The rest of
the bar length was covered by plastic pipes to debond the bar (Weber 2009).

Fig. 1. Curing of test specimen in water


49

The concrete recipe was devised so as to obtain a compressive strength of 75MPa (grade
C50/60 according to EC-2). The specimen were cast horizontally and tested vertically. This
configuration corresponds to the concreting direction in real life near-surface installations
of the bars. Each prism was reinforced with two closed carbon steel stirrups (d = 6 mm,
grade 500 steel) located at the beginning and in the middle of the embedment length. The
specimens were cured in water at room temperature. They were at least 28 days old when
tested.
3.3 Testing Concept
The central idea behind the testing scheme is to study the behaviour of the bars in cracked
concrete under extreme yet realistic environmental conditions.
To simulate these conditions alkaline concrete (Portland cement) was used on all specimen.
The specimens were pre-loaded to a slip of 1 mm at the free bar end at room temperature.
For the long-term bond creep tests the specimen were heated to 60C and were kept water
saturated at all times. This environment simulates real life conditions for service life spans
as they are required in real life projects. The load was sustained on the bars for at least 2000
hours without them showing any signs of increasing slip due to bond creep.
Bond Creep Test Sheme
25

Loading over peak

bond stress in MPa

20

15

unloading to 7,5 MPa


10
2000h creep at 60C in wet
alkaline concrete

design value for bond


0
0

1
2
slip at unloaded end of bar in mm

Fig. 2. Concept of bond creep testing


3.4 Pre-Loading of Specimen
In a first step a tensile load was applied on the bars at room temperature until a total slip of
1mm was measured at the end of the bar opposite to the load application. The slip was
recorded as a function of the (bond) force applied on the bars.
The maximum bond stress achieved by the 8mm bars during this pre-loading was between
20 and 26 MPa at a slip between 0.5 and 0.7 mm.

50

30

bond
force in kN
Verbundkraft
in kN

25
20
8 V10
8 V11

15

8 V12
10
5
0
0

0,2

0,4

0,6

0,8

1,2

1,4

slip at unloaded end in mm

1,6

Schlupf am lastabgewandten Ende in mm

Fig. 3. Pre-loading to 1mm slip d = 8 mm ComBAR bars


The maximum bond stress for the 16 mm bars was between 22 and 25 MPa at a slip
between 0.1 and 0.4mm.
120

bond
force in in
kNkN
Verbundkraft

100
80
16 V9
60

16 V10 002

40
20
0
0

0,2

0,4

0,6

0,8

slipamatlastabgewandten
unloaded end inEnde
mmin mm
Schlupf

1,2

Fig. 4. Pre-loading to 1mm slip d = 16 mm ComBAR bars


The maximum bond stress for the 25mm bars was between 16 and 19 MPa at a slip
between 0.3 and 0.4mm.
160

bond
force in kN
Verbundkraft
in kN

140
120
100

V1
V2

80

V3
60
40
20
0
0

0,2

0,4

0,6

0,8

1,2

slipam
at lastabgewandten
unloaded end in Ende
mm in mm
Schlupf

Fig. 5. Pre-loading to 1mm slip d = 25 mm ComBAR bars


51

3.5 Bond Creep Testing


After this pre-loading all specimen were unloaded and prepared for the long-term test. To
simulate real life conditions and to accelerate any aging processes in the bars the specimen
were saturated in water and heated to 60C prior to the application of the load. A constant
tensile load was then applied to the bars. The applied loads and the resulting bond stresses
are shown in Table 1.
bar diameter
[mm]
8mm
8mm
16mm
25mm

Table 1. Applied constant loads, bond stresses ComBAR


specimen
embedment
load [kN]
bond
stress
designation
length [mm]
[MPa]
V4-V6
40
9.60
9.55
V7-V12
40
8.07
8.03
V7-V10
80
30.31
7.54
V1-V6
100
70.47
8.97

The loads were sustained on the bars for at least 2000 hours. The tests were only declared
acceptable if there were no signs of an increase in the bond creep over this period of time.
The slip curves are shown in Figures 6 to 8.
The 8mm bars did not pass the test at a bond stress of 9.5 MPa (see Figure 6). They were,
however, able to sustain a bond stress of 8.0 MPa for up to 5000 hours.
1,0

0,9

displacement
bar end
[mm]
Weg amat
Stabende
[mm]

0,8

0,7

8-V4 9,5 MPa


8-V5 9,5MPa
8-V6 9,5MPa
8-V7 8MPa
8-V8 8MPa
8-V9 8MPa
8-V11 8MPa
8-V12 8MPa

0,6

0,5

not acceptable

0,4

0,3

0,2

0,1

0,0
0

500

1000

1500

2000

2500

Laufzeit
[h]
time
[hours]

3000

3500

4000

4500

5000

Fig. 6. Bond creep curves d = 8mm ComBAR bars at 60C, pre-loaded to 1mm slip
The 16 mm bars were able to sustain a permanent bond stress of 7.5 MPa without showing
signs of growing bond creep.

52

0,50
0,45

displacement
bar end
Weg am at
Stabende
[mm] [mm]

0,40
0,35
0,30

16-V7
16-V8
16-V9

0,25

16-V10

0,20
0,15
0,10
0,05
0,00
0

500

1000

1500

2000

2500

3000

3500

4000

4500

5000

5500

6000

[h]
timeLaufzeit
[hours]

Fig. 7. Bond creep curves 16mm ComBAR, 60C, bond stress 7.5 MPa, pre-loaded to
1 mm slip
The 25 mm bars were able to sustain a permanent bond stress of 9.0 MPa without showing
signs of growing bond creep.
0,50
25-V1
25-V2

displacement
at bar end [mm]
Weg am Stabende [mm]

0,45

25-V3
25-V4
25-V5

0,40

25-V6

0,35

0,30

0,25

0,20

0,15

0,10

0,05

0,00
0

500

1000

1500

2000

2500

3000

3500

4000

4500

5000

[h]
timeLaufzeit
[hours]

Fig. 8. Bond creep curves 25mm ComBAR at 60C, bond stress 9.0 MPa, pre-loaded to
1 mm slip
3.6 Inspection of Specimen after Testing
One concrete cube from each testing series was cut open after the test to inspect the GFRP
bars. Most of the ribs of the 8 and 16 mm bars had been sheared off in the tests. On the
25mm tests the concrete corbels had failed.
4. DERIVATION OF DESIGN VALUES OF BOND STRENGTH
Based on the results of these tests, the design values of the bond strength can be specified
for the specific GFRP reinforcing bar and for individual bar diameters thereof for service
53

lives up to 100 years. The characteristic value of the bond stress is derived by statistical
means from the bond stress sustained by at least five tests over at least 2000 hours without
showing any increase in the bond creep. It applies, of course, only to the concrete grade
tested in the series. In Germany a safety factor of 0.65 is applied to this characteristic value
to obtain the design value of the bond stress for that particular bar and the specific grade
concrete. (Schiel 2008)
In the case of the bar tested in this series the design values of the bond stress in normal
grade concretes were found to be essentially the same as those of steel reinforcement as it is
commonly used in central Europe.
5. CONCLUSIONS
The bond creep properties of GFRP rebars are a critical measure of their durability and
long-term load bearing characteristics. As the materials used in the production of the
various GFRP rebars available on the market differ greatly, and the bars themselves have
different surface profiles, geometries and textures, their bond properties differ greatly.
A testing procedure has been developed to determine the long-term bond behaviour (bond
creep) of GFRP reinforcing bars in pre-cracked concrete sections. In this procedure real life
environmental conditions are simulated by testing the bars in saturated highly alkaline
concrete prisms. To speed up the testing process and to simulate long-term applications of
the bars the tests are conducted at 60C.
In the tests the characteristic or guaranteed value of the bond stress is determined for the
tested concrete for a design service life of up to 100 years. The design value of the bond
stress for the particular grade concrete is derived from the results of these tests by applying
a safety factor of 0.65.
6. REFERENCES
1.
2.
3.
4.
5.

ACI Committee 440. 2004. Guide test methods for fiber reinforced polymers FRPs for
reinforcing or strengthening concrete structures. ACI 440.3R-04, American Concrete
Institute, Farmington Hills, MI, 40 p.
Canadian Standards Association (CSA). 2002. Design and construction of building
components with fibre reinforced polymers. CAN/CSA S806-02, Mississauga, ON,
177 p.
Franke, L. 1976. Einfluss der Belastungsdauer auf das Verbundverhalten von Stahl in
Beton, Verbundkriechen, Heft 268 DAfStb, Beuth Verlag, Berlin, Germany.
Schiel, P., Volkwein, A., Knab, F. 2008. Application for DIBt Certification of GFRP
Reinforcing Bars, ComBAR made by Schck Bauteile GmbH. expert report, Munich,
Germany.
Weber, A. 2009. Bond creep of ComBAR bars with d = 8 mm, d = 16 mm and d = 25
mm, Lab report, Schck Bauteile GmbH, Baden-Baden, Germany.

54

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

EFFECT OF VARIOUS SEVERE ENVIRONMENTAL EXPOSURES


ON CFRP AND SFRP-CONFINED CIRCULAR CONCRETE
COLUMNS
M. A. Mashrik1 and R. El-Hacha2
1
2

MSc Student, Department of Civil Engineering, University of Calgary, Calgary, Canada


Associate Professor, Department of Civil Engineering, University of Calgary, Calgary, Canada

ABSTRACT
This paper evaluates the performance of plain concrete cylinders strengthened with Steel
Fibre Reinforced Polymer (SFRP) and Carbon Fibre Reinforced Polymer (CFRP) sheets
exposed to different aggressive environments. A total of forty-five plain concrete cylinders
(150 mm diameter 300 mm height) were divided into five groups with nine specimens in
each group (three unconfined, three CFRP-confined, and three SFRP-confined). One group
was considered control and kept at room temperature (+22C). The remaining four groups
were subjected to four different environments: (a) prolonged high temperature of +45C for
135 days; (b) prolonged high temperature of +60C along with relative humidity of 96% for
42 days; (c) 90 freeze-thaw cycles; and (d) 471 freeze-thaw cycles. Each freeze-thaw cycle
was 7 hours and 8 minutes between -34C to +34C and a relative humidity of 75% for
temperature above +20C. Monotonic concentric uniaxial compression tests were
conducted on the specimens and the experimental results were correlated with statistical
analysis. Results showed that severe environments had a detrimental effect on the axial
strength of the CFRP wrapped cylinders, but not on that of the SFRP wrapped cylinders.
1. INTRODUCTION
Fibre Reinforced Polymer (FRP) materials have emerged as an economic and efficient
solution for strengthening structural concrete members. FRP sheets can be bonded onto
flexural members to increase flexural and shear capacity, or wrapped around the columns to
improve axial strength and ductility (El-Hacha et al., 1999). Such FRP-confined columns
may be exposed to aggressive environments during their service life. In Canada, exterior
columns (especially in bridges and parking structures) are likely to be subjected to
temperature extremes and freeze-thaw cycles. For the practical acceptance of FRP-confined
columns, it is necessary to evaluate the effects of aggressive environments on the
performance of the FRP confinement. This paper deals with the use of the newly developed
Steel Fibre Reinforced Polymer (SFRP) sheet to wrap plain concrete cylinders and compare

55

with the traditionally used Carbon Fibre Reinforced Polymer (CFRP) sheet; both at room
temperature and under different severe environments.
2. EXPERIMENTAL PROGRAM
2.1 Materials
2.1.1 Concrete
The target 28 day unconfined concrete compressive strength was 40 MPa while the average
of three 100200 mm concrete cylinders tested at 28 days was 43.1 MPa with a standard
deviation of 2.1 MPa.
2.1.2 Steel Fibre Reinforced Polymer (SFRP) Sheet
The SFRP sheet used in this project was type 32-20-12 made of unidirectional brass
coated ultra high strength twisted steel wires with a thickness of 1.23 mm and a net crosssectional area of 0.38 mm2/mm. According to the manufacturer, the ultimate tensile
strength and effective modulus for the SFRP sheet are 986 MPa and 66100 MPa,
respectively, and the ultimate tensile strain at failure is 1.5% (Hardwire, 2010a).
2.1.3 Carbon Fibre Reinforced Polymer (CFRP) Sheet
The CFRP sheet used in this project was SikaWrap Hex 230C made of unidirectional
carbon fibre fabric with a thickness of 0.381 mm. According to the manufacturer, the
ultimate tensile strength and modulus of elasticity are 894 MPa and 65402 MPa,
respectively, and the strain at failure is 1.33% (Sika, 2010b).
2.1.4 Epoxy Adhesive
Sikadur 330 was used to bond the SFRP and CFRP sheets around the cylinders. The
Sikadur330 is a two-component epoxy resin that has a tensile strength of 33.8 MPa,
modulus of elasticity of 3489 MPa and 1.9% elongation at break (Sika, 2010c).
2.2 Strengthening Procedure and Instrumentation
The concrete cylinders and FRP sheets were cleaned off dust before wrapping. The SFRP
and CFRP sheets were cut into desired lengths, allowing an overlap length of 100 mm.
Using the wet-lay up procedure, a thin layer of epoxy adhesive was applied to one side of
the FRP sheets and the sheets were wrapped around the cylinders in a continuous manner.
All cylinders were wrapped with one layer (300mm wide) of CFRP sheet or SFRP sheet
having almost the same axial stiffness (EACFRP= 7.48kN and EASFRP= 7.51kN, where E is
the modulus of elasticity of the FRP sheet and A is the area/300mm width).
The cylinders were instrumented with two strain gauges on two opposite sides. For FRP
wrapped cylinders, both were to measure circumferential strains; for unwrapped cylinders,
one was to measure circumferential strain and the other was to measure axial strain.
56

2.3 Test Set-up and Test Matrix


The cylinders were tested under monotonic concentric uni-axial compression load on a 2
MN compression testing machine at an average rate of 10 kN/sec until failure. Forty-five
cylinders were divided into five groups with nine specimens in each group (three
unconfined, three CFRP-confined, and three SFRP-confined). Table 1 presents the detailed
test matrix of the experimental program.

Specimen
Designation
RT-UW
RT-C
RT-S
HT-UW
HT-C
HT-S
HTRH-UW
HTRH-C
HTRH-S
FT90-UW
FT90-C
FT90-S
FT471-UW
FT471-C
FT471-S

Environmental
Exposure

Table 1. Test Matrix


Duration
Details
(Days)

Room
Temperature

22C

Prolonged High
Temperature

+45C

135

High
Temperature
+ Relative

+60C and
96% RH

42

Freeze-Thaw
Cycles
Freeze-Thaw
Cycles

"-34C" to
"+34C with
75%RH"
"-34C" to
"+34C with
75%RH"

27
(90 cycles)
140
(471 cycles)

Wrapping
Material
Unwrapped
CFRP
SFRP
Unwrapped
CFRP
SFRP
Unwrapped
CFRP
SFRP
Unwrapped
CFRP
SFRP
Unwrapped
CFRP
SFRP

No. of
Specimens
3
3
3
3
3
3
3
3
3
3
3
3
3
3
3

The specimen designation in Table 1 are explained as follows: except for the freeze-thaw
exposure, the first letters represent the environment (RT for room temperature, HT for
high temperature, HTRH for high temperature with relative humidity), followed by
letters representing wrapping material (UW for unwrapped, S for SFRP, C for
CFRP). For the freeze-thaw exposure, the first letters represent the environment condition
(FT for freeze-thaw), followed by number of cycles, and followed by wrapping material.
3. EXPERIMENTAL RESULTS AND DISCUSSION
In this section the results are reported and discussed in terms of ultimate strength, ductility,
stress-strain behaviour, and modes of failure of the concrete cylinders. The results are also
correlated with statistical analysis using the Single Factor Analysis of Variance (ANOVA)
performed to study the significance of different environments on the ultimate strength of
the cylinders. The ANOVA results are presented in Appendix A.
3.1 Ultimate Strength
The average ultimate strengths of the cylinders of different groups are presented in Table 2.
Table 2 shows that the average axial strength of the wrapped cylinders was superior to that
57

of the unwrapped cylinders, irrespective of the environmental conditions, indicating the


efficiency of FRP sheets as strengthening materials. The axial strength of the SFRP
wrapped cylinders was always superior to that of CFRP wrapped cylinders subjected to
same environment indicating that SFRP sheet is more effective than CFRP sheet in strength
enhancement of columns.

Specimen
Designation
RT-UW
RT-C
RT-S
HT-UW
HT-C
HT-S
HTRH-UW
HTRH-C
HTRH-S
FT90-UW
FT90-C
FT90-S
FT471-UW
FT471-C
FT471-S

Table 2. Summary of Axial Strength Results


Avg. Axial
Standard
% Increase or
% Increase
Strength
Deviation
Decrease
w.r.t. UW
(MPa)
(MPa)
w.r.t. RT
58.0
0.81

64.6
1.67
11.4

104.1
2.12
79.5

46.5
4.58

-19.8
57.1
1.66
22.8
-11.6
104.0
0.94
123.7
-0.1
51.2
7.07

-11.7
62.4
1.71
21.9
-3.4
101.8
2.00
98.8
-2.2
49.8
4.94

-14.1
61.2
0.66
22.9
-5.3
96.1
7.56
93.0
-7.7
53.6
2.37

-7.6
62.6
0.90
14.6
-5.0
106.3
0.70
98.3
2.1

% Increase
from CFRP
to SFRP

61.1

82.1

63.1

57.0

69.8

3.1.1 Effect of High Temperature


Table 2 shows that exposure to prolonged high temperature (HT) decreased the strength of
the unwrapped cylinders by 19.8% compared to room temperature exposure. The change in
average strength due to exposure to high temperature is statistically significant (Appendix
A). This is likely due to the micro-cracks developed in the concrete due to high
temperature, and some bond cracks at cement-aggregate interface as the Coefficient of
Thermal Expansion (CTE) are different for cement and aggregate. Prolonged high
temperature caused statistically significant changes in the average strength of the CFRP
wrapped cylinders (Appendix A). The decrease in strength of the CFRP wrapped cylinders
subjected to prolonged high temperature was 11.6% compared to the room temperature
cylinders. The reduction in ultimate strength of CFRP wrapped cylinders may be attributed
to the differences in the CTEs of the fibre and matrix, resulting in fibre slippage and matrix
cracking; and to carbon fibre degradation induced by high temperature. However,
prolonged high temperature did not affect the strength of the SFRP wrapped cylinders.

58

3.1.2 Effect of High Temperature along with Relative Humidity


Table 2 shows that exposure to combined high temperature and relative humidity (HTRH)
decreased the strength of the unwrapped, CFRP wrapped and SFRP wrapped cylinders by
11.7%, 3.4%, and 2.2%, respectively. The ANOVA analysis shows that this exposure did
not cause statistically significant difference in strength. Visual observation before testing
revealed that the steel fibres of the SFRP sheet were corroded at different locations around
the cylinders; however, it did not have any detrimental effect on the strength.
3.1.2 Effect of Freeze-Thaw Cycles
Appendix A shows that exposure to freeze-thaw cycles caused statistically significant
change in the strength of the unwrapped cylinders. Table 2 shows that the strength of the
unwrapped cylinders decreased by 14.1 and 7.6%, respectively, due to freeze-thaw cycles.
When exposed to freezing, the water in capillary pores in concrete freezes and expands
causing a tensile pressure in the concrete. During thawing, water goes back to the capillary
pores again. When, after thawing, another freezing cycle takes place, pressure is created in
the concrete again. In this way, repetitive freeze-thaw cycles causes damage to the concrete.
Exposure to 90 freeze-thaw cycles decreased the strength of the CFRP and SFRP wrapped
cylinders by 5.3% and 7.7%, respectively. Exposure to 471 freeze-thaw cycles decreased
the strength of CFRP wrapped cylinders by 5.0%, but increased the strength of the SFRP
wrapped cylinders by 2.1%. The change in strength was not statistically significant for the
SFRP wrapped cylinders, but was statistically significant for the CFRP wrapped cylinders.
The changes in strength between FT90 and FT471 groups were not also statistically
significant (Appendix A). The two effects of freeze-thaw cycles are thermal incompatibility
and polymer embrittlement. The first effect is related to the different CTEs of matrix and
fibre, as a result of which micro cracking and void generation may occur at matrix-fibre
interface. The second effect is polymer embrittlement at low temperature which increases
the strength and stiffness of the polymer. It may reduce the effectiveness of the matrix to
transfer stresses between fibres, or between the composite and the concrete (Green et al.,
2006). While observing the specimens before test, rust was found on the steel fibres of the
SFRP sheet at different locations; however, it did not affect the strength.
3.2 Ductility
Ductility can be defined as the ability to absorb inelastic energy without losing its load
capacity (Wang et al., 2005). In this study, the ductility index is determined as the area
under the stress-strain curve up to ultimate load. The average ductility indices of the
cylinders of different groups are presented in Table 3. Table 3 shows that the average
energy ductility index of the wrapped cylinders was superior to that of unwrapped
cylinders, irrespective of the environmental exposure. The average ductility of the SFRP
wrapped cylinders was always superior to that of CFRP wrapped cylinders subjected to
same environment, indicating SFRP sheet is more ductile than CFRP sheet. Severe
environments did not cause any statistically significant change in the ductility indices of the
wrapped cylinders, indicating severe environments had no significant impact on the
ductility of the wrapped cylinders.
59

Specimen
Designation
RT-UW
RT-C
RT-S
HT-UW
HT-C
HT-S
HTRH-UW
HTRH-C
HTRH-S
FT90-UW
FT90-C
FT90-S
FT471-UW
FT471-C
FT471-S

Table 3. Summary of Ductility Index Results


Avg. Ductility
Standard
% Increase
% Increase
Index
Deviation
or Decrease
w.r.t. UW
w.r.t. RT
(MPa-)
(MPa-)
35.4

661.4
160.9
1768.4

1233.7
76.7
3385.0

20.4
3.9

-42.4
572.0
92.4
2703.9
-13.5
1236.0
145.1
5958.8
0.2
7.6
1.9

-78.5
645.1
48.7
8388.2
-2.5
1100.3
96.9
14377.6
-10.8
7.5
7.5

-78.8
707.8
707.8
9337.3
7.0
1033.7
1033.7
13682.7
-16.2
20.2
9.2

-42.9
558.5
132.8
2664.9
-15.6
1075.6
180.4
5224.8
-12.8

% Increase
from CFRP
to SFRP

86.5

116.1

70.6

46.0

92.6

3.3 Stress-Strain Response

60

1060

50

884

40

707
RT
HT
HTRH
FT90
FT491

30
20
10
0
0

2000

4000

6000

530
353
177

120

2121

100

1767

80

1414

60

1060
RT
HT
HTRH
FT90
FT471

40
20
0

0
8000 10000 12000 14000

3000

6000

9000

12000

15000

707
353

Axial Compressive Load (kN)

1237
Axial Compressive Load (kN)

70

Axial Compressive Stress (MPa)

Axial Compressive Stress (MPa)

Figure 1 shows the typical stress-strain response of the CFRP and SFRP wrapped cylinders.

0
18000

Circumferential Strain ( )

Circumferential Strain ( )

(a)
(b)
Fig. 1. Typical stress-strain response of the cylinders wrapped with (a) CFRP; (b) SFRP
Figure 1 shows that the stress-strain curves of the wrapped cylinders were not altered by the
severe environments. The curves can be characterized by three zones: (1) linear elastic zone
roughly up to strain level of 500 with slope similar to that of unwrapped cylinders,
indicating the behaviour in this zone is similar to that of unwrapped cylinders. (2) nonlinear transition zone approximately between strain level of 500-6000 (CFRP) and
500-2500 (SFRP). In this zone, after the peak stress, the curve is nearly a plateau for
the CFRP wrapped cylinders, but ascending for the SFRP wrapped cylinders. It indicates
that for the CFRP wrapped cylinders, the rate for the confining sheet to contribute to the
60

load carrying capacity is nearly equal to the rate for the concrete to lose its load carrying
capacity, but for the SFRP wrapped cylinders, the former is higher than the latter. It implies
that SFRP wrapped cylinders show better strain-strain response compared to CFRP
wrapped cylinders. (3) linear ascending zone until failure, indicating concrete is completely
crushed and FRP sheet is fully activated.
3.3 Modes of Failure
Exposure to aggressive environments did not seem to have any significant effect on the
modes of failure of the specimens. All the unwrapped cylinders failed in conventional shear
or cone. The SFRP wrapped cylinders failed either by rupture of the SFRP sheets or by
combination of rupture and debonding at the overlap. All of the CFRP wrapped cylinders
failed by rupture of the CFRP sheets. The failure of the wrapped cylinders was always
brittle and sudden, with no prior warning except for some creeping sound of concrete
crushing. Figure 2 shows the typical failure modes of the wrapped cylinders.

Debonding

Fig. 2. Typical modes of failure of the wrapped cylinders


4. CONCLUSIONS
From the compression tests and statistical analysis of the CFRP and SFRP wrapped
cylinders, the following conclusions were drawn:
FRP confinement was found to be an effective and efficient way to increase the axial
strength and ductility of columns, both at room temperature and under severe
environments. However, for the same axial stiffness of the FRP sheet, one layer of
SFRP sheet was found to be more efficient than one layer of CFRP sheet in strength and
ductility enhancement.
Aggressive environments had a detrimental effect on the axial strength of the CFRP
wrapped cylinders, but not on that of the SFRP wrapped cylinders. Severe environments
had no significant effect on the ductility of the wrapped cylinders.
The axial stress-circumferential strain curves of the CFRP and SFRP wrapped cylinders
were not altered by the presence of aggressive environments. However, cylinders
wrapped with one layer of SFRP sheet exhibit better stress-strain behaviour than the
cylinders wrapped with one layer of CFRP sheet.
Exposure to aggressive environments did not have any significant effect on the failure
mechanism of the cylinders.
The SFRP sheets were susceptible to material degradation in the form of rust under
environments where moisture was involved. However, it did not have any negative
impact on the strength. The CFRP sheets were not susceptible to such damage.

61

5. REFERENCES
1.

2.
3.
4.
5.
6.

El-Hacha, R., Green, M., and Wight, G. 1999. CFRP wrapped concrete cylinders in
severe environmental conditions. Structural Faults and Repair 99, Proceedings of the
8th International Conference and Exhibition on Extending the Life of Bridges, Civil
and Building Structures, London, England, CD-Rom, 13-15 July, 12 p.
Hardwire Composite Armor Systems. 2010a. What is Hardwire?
(http://www.hardwirellc.com/solutions/reinforcments.html).
Sika Inc. 2010b. Sikawrap Hex 230C. Product Technical Data Sheet.
(http://www.sikaconstruction.com/tds-cpd-SikaWrapHex230C-us.pdf).
Sika Inc. 2010c. Sikadur 330. Product Technical Data Sheet, High-modulus.
(http://www.sikaconstruction.com/tds-cpd-Sikadur330-us.pdf).
Green, M.F., Bisby, L.A., Fam, A.Z., and Kodur, V.K.R. 2006. FRP Confined
Concrete Columns: Behaviour under Extreme Conditions. Cement and Concrete
Composites, 28: 928-937.
Wang, H. and Belarbi, A. 2005. Flexural Behavior of Fiber-Reinforced-Concrete
Beams Reinforced with FRP Rebars. Proceedings of the 7th International Symposium
on Fibre-Reinforced Polymer Reinforcement for Concrete Structures (FRPRCS-7),
ACI Special Publications SP-230, Kansas City, USA, 2: 895-914.

APPENDIX A. STATISTICAL ANALYSIS


The null hypothesis of the single factor ANOVA is: means of two or more populations are equal;
the alternate hypothesis is: at least one of the means is different. The ANOVA was conducted on a
95% confidence level. Table 4 shows the results of the ANOVA analysis.

Unwrapped

CFRP
Wrapped

SFRP
Wrapped

Table 4. Summary of the results of ANOVA analysis


Comparison
F-Ratio
F-Critical
P-Value
ALL
2.92
4.06
0.100256
RT
HT
18.30
7.71
0.012867
RT
HTRH
2.71
7.71
0.175314
RT
FT90
8.10
7.71
0.046592
9.35
7.71
0.037762
RT
FT471
1.45
0.294677
FT90 FT471
7.71
ALL
13.41
4.06
0.001735
RT
HT
30.65
7.71
0.005203
RT
HTRH
2.69
7.71
0.176562
RT
FT90
10.72
7.71
0.030652
8.51
0.043337
RT
FT471
7.71
0.09338
0.775163
FT90 FT471
7.71
ALL
2.56
4.06
0.128223
RT
HT
0.0018
7.71
0.968319
RT
HTRH
1.92
7.71
0.238117
RT
FT90
3.12
7.71
0.152056
3.12
0.151841
RT
FT471
7.71
5.44
0.080102
FT90 FT471
7.71

62

Decision
Not Significant
Significant
Not Significant
Significant
Significant
Not Significant
Significant
Significant
Not Significant
Significant
Significant
Not Significant
Not Significant
Not Significant
Not Significant
Not Significant
Not Significant
Not Significant

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

TEMPERATURE AND CURING TIME EFFECTS ON THE CREEP


RESPONSE OF FIBER REINFORCED POLYMER COMPOSITES
BONDED TO CONCRETE
Y. Jeong1, A. Jaipuriar2, M.M. Lopez1 and C.E. Bakis2
1
2

Dept. of Civil & Env. Eng. Pennsylvania State University, University Park, PA, USA
Dept. of Eng. Science and Mechanics, Pennsylvania State University, University Park, PA, USA

ABSTRACT
The objective of this investigation is to experimentally evaluate and numerically model
creep in carbon fiber reinforced polymer (CFRP) bonded to concrete and in neat epoxy.
Two test temperatures and epoxy curing times were considered. Based on the experimental
observations, an epoxy creep model was incorporated into the numerical modeling of the
time-dependent behavior of the pull-off specimens. Results indicate that the effects of
temperature and epoxy aging time on the time-dependent behavior of the FRP compositeconcrete bond are significant. These results provide critical data for characterizing the timedependent behavior of RC structures strengthened with bonded FRP composites.
1. INTRODUCTION
Fiber reinforced polymer (FRP) composites have emerged as a promising method for
rehabilitating existing reinforced concrete (RC) structures on account of their ease of
application, their corrosion resistance, and high strength-to-weight ratio [1]. While many
research activities on bonded FRP systems have been performed to improve predictive
models capable of explaining bond behavior [2], relatively few research efforts have
addressed bond performance under aggressive environmental conditions. In particular, not
much is known about long term FRP/concrete bond behavior under sustained loads. Recent
research [3] demonstrated that FRP-concrete bond behavior is affected by temperature,
sustained load and the presence of flaws at the interface. Specifically, significant creep
occurs in the epoxy when loaded within 7 days of epoxy application. Furthermore, it was
found that elevated temperature combined with sustained load led to significant bond stress
redistribution and a reduction in joint strength [3].
In the present investigation, experimental results from single shear (pull-off) specimens and
neat epoxy tensile coupons under sustained loading are obtained and used to develop finite
element models of FRP strips bonded to concrete. Two primary parameters were
63

considered: test temperature and the epoxy age following installation (curing time).
Experimental results indicate that the effects of temperature and epoxy curing time on the
time-dependent behavior of FRP composites to concrete are significant. The epoxy creep
model used in the numerical simulations captures the general trend of the CFRP timedependent deformations.
2. EXPERIMENTAL PROGRAM
2.1. Test Specimens and Materials
The concrete blocks (127 x 127 x 254 mm) and cylinders (100 mm diameter x 200 mm
long) were cast in one batch. All of the concrete specimens were cured in a moisture
chamber (95% relative humidity at 23C for 28 days). The average 28-day compressive
strength of the concrete was 33.9 MPa. A total of twelve pull-off specimens were
fabricatedthree for each test condition. Carbon FRP (CFRP) strips were attached to the
prepared concrete block according to manufacturers guidelines. A 25.4-mm-wide, single
ply of unidirectional carbon fiber strip was saturated with epoxy and bonded to each
concrete block as shown in Fig. 1. The CFRP composites bonded to concrete blocks were
cured by allowing them to cure in laboratory conditions (35% relative humidity at 20C)
for 7 or 90 days before applying sustained loading. According to the manufacturers
datasheet, full cure should be achieved after 7 days at 20C and 40% relative humidity. The
thickness of the impregnated CFRP sheet was found to be between 0.6 to 1.0 mm. Other
properties of the cured laminate and epoxy were obtained experimentally by Coronado and
Lopez [4]. Room-temperature material properties of CFRP and epoxy used in the finite
element models are shown in Table 1.
Table 1. Material properties used in the finite element models
Property
Unidirectional CFRP
Epoxy
Tensile strength
4.667 MPa
44.5 MPa
Elastic Modulus
208 GPa
2.8 GPa
Ultimate Elongation
2%
Poissons Ratio
0.39

(a)

(b)

Fig. 1. (a) Instrumentation of a pull-off specimen, (b) Fabrication of pull-off specimens.

64

2.2. Sustained Loading and Epoxy Creep Tests


The main parameters examined in the sustained loading of pull-off specimens were test
temperature during sustained loading (RT=20C, HT=30C) and the curing time of the
CFRP prior to sustained loading (7 days, 90 days), as shown in Table 2. Three replicate
specimens were tested for each combination of parameters.
Table 2. Test plan for pull-off specimens.
Specimen
Name
RT-7d
RT-90d
HT-7d
HT-90d

Curing time
(day)
7
90
7
90

Temperature during
sustained loading period
RT (20C)
HT (30C)
RT (20C)
HT (30C)

Sustained loading
period (day)
30
30
30
30

Pull-off specimens were instrumented with small stainless steel discs with machined pinholes for the use of a detachable mechanical (DEMEC) gauge. Seven target points were
fixed on the surface of the CFRP strip and nine reference points were fixed on the concrete.
By measuring the movement of each target point with respect to its corresponding reference
point, the longitudinal displacement of the CFRP strip along its length was calculated. The
accuracy of the DEMC gauge is 2.54e-3 mm. In this investigation, the creep deformation of
the CFRP strip was solely evaluated based on the variation of the longitudinal
displacements of target points D5 and D6 (Fig. 1(a)).
Sustained loading was applied to the prepared specimens as a percentage of a theoretical
debonding failure load, calculated using Cheng and Tengs bond model [5]. A sustained
load equivalent to 27% of the predicted ultimate pull-out load capacity was applied, which
corresponds to a load of approximately 2.67 kN. The sustained loading set-up consisted of a
steel frame in which the pull-off specimens were mounted. The sustained load was applied
by means of hooks and springs. Six loaded specimens were kept in room-temperature
laboratory conditions (20C) as shown in Fig. 2(a) while six others were kept at 30C in an
oven, as shown in Fig. 2(b).

(a)
(b)
(c)
Fig. 2. Creep tests(a) RT pull-off specimens, (b) HT pull-off specimens, and (c) neat
epoxy tensile coupons.
65

Neat epoxy creep tests were carried out in tension at temperatures of 20C and 30C (Fig.
(c)). Prior to testing, the epoxy specimens were aged under conditions similar to those for
pull-off specimens (room-temperature for either 7 or 100 days). Loading was set with dead
weights to provide a normal stress of 10 MPa in the epoxy. Longitudinal strains on the front
and back sides of the specimens were recorded with 25-mm extensometers and averaged to
compensate for bending. Three specimens were tested at each temperature. Further details
of the epoxy creep investigation are given in a companion paper [6].
2.3. Tests Results and Discussion
Results from the sustained loading period are presented in Figs. 3 and 4. Results are
presented as the longitudinal displacement at target points D5 and D6 as functions of time.
It can be observed that at target point D5, HT specimens exhibited larger displacements (of
the order of 0.127 mm) than the RT specimens (which remained very close to 0 mm). This
trend was very clear for all specimens regardless of the epoxy age. It appears that the
presence of a 30C temperature during sustained loading induces larger creep
displacements on the bonded CFRP strip. Target point D6 showed a similar trend for the
90-day cure specimens (with the exception of specimen HT-2). However, the 7-day cured
specimens showed a reversed trendi.e., more deformation occurs in RT specimens
compared to the HT case. Future investigation will determine the cause of the reversed
trend, such as microcracks at the bonded interface of the RT specimens.
Displacement (mm)

0.3
0.2

Displacement (mm)

HT-1
HT-2
HT-3
RT-1
RT-2
RT-3

D5

0.4

0.1
0
-0.1
0

10

20

Time (day)

30

0.3
0.2
0.1

HT-1
HT-2
HT-3
RT-1
RT-2
RT-3

D5

0.3
0.2

0
10

20

Time (day)

30

20

Time (day)

30

40

HT-1
HT-2
HT-3
RT-1
RT-2
RT-3

D6

0.1

-0.1
0

10

(b) Point D6, 7 day


Displacement (mm)

Displacement (mm)

(a) Point D5, 7 day


0.4

D6

0
-0.1
0

40

HT-1
HT-2
HT-3
RT-1
RT-2
RT-3

0.4

0.4
0.3
0.2
0.1
0
-0.1
0

40

(c) Point D5, 90 day

10

20

Time (day)

30

40

(d) Point D6, 90 day

Fig. 3. Effect of temperature on the longitudinal displacement of bonded CFRP strips with
different curing conditions.

66

It was also found that after the instantaneous elastic displacement occurs, the majority of
displacement changes occurred in the first 15 days for the HT specimens. In contrast, the
RT specimens showed that most of the change in time-dependent displacement occurred
during the first 3 days. It can be concluded that the presence of a 30C temperature results
in larger CFRP displacement due to creep, irrespective of the age of the epoxy, and it
stabilizes over a longer period of time than for specimens tested at 20C.
The influence of the epoxy curing time on the creep behavior of bonded CFRP strips can be
observed in Fig. 4. The 7-day cured specimens showed larger creep displacements than 90day cured specimens when the sustained loading was conducted at 20C. Aside from the 7day-cured, 20C-tested outliers (possible microcracking), it can be concluded that
increasing the temperature to 30C during sustained loading results higher creep
displacements and little disparity in the displacements for the 7-day and the 90-day cured
specimens. Therefore, at room temperature it can be expected that larger curing times can
have beneficial effects in reducing creep deformations. Increasing the temperature
decreases the curing time effect, but creep values could be larger in magnitude.

0.1

0.05

Displacement (mm)

Displacement (mm)

0.5

7d-1
7d-2
7d-3
90d-1
90d-2
90d-3

D5

-0.05
0

10

20

Time (day)

30

0.4
0.3

0.1

(a) Point D5, 20C


0.5

0.3
0.2

20

Time (day)

30

40

0.5

7d-1
7d-2
7d-3
90d-1
90d-2
90d-3

D6

Displacement (mm)

Displacement (mm)

0.4

10

(b) Point D6, 20C


7d-1
7d-2
7d-3
90d-1
90d-2
90d-3

D5

D6

0.2

0
0

40

7d-1
7d-2
7d-3
90d-1
90d-2
90d-3

0.1

0.4
0.3
0.2
0.1

0
0

10

20

Time (day)

30

0
0

40

10

20

Time (day)

30

40

(c) Point D5, 30C


(d) Point D6, 30C
Fig. 4. Effect of curing time on the longitudinal displacement of bonded CFRP strips
exposed to different temperatures during sustained loading.

67

3. NUMERICAL PROGRAM
3.1. Creep Constitutive Model
A power-law model was used to describe the creep behavior of the epoxy used in the CFRP
strengthening system. This model is only valid in cases where the stress state remains
essentially constant and low in relation to the yield strength of the polymer, as is assumed
for the sustained loading stage of the pull-off specimens. The time hardening form of the
power-law model can be written as follows:
=Aqntm

(1)

Based on the experimental creep test results, the uniaxial equivalent deviatoric creep strain
rate of the epoxy, (min.-1) was modeled in terms of the uniaxial equivalent deviatoric
stress, q (Pa), and time, t (min.) using Equation 1 [7]. The creep model parameters A and m
were obtained by curve-fitting the model to the experimental creep strain data (Table 3).
The stress exponent n is assumed to be 1 for linear viscoelastic behavior.

A
n
m

Table 3. Creep law parameters used in the finite element analysis.


RT-7d
RT-90d
HT-7d
HT-90d
1.2e-10
1.0e-12
4.3e-10
1.0e-12
1
1
1
1
-0.750
-0.325
-0.950
-0.370

3.2. Finite Element Modeling


Time-dependent nonlinear finite element analyses of the pull-off specimens during
sustained loading were performed using ABAQUS 6.9-2. The finite element mesh used to
model the pull-off specimen is shown in Fig. 5. The specimen was modeled using a 2-D
plane strain analysis. Concrete, epoxy, damage band (concrete-epoxy interface) and CFRP
were modeled using 4-node bilinear elements (CPE4R). Tie contacts were used between
meshes with different element sizes. Normal and shear stresses can be determined along the
entire tie interaction. A plastic damage model was used to predict the constitutive behavior
of the damage band. Details of the formulation can be found elsewhere [2]. The CFRP strip
is assumed to behave linear-elastically in tension up to the failure stress.

Fig. 5. Finite element mesh used for modeling the pull-off specimen.

68

3.3. Numerical Results


The FE model described in this paper was built assuming that creep deformation occurs
only in the epoxy. One of the objectives of the modeling effort is to determine the
suitability of this assumption for the given experimental conditions. Another objective is to
evaluate the capability of the power-law model in simulating creep of bonded CFRP strips.
Figure 6 shows the experimental and numerically simulated longitudinal displacements of
point D6 as functions of time. It can be observed that the FE model captures the general
trend of the time-dependent deformations for all temperatures and epoxy ages, except for
the 7-day cured specimens with sustained loading at 20C (RT-7d). As mentioned earlier,
additional investigation needs to be done on these specimens to determine whether or not
microcracking contributed to the anomalous results. The best prediction was found for the
7-day cured specimen with sustained loading at 30C (HT-7d).
0.5

Exp-1
Exp-2
Exp-3
FEM

0.4
0.3

Displacement (mm)

Displacement (mm)

0.5

0.2
0.1
0
0

10

20

30

Time (day)

40

0.3
0.2
0.1
0
0

50

Exp-1
Exp-2
Exp-3
FEM

0.4

10

20

(a) RT-7d
Exp-1
Exp-2
Exp-3
FEM

0.4
0.3
0.2
0.1
0
0

10

20

30

Time (day)

40

50

(b) RT-90d
0.5

Displacement (mm)

Displacement (mm)

0.5

30

Time (day)

40

0.3
0.2
0.1
0
0

50

(c) HT-7d

Exp-1
Exp-2
Exp-3
FEM

0.4

10

20

30

Time (day)

40

50

(d) HT-90d

Fig 6. Longitudinal displacement of target point D6 as function of time from experimental


(Exp) and numerical (FEM) results.
5. CONCLUSIONS

Results from this investigation indicate significant effects of curing time and temperature
during loading on the time-dependent creep behavior of CFRP composite bonded to
concrete. The higher temperature resulted in larger creep displacements for either curing
time. Longer curing time resulted in less creep deformation only in the specimen loaded at
room-temperature. The epoxy creep model used in the numerical simulations captures the
general trend of the time-dependent CFRP deformations for the temperature and curing
69

times evaluated, with the exception of the 7-day cured specimens with sustained loading at
20C. The possibility of microcracking in the latter specimens needs to be investigated in
future research.
6. ACKNOWLEDGEMENT
This paper is based upon work supported by the National Science Foundation under Grant
No. CMMI-0826461 and an REU supplemental grant. The authors would like to
acknowledge the help provided by Jeff Flood during the experimental work.
7. REFERENCES
1.
2.
3.
4.
5.
6.

7.

ACI Committee 440. 2008. Guide for the Design and Construction of Externally
Bonded FRP Systems for Strengthening Concrete Structures. ACI 440.2R-08.
American Concrete Institute, Farmington Hills, MI, 76 p.
Coronado, C.A., and Lopez, M.M. 2010. Numerical Modeling of Concrete-FRP
Debonding Using a Crack Band Approach. Journal of Composites for Construction,
ASCE, 14(1): 1121.
Gullapalli, A., Lee, J.H., Lopez, M.M., and Bakis, C.E. 2009. Sustained Loading and
Temperature Response of FRP-Reinforced-Polymer-Concrete Bond. Transportation
Research Record, 2131: 155162.
Coronado, C.A., and Lopez, M.M. 2007. Damage Approach for the Prediction of
Debonding Failure on Concrete Elements Strengthened with FRP. Journal Composites
for Construction, 11(4): 391400.
Chen, J.F., and Teng, J.G. 2001. Anchorage Strength Models for FRP and Steel Plates
Bonded to Concrete. Journal of Structural Engineering, 127: 784791.
Jaipuriar, A., Flood, J., Bakis, C.E., and Lopez, M.M. 2011. Effects of Aging on
Behavior of Epoxy used in FRP Structural Strengthening. Proc. 4th Intl. Conf.
Durability and Sustainability of Fibre Reinforced Polymer (FRP) Composites for
Construction and Rehabilitation, CDCC 2011, eds: Benmokrane, B., El-Salakawy, E.
& Ahmed, E., 20-22 Jul. 2011, Quebec City, Quebec, Canada, 8 p.
Shames, I.H., and Cozzarelli, F.A. 1992. Elastic and Inelastic Stress Analysis, PrenticeHall, New Jersey, USA, 722 p.

70

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

BOND BEHAVIOUR OF GFRP BARS TO CONCRETE SUBJECTED


TO FATIGUE LOADING AND FREEZE-THAW CYCLES
J.R. Alves1, A. El-Ragaby2 and E.F. EL-Salakawy3
1

Structural Engineer, Manitoba-Hydro, Winnipeg, MB, Canada


Assistant Professor, Dept. of Civil & Env. Eng., University of Windsor, Windsor, ON, Canada
3
Associate Professor and CRC, Dept. of Civil Eng., University of Manitoba, Winnipeg, MB, Canada
2

ABSTRACT

Due to the mismatching thermal properties between FRP bars, particularly glass, and
concrete, bond of FRPs to concrete is susceptible to more damage as a result of temperature
changes. Whether FRP-reinforced concrete elements can be designed to utilize the full
tensile strength of FRP bars or not depends largely on their bond characteristics to concrete.
Currently, few to none experimental test data are available on the bond behaviour of FRP
bars in concrete elements under different loading and environmental conditions. Therefore,
this research program is designed to investigate experimentally the durability of FRP bond
to concrete elements subjected to different combinations of loading and environments. An
FRP-reinforced concrete specimen was developed to apply axial-tension fatigue or
sustained loads to GFRP bars eccentrically located within the concrete environment. Test
specimens were subjected simultaneously to the dual effects of 250 freeze-thaw cycles (FT)
and sustained load (SUS) followed by 1,000,000 fatigue load cycles. A total of six test
specimens were constructed and tested. The test parameters included bar diameter and
concrete cover. After conditioning, each test specimen was sectioned to two replicates
(halves) for pull-out test. Another series of twelve unconditioned standard pull-out
specimens were constructed and tested as control. It is concluded that freeze-thaw cycles
along with sustained load followed by fatigue loads resulted in significant decrease in the
bond strength compared to the unconditioned specimens.
1. INTRODUCTION
Glass fibre-reinforced polymer (GFRP) bars are more attractive to the construction industry
due to its lower cost compared to other types of FRP materials. In North America, bridge
deck slabs and parking garages are prime examples where GFRP reinforcements are
currently being used. These elements are continuously subjected to wide combinations of
fatigue and/or sustained load from traffic as well as thermal loads due to freeze-thaw cycles
and temperature fluctuations. Such environmental and loading conditions may affect the
mechanical properties of the concrete itself, the FRP reinforcement, and interface or bond
between them. For reinforced concrete structures, the transfer of stresses between the
71

concrete and the reinforcement, both at the serviceability and at the ultimate limit states, is
strongly dependent on the quality of bond. Therefore, bond of FRP reinforcing bars to
concrete elements subjected to thermal and mechanical loads might be a critical design
factor that needs to be investigated.
Research work done to date on the bond characteristics of FRP bars in concrete investigates
mainly the bond strength under static/monotonic load. Currently, few experimental test data
is available regarding the effects of fatigue loading cycles with constant and variable stress
ranges, impact loads as well as environmental effects such as, freeze-thaw cycles, wet-dry
cycles, high and low temperatures. The bond behaviour of FRP bars is influenced by many
factors such as adhesion, bar diameter, bar surface pattern and shape. Weber (2005) and
Esfahani et al. (2005) noticed the increase in bond strength with the increase of concrete
cover up to a value of 2.0db, where db is the bar diameter. Tepfers et al. (1998) and Galati
et al. (2005) found that concentric placement of the bar reduced the bond strength by
approximately 30% to 50% compared to eccentric specimens with 2.0db concrete clear
covers. On the other hand, Benmokrane et al. (1996) and Achillides and Pilakoutas (2004)
reported that larger size bars produce smaller bond stresses and they also lose their adhesive
bond earlier than small size bars (in case of sand-coated GFRP bars). Therefore, in order to
achieve the same bond stress, larger bar diameters need longer embedment lengths.
Furthermore, different loading conditions also were found to affect the bond quality of FRP
bars to concrete. Bakis et al. (1998) investigated the influence of cyclic load on the bond
behaviour of GFRP-reinforced beams. FRP bars with difference surface patterns were
investigated. The authors reported that the cumulative slip increased with increase in the
peak load or with increase in the number of cycles regardless of the surface pattern.
Moreover, Katz (2000) observed 19% to 80% reduction in the bond strength of GFRP in
concrete specimens depending on the surface conditions. The GFRP bars with smooth
surface showed the highest loss in bond strength, while bars containing sand-coating and
helical wraps presented the lowest loss. He and Sun (2006) found that pullout specimens
with GFRP bar containing helically winded glass strands subjected to fatigue loading
showed up to 40% smaller bond resistance and up to 6 times higher peak slip than those
without any loading history.
2. THE EXPERIMENTAL PROGRAM
2.1 Test Specimens
This experimental program is divided into two series: (C) Control Static pullout and (I)
Freeze-thaw cycles along with sustained loads followed by fatigue loading. One of the main
challenges of this research was to develop a test specimen that meets the following
requirements:
Apply axial tension sustained or fatigue loading on FRP bars eccentrically located
within concrete with different concrete cover thickness;
Sustain large number of fatigue load cycles without premature failure of the anchor;
Can be moved easily and placed inside the environmental chamber for the
environmental conditioning;
Allow to perform monotonic pull-out test after conditioning.

72

An FRP-reinforced concrete specimen for Series I was developed to fulfill these


requirements. The specimen consists of three concrete blocks: the test portion (middle
block) and the two end anchor blocks. The three portions were connected together by the
GFRP bar as shown in Fig. 1. The middle block had 200-mm square cross section to meet
the requirement of standard pull-out test specified by the Canadian Standard CSA-S806-02
(CSA-2002). The length of the middle block was 405-mm such that after conditioning, each
specimen was split into two 200-mm cube replicates for pullout test (considering a 5-mm
clearance for the cutting saw). The GFRP bar was always eccentric in the middle block,
while the end anchors were always adjusted to be concentric with respect to the GFRP bar.
Each anchor contained six 16 mm-diameter threaded rods placed perpendicular to the axis
of the GFRP bar to apply the specified loads.
Series C contained 12 unconditioned control specimens to assess the bond characteristics of
the used GFRP bars as well as to provide a datum for the other test specimens using a
standard procedure provided by the CAN/CSA-S806-02 (CSA 2002). The control specimen
consists of a 200-mm concrete cube with a 5db embedment length with eccentric placement
of the bar. Steel pipes were used as anchors, and were cast with expansive grout prior to the
casting of the concrete specimens.

(a) Plan view of test specimen

(b) Photo for the constructed test specimen

Fig. 1. Details of test specimen for Series I


The test parameters included the GFRP bar diameter and the concrete cover. Two bar
diameters, No.16 and No. 19 GFRP bars (15.9 and 18.9 mm in diameter, respectively) were
investigated. Also, three different clear concrete covers of 1.5 db, 2.0 db, and 2.5 db were
studied. These parameters resulted in having a total of 12 specimens in Series I and II; six
specimens each. In each series, the test parameters were bar diameter and concrete cover
where each group was subjected to different conditioning scheme. Table 1 gives the details
of the test program.
73

2.2 Materials Properties


The specimens were constructed using normal weight concrete with an average 28-day
compressive strength of 50 MPa. Since the concrete was cast on several stages, it was
mixed in the McQuade Heavy Structures Laboratory at the University of Manitoba using a
mechanical mixer to obtain consistent concrete properties. At least twelve 100 by 200 mm
and six 150 by 300 mm cylinders were prepared for each concrete batch and cured under
the same condition as their reference specimen to evaluate the compressive and tensile
strength of concrete at the test date.
Table 1. Test matrix and details of the experimental program
Specimens

Name*

Sustained load

Freeze-thaw cycles

Fatigue load

C.16.15
C.16.20
C.16.25
No
No
No
conditioning
conditioning
conditioning
C.19.15
C.19.20
C.19.25
S.FT.F.16.15
S.FT.F.16.20
1,000,000 cycles
S.FT.F.16.25
Series I
30% of GTS
250 Cycles
at 1.5 Hz and 25%
S.FT.F.19.15
of GTS
S.FT.F.19.20
S.FT.F.19.25
Notes: GTS is the guaranteed tensile strength of the GFRP bar.
* C refers to the control specimens; S stands for sustained load, FT freeze-thaw cycles, F
fatigue load, the first two numbers represent the bar diameter (No.16 or No.19), the following
two numbers stands for the concrete clear cover (1.5, 2.0 or 2.5db). So, for example,
S.FT.F.16.15 stands for a specimen subjected to sustained load and freeze-thaw cycles followed
by fatigue loading, considering bar diameter No. 16 mm and concrete clear cover of 1.5db.
Series C

The GFRP bars used in this study (V-RODTM) had a sand-coated surface and were made of
continuous longitudinal fibres impregnated in a thermosetting vinyl-ester resin with a fibre
content of 73% by weight (Pultrall Inc. 2007). Table 2 lists the mechanical properties of
the reinforcing bars as determined by tensile tests on representative sample in accordance
with CAN/CSA-S806-02 (CSA 2002).
2.3 Loading and Environmental Conditioning Schemes
Specimens of series I were subjected to the dual effects of sustained load and freeze-thaw
cycles simultaneously followed by fatigue load cycles. A sustained load of 30% of the
guaranteed tensile strength of the GFRP bar (GTS) was selected (CSA 2002).
Consequently, the applied sustained loads were 42 and 54 kN for No.16 and No. 19 GFRP
bars, respectively. Moreover, the fatigue loading consisted of repeated tensile stress cycles
(tension-tension) with peak loads of 25% of the GTS according to the limits permitted by
the CAN/CSA-S6-06 (CSA 2006). Therefore, the peak tensile load for a fatigue cycle was
74

35 and 45 kN for No.16 and for No. 19, respectively. The lowest tensile load was set to 5
kN for all fatigue cycles. The specimens were conditioned to 1,000,000 cycles in a
sinusoidal wave.
Table 2. Properties of GFRP bars
Bar
Size

Bar
Diameter
(mm)

No.16
No.19

15.9
18.9

Tensile
Modulus of
Elasticity
(GPa)
48.2
47.6

Ultimate
Tensile
Strength
(MPa)
751
728

Ultimate
Tensile
Strain
(%)
1.56
1.53

Coefficient of Thermal
Expansion (at 10-6/oC )
Longitudinal
6.4
6.0

Transverse
29.1
29.0

The freeze-thaw conditioning was designed and applied in accordance with ASTM standard
specification C66603 (ASTM 2003). According to this standard, to simulate the freezethaw effect, the temperature inside the concrete at the level of reinforcement should ramp
between -18 and +4oC allowing for 67% of the cycle duration under freeze conditions.
After several attempts, it was concluded that ramping the environmental chamber
temperature from -25oC to +15oC and maintain the freezing period (at -25oC) for five hours
and the thawing period (at +15oC) for 3 hours satisfied the ASTM C666-03 requirement. A
total of 250 cycles were applied in approximately 4 months at a rate of 2 cycles/day. The
relative humidity was forest to 50% during freezing and 95% during thawing.
2.4 Test Set-up and Instrumentations
The sustained load was applied to the specimens of Series I through a prestressing bed. A
set of steel plates were used to apply the load. Steel angles and two 38-mm diameter thread
rods were used to hold the anchors in place, thus sustaining the load. To avoid relative
movement between concrete and the steel angles, nuts were used in both sides of the
angles. This mechanism allows moving the specimen in and out of the environmental
chamber as well as adjusting the applied load by turning the locking nuts toward or against
each other to maintain constant sustained load. Figure 2 shows a photo for the sustained
load test set-up. A total of twelve 6-mm long electrical strain gauges were attached to the
GFRP bars at the middle of the test block. During prestressing, the load was monitored by a
load cell, while the strains in the GFRP bars were monitored through the strain gauge.
Moreover, thermocouples were attached to the GFRP bars in two specimens: S.FT.F.16.15
and S.FT.F.19.25, which represented the specimens with the smallest and largest concrete
covers. These thermocouples were attached to verify the temperature inside the specimens
during the freeze-thaw cycles.
A universal 1000-kN MTS machine was utilized to apply either the fatigue cyclic loading
or the monotonic pull-out loading. The fatigue load cycles were applied to the test
specimens of Series I in sinusoidal wave at rate of 1.5 Hz under load-controlled conditions.
The monotonic pull-out load was also applied at a load-controlled rate of 10 kN/min in
accordance with the standard test method of CAN/CSA-S806-02 (CSA 2002). A high
accuracy LVDT was used to record the free-end slip of GFRP bars during the pull-out test.
Figure 2 shows details of the fatigue loading and pull-out tests set-up.

75

(a) Sustained Load

(b) Freeze-Thaw Conditioning

(c) Fatigue loading

(d) Pull-out test

Fig. 2. Details of test set-up for different loading and environmental conditioning
TEST RESULTS AND OBSERVATION

In this section, test results regarding bond strength and slip measurements are presented in
terms of bond stress versus free-end slip for different conditioning schemes. The term slip
represents the relative displacement between the concrete block and the GFRP bar. The
bond stress was considered constant along the embedded length. Such assumption is
acceptable since a short embedment length of 5db was used.
2.5 Effect of Conditioning
Figures 3 and 4 show typical bond-slip relationship obtained from the pull-out tests carried
out on the conditioned specimens. For specimens in Series I, all the conditioned test
specimens showed degradation in bond properties between the GFRP bars and concrete.
This degradation was observed in terms of decrease of the bond strength, increase of the
slip, or both. The bond strength decreased by about 20 to 50% for specimens S.FT.F.19.25
and S.FT.F.16.25 with significant increase in slip in the former specimen when compared
to their unconditioned counterparts. It is clear that the fatigue loading largely degraded the
bond resistance of sand-coated GFRP bars.
2.6 Effect of Concrete Cover
Figures 5 shows a comparison of the bond-slip relationship of test specimens of Series I
with different concrete cover thickness, which seems to play an important role. From Fig. 5,
it is clear that the degree of decrease in the bond strength due to combined freeze-thaw,
sustained loading and fatigue conditioning depends on the concrete cover. Increasing the
concrete cover by about 33% (from 1.5db to 2.0db) and 25% (from 2.0db to 2.5db) increased
the residual bond strength (i.e. decrease the loss in bond strength) by about 15 and 25%,
respectively. On the other hand, all control specimens showed constant increase of about
15% in the bond strength when increasing the concrete cover from 1.5db to 2.0db or from
2.0db to 2.5db. This can be explained as follows. Increasing the concrete cover increases the
confinement pressure on the bar which improves the bond strength. Also, cover thickness
plays an important role in decreasing the temperature gradient effects at reinforcement level
which reduces the thermal effects due to temperature fluctuation. However, it can be
concluded that the effect of fatigue loading is more dominant than that of the freeze-thaw

76

conditioning so that the residual bond strength after conditioning is less than that of the
counterpart control specimen.

12

12

Bond Stress (MPa)

15

Bond Stress (MPa)

15

9
6
Series C
(Control

9
6
Series C (Control)
Series I (S.FT.F)

3
0

0
0

0.04

0.08
0.12
0.16
Free-End Slip (mm)

0.2

0.04

0.08
0.12
0.16
Free-End Slip (mm)

0.2

Fig. 4. Effect of different conditioning


schemes on the bond-slip behaviour of GFRP
bar No. 16

Fig. 3. Effect of different conditioning


schemes on the bond-slip behaviour of
GFRP bar No. 19

Figure 6 shows comparisons of the bond strength of GFRP bars to concrete after
conditioning. It can be noticed that all test specimens of Series I experienced decrease of
bond strength after conditioning compared to their counterparts of Series C. Also, it is
noticeable that the larger the bar diameter, the larger the loss in bond strength after
conditioning. This may be due to the effect of thermal stresses on the expansion of the bar
cross section, which is more pronounced for the larger bar diameter (No. 19). Therefore, the
largest cover, 2.5 db, appears to have resisted the thermal loads and was capable of
providing significant increase in bond resistance. For smaller covers (1.5 and 2.0db), the
enlargement in the bar diameter may have caused damage in the surrounding concrete.
From these figures, it can be concluded that using concrete cover of 2db and 2.5db resulted
in the highest bond strengths for GFRP bars No. 16 and No. 19, respectively, either for the
control or the conditioned specimens.
However, the ISIS Design Manual No 3 (ISIS Canada 2007) and CAN/CSA-S6-06 design
code (CSA 2006) recommend the use of the clear covers of 2.5db for external exposure with
a minimum value of 35 10 mm. The CAN/CSA-S806-02 code (CSA 2002) recommends
clear covers of 3.5db with a minimum value of 40 mm for prestressed elements. Compared
to test results, these provisions are not in good agreement with test results for No. 19 mm
GFRP bars subjected to freeze-thaw cycles and fatigue loads. The ACI 440.1R-06 (ACI
Committee 440) only provides a maximum limit of 3.5db for the clear concrete cover.

77

19.15
16.15

12

19.20
16.20

15

19.25
16.25

Bond strength (MPa)

Bond Strength (MPa)

15

Series C (control)
Series I (S.FT.F)

10

9
6
3

5
0

0
0

0.04

0.08
0.12
0.16
Free-End Slip (mm)

0.2

Fig. 5. Effect of concrete cover on the bondslip behaviour of GFRP bars subjected to
Freeze-Thaw cycles, sustained load, and
fatigue

16.15 16.2 16.25 19.15 19.2 19.25


Test Specimens

Fig. 6. Comparison of the bond strength of


GFRP bars to concrete after being
conditioned under different conditioning
schemes

2.7 Mode of Failure


In this section, the mode of failure was analyzed. Five types of failure could be observed in
this experimental program as shown in Fig. 7. Tastani et al. (1984) reported also similar
failure modes for bond failure of FRP bars. The following presents brief description of the
modes of failure obtained for the three test series:
Pure Pullout Failure; where the bar is pulled out of the specimen without any
splitting/cracking in the concrete.
Concrete Cover Failure; where a crack in the smallest cover is suddenly formed with
immediate loss of bond resistance.
V-Shaped Concrete Cover Failure; in this case, it is thought that a cover failure has
already occurred. The final failure is observed with the v-shaped concrete failure,
Diagonal Failure; probably due to the lack of homogeneity of the concrete, a failure
which should follow the V-shape, deviated and cracked the cover in the lateral direction,
Concrete Block Split; where a crack, which started in the smallest cover, propagates
through the whole concrete block.
In Series C, it was observed that by increasing the concrete clear cover, the failure mode
changed. For specimens with No. 16 bars, the clear cover of 2.0db was enough to ensure the
pure pullout failure, while for No. 19 bar the same concrete clear cover (2.0db) was not
enough to prevent cracking. The concrete deterioration due to freeze-thaw conditioning is
visible in both the specimens and cylinders. The concrete compressive and tensile strength
of the cylinders that were kept inside the environmental chamber during the freeze-thaw
conditioning were 15% less than those of the control ones. In this case, the decrease in the
concrete strength also contributed to the splitting mechanism. For Series I, half of the
specimens failed in pure pullout and half failed by cracking. In some specimens, which
failed by concrete block splitting, the surface print of the sand-coating layer on the concrete

78

interface could be clearly seen This observation indicates that significant deterioration
occurred in the sand-coating layer due to the fatigue loading.

(a) Pure pullout


failure

(b) Concrete
cover failure

(c) V-Shaped
concrete failure

(d) Diagonal concret (e) Concreteblock


splitfailure
cover failure

Fig. 7. Different modes of bond failure


5. CONCLUSIONS
Based on the experimental results, the following conclusions can be drawn:
1. In Series C (control specimens), unconditioned specimens presented very stiff
behaviour with no slip up to failure. It can be observed that by increasing the concrete
clear cover, the failure mode changed. For specimens with No. 16 bars, the clear cover
of 2.0db was enough to ensure the pure pullout failure, while for No. 19 bar the same
concrete clear cover (2.0db) was not enough to prevent concrete cracking.
2. Investigating the results of Series I (freeze-thaw cycles and sustained load followed by
fatigue loading), it can be concluded that the effect of fatigue loading is more dominant
than the freeze-thaw cycles, since all specimens in Series II showed significant loss in
bond strength of at least 23%. For specimens with No. 16 bars, it seems that the effect
of fatigue loading counteracts that of the freeze-thaw cycles. While for specimens with
No. 19 bars and covers of 1.5 and 2.0db, the adverse effects of both conditionings are
added together.
3. It seems that a concrete clear cover of 2.0db is an optimum for sand-coated GFRP bars.
However, if subjected to the effect of freeze-thaw cycles, bar No. 19 presented 25%
higher bond resistance than unconditioned specimens for the larger concrete cover of
2.5db.
6. ACKNOWLEDGEMENTS
The authors wish to express their gratitude and sincere appreciation for the financial
support received from the Natural Science and Engineering Research Council of Canada
(NSERC), through Canada Research Chairs program. Also, the equipment provided by a
Canada Foundation for innovation (CFI) grant is greatly appreciated. The help received
from the technical staff of the McQuade Heavy Structural Laboratory in the department of
civil engineering at the University of Manitoba is also acknowledged.

79

7. REFERENCES
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.

Achillides, Z. and Pilakoutas, K. 2004. Bond behaviour of fibre reinforced polymer


bars under direct pullout conditions. ASCE, Journal of Composites for Construction, 8,
No. 2: 173-181.
Bakis, C., Al-Dulaijan, S., Nanni, A., Boothby, T. and Al-Zahrani, M. 1998. Effect of
cyclic loading on bond behaviour of GFRP rods embedded in concrete beams.Journal
of Composites Technology and Research, JCTRER, 20, No. 1: 29-37.
Benmokrane, B., Tighiouart, B. and Chaallal, O. 1996. Bond strength and load
distribution of composite GFRP reinforcing bars in concrete. ACI Materials Journal,
93(3): 246-253.
CSA. 2002. Design and construction of building components with fibre reinforced
polymers, CAN/CSA-S806-02, Canadian Standards Association, Toronto, Canada.
CSA. 2006. Canadian highway bridge design code, CAN/CSA-S6-06, Canadian
Standards Association, Toronto, Canada.
Galati, N., Nanni, A., Dharani, L., Focacci, F. and Aiello, A. 2005. Thermal effects on
bond between FRP rebars and concrete. Elsevier Composites: Part A: 1223-1230.
He, Z. and Sun, Y. 2006. Experimental study on bond behaviour of GFRP rod in
concrete subjected to stress reversal. Federation Internationale du Beton, FIB, Session
10. ID 10-29.
ISIS Canada. 2006. Durability of fibre reinforced polymers in civil infrastructure,
Technical Report, The Canadian Network of Centres of Excellence on Intelligent
Sensing for Innovative Structures, ISIS, Winnipeg, Manitoba, Canada.
ISIS Canada. 2007. Reinforcing concrete structures with fibre reinforced polymers,
ISIS-Manual No.3, version 2, The Canadian Network of Centres of Excellence on
Intelligent Sensing for Innovative Structures, ISIS, Winnipeg, Manitoba, Canada.
Katz, A. 2000. Bond to concrete of FRP rebars after cyclic loading. ASCE, Journal of
Composites for Construction, 4(3): pp. 137-144.
Koller, R., Chang, S. and Xi, Y. 2006. Fibre-reinforced polymer bars under freezethaw cycles and different loading rates. Journal of Composite Materials, 41(1): 5-25.
Mashima, M. and Iwamoto, K. 2004. Bond characteristics of FRP rod and concrete
after freezing and thawing deterioration. ACI Structural Journal, 138: 51-70.
Pulltral Inc. 2007. V-ROD-Product guide specification. Pultral inc, Thetford Mines,
Quebec, Canada.
Shahidi, F., Wegner, L. and Sparling, B. 2006. Investigation of bond between fibrereinforced polymer bars and concrete under sustained loads. Canadian Journal of Civil
Engineering, 33:.1426-1437.
Tastani, S., Pantazopoulou, S. and Karvounis, P. 1984. Local bondslip characteristics
of GFRP bar. ACI Structural Journal, 2(230):1481-1496.
Tepfers, R., Hedlund, G. and Rosinski, B. 1998. Pull-out and tensile reinforcement
splice tests with GFRP bars. Second International Conference on Composites in
Infrastructure, ICCI-98, Tucson, Arizona, 14p.
Won, J. and Park, C. 2006. Effect of environmental exposure on the mechanical and
bonding properties of hybrid reinforcing bars for concrete structures. Journal of
Composite Materials, 40(12): 1063-1074.

80

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

DURABILITY CHARACTRISTICS OF NEW GFRP DOWELS FOR


CONCRETE PAVEMENT
M. Montaigu1, M. Robert2 and B. Benmokrane3
1

Master Student, Department of Civil Eng., University of Sherbrooke, Quebec, Canada


Postdoctoral Fellow, Department of Civil Eng., University of Sherbrooke, QC, Canada
3
Professor, Department of Civil Engineering, University of Sherbrooke, Quebec, Canada
2

ABSTRACT
Dowel bar commonly used for load transfer in concrete pavement slab is traditionally made
from steel. However, once the steel dowel bar corrodes, it may cause faults, such as binding
due to lockout of the dowel bar in concrete pavement, level differences resulting from
spalling or decreased efficiency of load transfer. To solve this problem, many studies on the
performance and feasibility of FRP dowel bars are being undertaken. In particular, glass
fibre-reinforced polymer (GFRP) dowel bar is now considered as a potential solution.
However, GFRP dowels could be damaged by the capillary water of concrete, which
typically has high alkalinity. This study is conducted through a research collaboration
project between the University of Sherbrooke and the Ministry of Transportation of Quebec
(MTQ, Pavement Division). One of the main objectives of this research project was to
evaluate the durability characteristics of new generation of GFRP dowel bars manufactured
using vinylester resin, in simulated environmental conditions experienced by concrete
pavement. An accelerated aging test method was applied to evaluate the long-term
durability of the GFRP dowel bar. The results of this research have contributed to introduce
this type of GFRP dowel bars with enhanced durability characteristics for concrete
pavement applications in the province of Quebec.
1. INTRODUCTION
FRP composites, mainly based on thermoset polymers and glass or carbon fibres, are being
used in infrastructures exposed to harsh conditions involving de-icing salts or marine
environments. A typical dowel bar is more vulnerable to corrosion when it is used to
transfer loads in concrete pavement due to the penetration of water by the joint opening [1].
When corrosion occurs in a dowel bar, freezing could be caused by the expansion, followed
by fractures due to the curling of the concrete pavement slab. Ministry of Transportation of
Quebec (MTQ) has encountered these types of problem with steel reinforced jointed
concrete pavement. More than 350 km of jointed concrete pavement have been rebuilt since
1994 in the province of Quebec. To minimise the corrosion problems, the MTQ is actually
using epoxy-coated steel dowels of 25.1 to 38.1 mm in diameter in the design of new
81

concrete pavement slabs of 225 to 325 mm in thickness. To solve corrosion problems of


metallic dowels (conventional steel, epoxy-coated, etc.), GFRP dowel bars could be
considered now as a feasible solution [2, 3] with high industrial productivity [4, 5].
The MTQ is looking to improve the long-term performance, durability, and safety of
jointed concrete pavement using GFRP dowels. However, long-term durability
performance of GFRP dowels, such as in high alkalinity environment, needs to be
investigated [6].
This study investigates the durability of GFRP dowels using accelerated ageing alkaline
tests at different temperatures and period of time. GFRP specimens were immersed at
different temperatures (up to 60 C.) in alkaline cement Portland solution for a period of 180
days, then tested under transverse direct shear. Based on prediction modelling, long-term
transverse shear strength of the tested GFRP dowels was determined.
2. EXPERIMENTAL PROGRAM
2.1 Material
GFRP dowels made of continuous high strength E-glass fibres impregnated in vinylester
(VE) resin using the pultrusion process were used in this study. Durability tests were
performed on several diameters from 25.4 to 44.5 mm. The initial properties (reference) of
the GFRP dowels were determined by preliminary tests. As there was no effect of the
diameter on the degradation of dowels (less than 2%), the durability study reported in this
paper only presents results for 34.9 and 38.1 mm-diameter dowels. Note that these two
diameters are the most employed diameters in roads infrastructure. The reference
mechanical and physical properties of the tested 34.9 and 38.1 mm-diameter GFRP dowels
are summarized in Table 1.
All GFRP dowels were cut into 300 mm lengths so that the direct shear test can be
performed using ACI 440.3R-04. The bars were divided into two series; 1) the
unconditioned reference samples; 2) conditioned samples immersed in alkaline solution.
2.2 Conditioning of the GFRP Dowels
Accelerated ageing tests were conducted according to the method DBT-2 Recommended
FRP Dowel Bar Test Protocol, proposed by Market Development Alliance of the
Composite Institute [7], in accordance with the ASTM C 581 standard. Cement extract
solution, which is a real representative solution of concrete pores water, was prepared by
mixing commercial Portland cement type 30 (highest rate of CaO) with tap water to get a
pH of 12.60, measured at the beginning of the test and maintained constant throughout the
conditioning period with Ca(OH)2 and cement. Specimens were immersed at three different
temperatures (23, 50 and 60C). Specimens were place in wood containers waterproofed
with a high density polyethylene film and hermetically closed to avoid excessive
evaporation and changes of the pH solution (Figure 1). For each diameter and resin type of
dowel, six specimens were removed from solution and tested under short beam shear tests
after 30, 60 and 180 days at 23, 50 and 60C of conditioning to measure the mechanical
properties of aged dowels.
82

Mechani

Physical

Table 1. Mechanical and physical properties of 34.9 and 38.1 mm diameter GFRP dowels
Property
VE-based GFRP dowels
34.9 mm
38.1 mm
Fibre content (%)
80.7
80.6
Cure ratio (%)
100
100
Tg (C)
124
123
Moisture uptake (%)
0.06
0.07
Relative density
2.12
2.14
-6
-1
LTE long. (x10 C )
7
7.6
-6
-1
LTE transv. (x10 C )
23.5
24
Direct shear strength (MPa)
184
173
Short beam shear strength (MPa)
61.1
53.9
4 points flexural strength (MPa)
1210
1077
Flexural modulus of elasticity (GPa)
50.3
51.6

Fig. 1. GFRP dowels ready for conditioning


2.3 Direct Transverse Shear Tests
Direct shear tests were conducted to characterize the GFRP dowels according to ACI
440.3R-04 guide Test Method B.4. The setup consists in a 230 x 100 x 110 mm steel base
equipped with lower blades spaced of 50 mm face to face, allowing the double direct shear
of the specimen by an upper blade, as shown in Figure 2. Six unconditioned specimens of
300 mm long were tested as reference for each type of dowel at the laboratory conditions
using MTS 810 testing machine equipped with a load cell of 500 kN. The displacement
speed was 1.5 mm/min giving between 30 and 60 MPa/min, until the failure of the
specimen. The direct shear strength is given by the equation 1.
P
(1)
u s
2A
Where u is the direct shear strength (MPa); Ps is the failure load (N); and A is the specimen
section (mm).
83

Fig. 2. Direct shear test setup


2.4 Prediction of Long-Term Behavior
Arrhenius concept [8] was used for the prediction of long-term behaviour of GFRP dowels.
The equation 2 expresses the Arrhenius relation, in terms of the degradation rate [9].
k Ae

Ea
RT

(2)

Where k is the degradation rate; A is a constant relative to the material and degradation
process; Ea is the energy of activation of the reaction; R is the universal gas constant; and T
is the temperature in C. The primary assumption of this model is that only one dominant
degradation mechanism of the material operates during the reaction and that this
mechanism will not change with time and temperature during the exposure.
2.5 Differential Scanning Calorimetry (DSC)
Twelve-milligram to 15-milligram specimens from both unconditioned and aged samples
were sealed in aluminum pans and analyzed in a TA Instruments DSC Q10 calorimeter
equipped with a refrigerated cooling system. Analysis was conducted in modulated DSC
mode. Specimens were heated from 25C to 195C at a rate of 5oC/min. Glass transition
temperature was determined by DSC for each specimen in accordance with ASTM D 1356
standard. Two scans were performed for each specimen. The first scan is useful to
determine the difference of Tg between reference and conditioned specimens. If a decrease
of Tg is observed for conditioned samples, this is an indication of plasticizing effect or
chemical degradation. The second scan gives information about the mechanism of
degradation and if it is irreversible.

84

2.6 Fourier Transform Infrared Spectroscopy (FTIR)


Fourier transform infrared spectroscopy (FTIR) spectra were recorded using a Nicolet
Magna-550 spectrometer equipped with an attenuated total reflectance device. Fifty scans
were routinely acquired with an optical retardation of 0.25 cm to yield a resolution of 4
cm1. 25.4 mm diameter dowels were investigated after 180 days conditioning at 60C.
2.7 Scanning Electronic Microscopy (SEM)
Scanning electron microscopy (SEM) observations and image analysis were performed to
observe the microstructure of vinylester-based dowels specimens before and after aging in
alkaline solution during 180 days at 60oC. All specimens observed in the SEM were first
cut, polished, and coated with a thin layer of gold-palladium by a vapor-deposit process.
After coating the surfaces, microstructural observations were performed on a JEOL JSM840A SEM. These observations were conducted to see the potential degradation of polymer
matrix, glass fibers, or interfaces, if any.
3. EXPERIMENTAL RESULTS
3.1 Direct Shear Strength
Table 2 shows the direct shear strength of vinylester-based dowels after different
conditionings in alkaline solution. Figure 3 shows the retention of the direct shear strength
after 180 days of ageing of the GFRP dowels at various temperatures. As shown in the
Figure 3, vinylester-based dowels were slightly affected by ageing in alkaline solution with
a retention of more than 90% of the initial shear performance after 180 days at 60C in
alkaline solution (pH=12.60). All specimens have kept elastic behaviour until the shear
failure of the fibres (Figure 4). It can be seen that the degradation rate between 23C and
50C is nearly the same that between 50C and 60C, characterizing the exponential effect
of the temperature on mechanisms of degradation.
Table 2. Direct shear properties of aged GFRP dowels
Time of
Temperature
Mean Shear
Standard deviation
o
immersion (days)
( C)
Strength (MPa)
(MPa)
35 mm 38 mm
35 mm
38 mm
0
23
184.3
173.4
1.81
3.19
23
182.1
177.7
2.67
1.24
30
50
180.0
175.7
2.43
2.54
60
178.1
171.0
3.65
1.87
23
180.3
181.9
1.97
1.93
60
50
175.7
177.6
4.17
1.84
60
172.4
168.6
1.85
0.86
23
178.9
176.5
1.87
2.14
180
50
172.7
169.8
3.11
0.73
60
167.5
157.7
3.65
3.27

85

Fig. 3. Direct shear strength retention of GFRP dowels after conditioning in alkaline
solution at 60oC

a)

b)
Fig. 4. Typical mode of failure of 34.9 mm GFRP dowels tested under shear: a)
longitudinal view, b) cross-sectional view
3.2 Effect on Polymer Matrix
Table 3 gives the values of Tg before and after aging in alkaline solution during 180 days at
60oC. No significant effect of aging was observed on vinylester-based dowels. It can also
be seen that no shift of Tg was measured during the second run, leading to the conclusion
that the initial cure ratio of the vinylester resin was very high.
A FTIR analysis of unconditioned dowel and specimens aged in alkaline solution during
180 days at 60oC was conducted (Figure 5). The most interesting region of the FTIR
spectra is located between 3300 cm-1 and 3600 cm-1, which corresponds to the stretching
mode of the hydroxyl groups of the vinylester resin. When hydrolysis reaction occurs, new
hydroxyl groups are formed and the corresponding infrared band increases. Changes in the
peak intensity are quantified by determining the ratio of the OH- peak to the carbonhydrogen stretching peak of the resin, which is not affected by the conditioning. The
experimental ratios of the OH peak to the carbon-hydrogen stretching peak of the core and
86

the surface of vinylester-based dowel immersed in alkaline solution for 180 days at 60oC
were 0.44 and 0.54, respectively, compared to 0.49 for unconditioned samples. The
hydroxyl peak did not show any significant changes. This observation lead to the
conclusion that no chemical degradation of the vinylester resin occurred during the
immersion of the dowels in alkaline solution at 60C for 180 days.
Table 3. Tg of reference and aged dowels
Tg run 1 (oC)
35 mm
38 mm
Reference
123
124
o
180 days in alkaline solution at 60 C
122
123
Conditioning

Tg run 2 (oC)
35 mm
38 mm
124
123
123
122

Fig. 5. FTIR spectra for unconditioned and aged samples


3.3 Microstructural Observations
3.3.1 External Surface
SEM observations of external surface of dowels were performed to investigate the surface
deterioration of the polymer matrix after conditioning in alkaline solution. Figure 6 presents
micrographs of the surface of reference and aged dowels.

a)
b)
Fig. 6. Micrographs of the dowels surface for: a) 34.9 mm VE reference dowel; b) 34.9 mm
VE dowel aged in alkaline solution during 180 days at 60oC
87

It can be observed that no degradation of the vinylester matrix has occurred after 180 days
at 60C aging in alkaline solution. No increase of the number or dimensions of the pores
was observed and the surface remains intact without any cracking or microcracking. No
dimension or weight changes of the dowels have been noticed after the accelerated ageing
in alkaline solution.
3.3.2 Micro-structural Effects
Figure 7 presents micrographs for both reference specimens and specimens aged in alkaline
solution during 180 days at 60oC. The visual and microstructural observations showed no
significant damage on vinylester-based dowel after 180 days of immersion in the alkaline
solution at the highest temperature (60oC). Observations of the fibre/matrix interface and of
the microstructure, in general, demonstrate that the conditionings of vinylester-based dowel
in alkaline solution do not affect the microstructural properties of the GFRP dowel (Figure
7b).

(a)
(b)
Fig. 7. Micrographs at the fibre/matrix interface before mechanical tests for: a)VE reference
dowel, b) 34.9 mm VE dowel aged in alkaline solution during 180 days at 60oC
3.4 Service Life Prediction
Long-term predictions of mechanical properties can be made according to the method
DBT-2, based on shear strengths obtained after 30, 60 and 180 days at three temperatures
of conditioning. Simulations are proposed for the vinylester-based dowels of 34.9 mm
diameter. Following the procedure proposed by Bank et al.[10], the natural logarithm of
time to reach a set of levels of normalized performances versus 1/T, expressed as the
inverse of absolute temperature (1000/K), was used to predict the service life at the Mean
Annual Temperature (6.2oC) in Montral, Qubec, Canada. A coefficient of determination
(R2) value close to 1 is desired. However, the ASTM procedures recommend a minimum
value of 0.80 for acceptability and the obtained R2 values are between 0.96 and 0.99.
Predictions are made for direct shear strength retention as a function of time for an
immersion at 6.2C and the general relation between the PR and the predicted service life at
the average temperature of 6.2oC are drawn (Figure 8). It can be seen from Figure 8 that
vinylester-based dowels present a very high durability in concrete pavement environment.
In fact, the predicted service life of vinylester-based dowels immersed in alkaline solution
88

at an isotherm temperature of 6.2oC to reach a PR less than 90% can be estimated to be


infinite.

Fig. 8. Long-term predictions for 34.9 mm dowels


4. CONCLUSIONS
Based on the results of this study, the following conclusions may be drawn:
- The tested GFRP vinylester-based dowels present a very high durability after 180
days of accelerated ageing in alkaline solution at 60C. Shear strength retention of
90 % was measured after 180 days of immersion in alkaline solution at 60oC.
- The microstructural analysis has shown that no damage has occurred to the internal
microstructure or at the fibre/matrix interface for vinylester-based GFRP dowel.
- According to the long-term predictions, the shear strength retention of vinylesterbased GFRP dowels decreased by less than 10% after 200 years. This result clearly
demonstrates the high stability of vinylester-based GFRP dowels in concrete
pavement environment. For steel reinforced concrete pavement, the ACPA
estimates that steel dowel bars can fail in as little as 7 to 15 years depending on
design and location due to corrosion problems.
5. ACKNOWLEDGEMENTS
The authors would like to express their special thanks and gratitude to the Ministry of
Transportation of Quebec, the Natural Science and Engineering Research Council of
Canada (NSERC), the Fonds quebecois de la recherche sur la nature et les technologies
(FQRNT), Pultrall Inc. (Thetford Mines, Qubec) for donating the GFRP dowels, and the
technical staff of the structural lab of the Department of Civil Engineering at the University
of Sherbrooke.
89

6. REFERENCES
1.

Mauricio, M., Cruz, C.J., Jieying, Z.; Harvey, J.T., Monteiro, P.J.M.; Abdikarim, A.
2005. Laboratory evaluation of corrosion resistance of steel dowels in concrete
pavements. Pavement Research Center, Institute of Transportation Studies, University
of California: Davis, Berkeley.
2. Eddie, D., Shalabi, A., and Rizkalla, S. 2001. Glass-Fiber-Reinforced Polymer Dowels
for Concrete Pavements. ACI Structural Journal, 98(2): 201-206.
3. Porter, M. 2002. Assessment of dowel bar research. Final Report, CTRE, Department
of civil and construction engineering, Iowa State University, Ames, USA, 81 p.
4. Highway Innovative Technology Evaluation Center (HITEC). 1998. HITEC evaluation
plan for Fiber reinforced polymer composite dowel bars and stainless dowel bars. Ohio
Department of Transportation, Ohio, USA.
5. Abo-Qudais, S.A. and Al-Qadi, I.L. 2000. Dowel bars corrosion in concrete pavement.
Canadian Journal of Civil Engineering, 27: 1240-1247.
6. Won, J., Cho, Y. and Jang, C. 2006. The Durability of Glass Fibre-Reinforced Polymer
Dowel after Accelerated Environmental Exposure. Polymers & Polymer Composites,
14(7): 719-730.
7. Litherland, K.L., Okley, D.R., and Proctor, B.A. 1981. The use of accelerated aging
procedures to predict the long term strength of GRC composites. Cement and Concrete
Research, 11: 455-466.
8. Market Development Alliance (MDA). 1998. Recommended FRP Dowel Bar
Durability Test Protocol. Dowel Bar Team 2 of the SPI Composites Institute, Harrison,
USA, 26 p.
9. Nelson, W., 1990. Accelerated testingStatistical models, test plans, and data
analyses. Wiley, New York, 601 p.
10. Bank, L.C, Gentry, T.R, Thompson, B.P, and Russel, J.S. 2003. A Model Specification
for Composites for Civil Engineering Structures. Construction and Building Materials,
17(6-7): 405-437.

90

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

EVALUATION OF THE DURABILITY OF CFRP SHEETS AFTER CLIMATIC


EXPOSURE
I. Nishizaki1, P. Labossiere2, K. W. Neale2, M. Demers3 and T. Tomiyama4
1

Team Leader, Public Works Research Institute, Tsukuba, Japan


Professor, Universite de Sherbrooke, Canada
3
Research Engineer, Universite de Sherbrooke, Canada
4
Senior Researcher, Public Works Research Institute, Tsukuba, Japan
2

ABSTRACT
Structural reinforcement with CFRP sheets is a widely used method for strengthening
concrete structures such as girders, decks and piers. While the efficiency of this method is
well documented, there are still questions about its long term durability. Most of the
original studies on the long-term behaviour of CFRP have been done for aircraft or marine
structures. However, many conditions such as materials for these applications,
environmental factors and expected life time are different than those in structural civil
engineering. The authors carried out a series of exposure tests under natural climates
conditions for CFRP sheets in three characteristic locations in the world, Sherbrooke
(Canada), Tsukuba (mainland Japan) and Oogimi (Okinawa). The natural exposure tests
were continued for ten years and the mechanical properties of the specimens from the three
locations, such as tensile strength and in-plane shear, were evaluated at regular interval.
Although the results showed a slight decrease in mechanical properties over the exposure
period, the tested materials exhibit properties that still allow their use after ten years.
1. INTRODUCTION
The use of composite materials is becoming an increasingly popular method of repairing
and strengthening ageing concrete structures around the world. While the efficiency of this
method is well documented, there are still questions about its long term durability. Most of
the previous studies on the durability of CFRP have been done for aircraft or marine
structures [1-4], and many conditions such as materials, environmental factors and expected
life time are different for structural civil engineering applications. The authors carried out a
series of exposure tests under natural climates conditions for CFRP sheets in three
characteristic locations in the world, Sherbrooke (Canada), Tsukuba (mainland Japan) and
Oogimi (Okinawa). Intermediate results were reported previously [5-7], however the
natural exposure tests were originally planned for ten years, and this report shows some
results from the ten year exposed specimens.

91

2. METHODS
2.1 Specimens
The results from two types of specimens are presented herein as follows:
(1) Laminate 0
Table 1 provides a summary of the specimens. Two types of continuous CF fiber sheet
(products A and B) supplied by different manufacturers were used for the specimens. Both
CF fiber sheets were PAN type. Matrix resins were selected from the recommended
products of each manufacturer (an epoxy resin in both cases). Four-ply CFRP
unidirectional laminates, 250 x 300 mm in size with the fibers in the longitudinal direction,
were prepared with each set of materials. Additional laminates with top coating
(acrylurethane resin) were also prepared to evaluate the protective effect of the latter. The
purpose of Laminate 0 specimens was to evaluate the durability of the fiber itself, hence
after exposure, it was tested in tension in the longitudinal direction.

250

Laminate45-A
Laminate45-B
Laminate45-AC

Direction for 1 st & 4 th layer

Laminate system
4-ply, UD

4-ply, [45/-45]s for the


testing direction

Direction for 2 nd & 3 rd layer

Code
Laminate 0-A
Laminate 0-B
Laminate 0-AC

Table 1. Summary of the specimens


Materials
PAN type CF & Epoxy resin (product A)
PAN type CF & Epoxy resin (product B)
PAN type CF & Epoxy resin (product A)
Top coating (Acrylurethane resin)
PAN type CF & Epoxy resin (product A)
PAN type CF & Epoxy resin (product B)
PAN type CF & Epoxy resin (product A)
Top coating (Acrylurethane resin)

250

250

Fig. 1. Outline of Laminate45

92

Unit: mm

(2) Laminate 45
Figure 1 shows the general configuration of Laminate45 specimens. Tensile tests were to be
performed after exposure in the 45-degree direction, hence this specimen has a [45/-45]s
configuration. The purpose of these specimens was to evaluate how much the shear
properties of the laminates would be affected by the condition between fiber and matrix
resin. Table 1 also provides a summary of Laminate45. The three type of the specimens
prepared as Laminate45 were identical in configuration to the Laminate0 series.
Five identical specimens were prepared for each laminate configuration and for each
exposure site. One set of additional specimens was also tested before exposure, in order to
obtain reference values for each configuration.
In addition to the above specimens, three other types of specimens were also prepared for
the exposure test program. They are configuration (3) to (5), as follow:
(3) CF-sheets: CF sheet without matrix resin
(4) Matrix resin plates: Matrix resin plate used for the laminates
(5) Concrete columns: Concrete columns reinforced with laminate-A or B
The results for these specimens are not reported herein.
2.2 Exposure Test
The exposure tests were initiated in 1997 at the three locations identified above. Table 2
shows the main climatic characteristics of each location.
Table 2. Characteristics of the exposure sites
Exposure locations Latitude Annual mean Annual
mean Climate
temperature
rain fall
Sherbrooke
4537N 4.1C
1084mm
Cold and covered
(Quebec, Canada)
with snow in winter
Tsukuba
3567N 13.6C
1505mm
Moderate climate
(Ibaraki, Japan)
Oogimi
2648N 22.4C
2036mm
Subtropical climate,
(Okinawa, Japan)
close to sea shore
The laminate specimens were installed vertically on steel exposure racks using an
aluminum frame. Five specimens of each type were placed in the each of the three exposure
sites. Exposed specimens were scheduled to be recovered after one, three, five, seven and
ten years of exposure. Fig.2, 3 and 4 show the specimens undergoing exposure under
natural conditions at the three exposure sites.

93

Fig. 2. Specimens under the exposure test (Sherbrooke)

Fig. 3. Specimens under the exposure test (Tsukuba)

Fig. 4. Specimens under the exposure test (Oogimi)

94

2.3 Evaluation Methods


All the specimens were recovered according to the planned schedule. Five coupons were
cut out from each retrieved specimen and mechanical properties were examined with the
coupons. The evaluation methods were as follows.
- For Laminate0: tensile test- ASTM D 3039-76 Standard Test Method for Tensile
Properties of Fiber-Resin Composites
- For Laminate45: in-plane shear test- ASTM D 3518M-91 Standard Practice for In-Plane
Shear Stress-Strain Response of Unidirectional Polymer Matrix Composites
3. RESULTS
3.1 Differences between Materials
Fig.5 shows the average tensile properties measured from Laminate0 specimens while Fig.6
shows the results of shear properties from Laminate45 specimens.
1.2
Tensile modulus/Initial values

Tensile strength/Initial values

1.2
1.0
0.8
0.6
0.4
Laminate0-A

0.2

Laminate0-B
2

10

0.8
0.6
0.4
Laminate0-A

0.2

Laminate0-B

0.0
0

1.0

0.0

12

Exposure years

6
8
Exposure years

10

12

a) Tensile strength
b) Tensile modulus
Fig. 5. Tensile properties of the Laminate0 specimens after exposure variation between
products (Average of the three exposure sites)
In-plane shear modulus/Initial values

In-plane shear strength/Initial values

1.2
1.0
0.8
0.6
0.4
Laminate45-A

0.2

Laminate45-B

0.0
0

6
8
Exposure years

10

1.2
1.0
0.8
0.6
0.4
Laminate45-A

0.2

Laminate45-B

0.0
0

12

6
8
Exposure years

10

12

a) In-plane shear strength


b) In-plane shear modulus
Fig. 6. In-plane shear properties of the Laminate45 specimens after exposure variation
between products (Average of the three exposure sites)
95

The tensile strength of the laminates did not show significant decrease even after ten years
of exposure. On the other hand, the shear strength of the laminates showed a slight
decrease. The decrease of shear strength was almost totally attained after three years of
exposure and remained quite stable afterwards. This suggests that the carbon fiber itself has
a good durability after ten years of exposure. However the interface between matrix resin
and carbon fiber appears to show a slight deterioration. The difference between products A
and B is not clear, however Laminate-B seems slightly more sensitive to deterioration than
Laminate-A. The changes in modulus for each case seem to be more significant than
changes in strength, but the difference was not so remarkable in either cases.
3.2 Effects of the Exposure Conditions
Fig.7 and 8 show the results of tensile and shear properties measured from specimens of the
three exposure sites. The data shown are the average values of uncoated specimens
(Laminate0-A and B).
1.2
Tensile modulus/Initial values

Tensile strength/Initial values

1.2
1.0
0.8
0.6
0.4

Sherbrooke
Tsukuba

0.2

Oogimi
2

6
8
Exposure years

10

0.8
0.6
Sherbrooke

0.4

Tsukuba
0.2

Oogimi

0.0

0.0
0

1.0

12

10

12

Exposure years

a) Tensile strength
b) Tensile modulus
Fig. 7. Tensile properties of the Laminate0 specimens after exposure differences between
exposure sites (Average values of the two products)
In-plane shear modulus/Initial values

In-plane shear strength/Initial values

1.2
1.0
0.8
0.6
Sherbrooke

0.4

Tsukuba
0.2

Oogimi

0.0
0

6
8
Exposure years

10

12

1.4
Sherbrooke

1.2

Tsukuba

1.0

Oogimi

0.8
0.6
0.4
0.2
0.0
0

6
8
Exposure years

10

12

a) In-plane shear strength


b) In-plane shear modulus
Fig. 8. In-plane shear properties of the Laminate45 specimens after exposure differences
between exposure sites (Average values of the two products)
96

The results suggest that the deterioration in Oogimi generally seems to be slightly larger
than in other locations. However the difference may not be significant. Considering the
difference between climatic conditions at the exposure sites, the effect of the exposure
condition seems not to be strong in the tested cases. However the tested cases were carried
out in three locations in the world, the condition was not still comprehensive, hence the
strength of the effect of the environmental condition should be evaluated under wider
variations in the future.
3.3 Protective Effect of the Top Coating
Fig.9 and 10 show the protective effect of the top coating. Each data point is the average
value of three exposure sites for Laminate0-A and Laminate0-AC.
1.2
Tensile modulus/initial values

Tensile strength/initial values

1.2
1.0
0.8
0.6
0.4
Laminate0-A
0.2

Laminate0-AC

1.0
0.8
0.6
0.4
La minate0-A

0.2

La minate0-AC

0.0

0.0
0

6
8
Exposure years

10

12

10

12

Exposure years

1.2

In-plane shear modulus/Initial values

In-plane shear strength/Initial values

a) Tensile strength
b) Tensile modulus
Fig. 9. Tensile properties of the Laminate0 specimens after exposure - Effect of the top
coating (Average values of the three exposure sites)

1.0
0.8
0.6
0.4
Laminate45-A
0.2

Laminate45-AC

0.0
0

10

1.2
1.0
0.8
0.6
0.4
La minate45-A
0.2

La minate45-AC

0.0
0

12

Exposure years

6
8
Exposure years

10

12

a) In-plane shear strength


b) In-plane shear modulus
Fig. 10. In-plane shear properties of the Laminate45 specimens after exposure tests - Effect
of the top coating (Average values of the three exposure sites)
There is a previous study that reports that a top coating has a good protective effect against
deterioration of mechanical properties of pultruded GFRP [8]. However, Fig.9 a) and
Fig.10 a) show that the coating does not provide an effective protection for CFRP laminates
97

and does not prevent strength degradation. Fig.9 b) and Fig.10 b) suggest that a top coating
may prevent a modulus reduction; however, this effect seems to be very small.
4. CONCLUSIONS
Outdoor exposure tests were conducted for ten years in three locations in Canada and Japan
for two types of laminates with continuous CF sheets, and the changes of the tensile and inplane shear properties were evaluated.
The tensile strength of the laminates did not show significant decreases even after ten years
of exposure while, on the other hand, the shear strength of the laminates showed a small
decrease. This result suggests that the carbon fiber itself has a good durability after ten
years of exposure. However, the interface between matrix resin and carbon fiber shows a
slight deterioration. The difference between equivalent products from two manufacturers
was not clear.
Despite the important differences in climatic conditions between the exposure sites, it did
not correlate with differences in strength and modulus between specimens.
The anticipated protective effect of a top coating against the reduction of mechanical
properties was not observed for the CFRP laminates tested, whereas such an effect had
clearly been found for GFRP in a previous study.
Overall, although the results showed a slight decrease in mechanical properties over the
exposure period, the products exhibit properties that still allow their use after ten years.
5. ACKNOWLEDGEMENT
This work was supported in Japan by JSPS KAKENHI (21360209) and, until 2009, by the
ISIS Canada Network of Centres of Excellence.
6. REFERENCES
1.
2.
3.
4.
5.

6.

Backer, D. J. 1984. Flight Service Evaluation of Composite Components on Bell 206L


and Sikorsky S-76 Helicopters. Journal of American Helicopter Society, 29 (2): 3-11.
Dexter, H. B. and Backer, D. J. 1984. Worldwide Flight and Ground-based Exposure
of Composite Materials. NASA Conference Publication, NASA-CP-2321, pp. 17-49.
Yamaguchi, T., Momoshima, Y. and Shirota, T. 1998. Long-term outdoor exposure
test for Carbon Fiber Composites, Reinforced Plastics, 44 (2): 56-62. (in Japanese).
Kudo, R., Ben, G., Hojo, H. and Okubo, H. 1999. Durability and undestructive
prediction of bending strength of CFRP under accelerated exposure condition. Journal
of Japanese Society for Composite Materials, 25 (1): 23-29. (in Japanese).
Labossiere, P., Neale, K. W. and Nishizaki, I. 2003. Effect of Different Long-term
Climatic Conditions on FRP Durability, Proceedings of the Sixth International
Symposium on FRP Reinforcement for Concrete Structures (FRPRCS-6), ed. Tan,
K.H., Singapore, 8 10 July, 2: 779-784.
Nishizaki, I., Labossiere, P., and Sarsaniuc, B. 2005. Durability of CFRP sheet
reinforcement through exposure tests. Proceedings of the 7th International Symposium
98

7.

8.

on Fiber Reinforced Polymer Reinforcement for Reinforced Concrete Structures


(FRPRCS-7), ACI SP-230, Kansa City, MO, USA, 5-9 November, pp. 1419-1428.
Demers, M, Labossiere, P., Nishizaki, I., Sarsaniuc, B. and Neale, K.W. 2007.
Durability of CFRP sheets under natural climatic conditions. Proceedings of the Third
International Conference on Durability and Field Applications of Fibre Reinforced
Polymer (FRP) Composites for Construction (CDCC2007), eds. Benmokrane, B. & ElSalakawy, E., Quebec City, Quebec, Canada, 22-24 May, pp. 151-158.
Nishizaki, I., Kishima, T., and Sasaki, I. 2007. Deterioration of mechanical properties
of pultruded FRP through exposure tests. Proceedings of the Third International
Conference on Durability and Field Applications of Fibre Reinforced Polymer (FRP)
Composites for Construction (CDCC2007), eds. Benmokrane, B. & El-Salakawy, E.,
Quebec City, Quebec, Canada, 22-24 May, pp. 159-166.

99

100

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

DURABILITY OF EXTERNALLY BONDED FRP COMPOSITES:


COMPARISON OF ACCELERATED TEST METHODS WITH REAL
TIME EXPOSURE
H.R. Hamilton1, C.W. Dolan2 and J.E. Tanner3
1

Assoc. Prof., Dept. of Civil & Coastal Eng., Univ. of Florida, USA
Prof., Dept. of Civil and Arch. Eng., Univ. of Wyoming, USA
3
Assoc. Prof., Dept. of Civil and Arch. Eng., Univ. of Wyoming, USA
2

ABSTRACT
The durability of CFRP composite reinforcement externally bonded to concrete is affected
by moisture exposure. This paper provides a comparison of accelerated aging of externally
bonded CFRP composites with that of similar specimens exposed to the environment.
Small beam test specimens (100 x 100 x 350 mm) were constructed with bonded CFRP
reinforcement and exposed to both accelerated conditioning using heated water and to real
time weathering. Two extreme environments were chosen for weathering. One was in a
tidal zone in coastal brackish water in Florida, which resulted in wetting and drying. The
other was in Laramie, WY, which exposed the specimens to extreme temperatures, UV, and
freeze-thaw. The accelerated conditioning protocols were submerged exposure in 60C (140
F) water for 60 days and 100% relative humidity at 60 C (140 F) for 60 days. The
specimens were then tested in three-point bending (flexure) to determine the loss in
capacity due to the exposure.
1. INTRODUCTION
The strength of CFRP composites bonded to concrete deteriorates when exposed to
hygrothermal conditions. The deterioration occurs principally on the bond plane, which is
the interface between the epoxy adhesive and the concrete. The exact mechanism of the
deterioration is not well known; however deterioration is accelerated under elevated
temperature, immersion in water or sustained loads. Tests on a several CFRP systems
indicated that the procedures in this report were capable of differentiating a wide range of
responses of deterioration of CFRP systems subjected to the test procedures.
Environmental exposure conditions and test procedures assure uniform reporting and
evaluation of durability strength reduction factors to field application environments. The
methodology presented in this report allows engineers to obtain standardized comparisons
of alternative CFRP systems, apply the strength reduction factors to design, and to make
decisions on the applicability of a CFRP system for a given environment.
101

2. RESEARCH PROGRAM
The research reported in this paper was part of a larger research project examining the
durability of externally bonded CFRP composites. As part of the research a literature
review was conducted regarding the various potential mechanisms of degradation that
might affect externally bonded composites. It was found that moisture is the most likely
candidate. For bond critical applications, moisture degradation of the bond line can be the
most serious degradation mechanism. Consequently, the bulk of the research program
focused on the development of a simple, convenient bond test and the accelerated
conditioning protocols that can be used to evaluate the susceptibility of a CFRP system to
moisture.
The recommended accelerated conditioning is divided into two primary environments: Wet
and Air. Wet environment calls for test specimens to be submerged in 60 deg. C water for
60 days. Air environment calls for exposure to 100% relative humidity at 60 deg. C for 60
days. (Hamilton et al. 2011)
Along with extensive accelerated conditioning testing, real time outdoor testing was also
conducted. This paper details the results of the real time exposure and compares the results
to those of the accelerated conditioning. Two real time exposure conditions were selected
to compare to the accelerated conditioning. One was in a tidal zone in brackish water in
Florida. The other was on the roof of a building in Wyoming. The Florida exposure
involved wetting and drying of the specimens with little solar exposure while the Wyoming
specimens were exposed to freeze-thaw and solar. Specimens constructed and tested at the
Univ. of Wyoming are denoted as Group 1 and those constructed and tested at the Univ. of
Florida are denoted as Group 2.
3. SPECIMEN CONSTRUCTION
The test specimen is a 100 mm x 100 mm x 356 mm (4 in. x 4 in. x 14 in.) beam, tested in
three-point bending (flexure) on a 305-mm (12-in.) span length (Fig. 1). The specimen is
prepared by providing a full-width half-depth saw cut approximately 2.5 mm (0.1 in.) wide
at midspan. Further details of the development of this specimen configuration can be found
in Gartner et al. (2011). Application of a CFRP system and surface preparation is in
accordance with NCHRP Report 514 (Mirmiran et al. 2004).
To determine if the proposed test procedures could differentiate resistance to the
accelerated aging, two - unidirectional wet-layup CFRP systems (denoted as A and B) and
one unidirectional carbon laminate system (denoted as C) were selected to reinforce the
specimens. Systems A, B, and C were commercially available CFRP composites in which
the fiber and epoxy were supplied jointly. All of these systems consisted of bisphenol A
epoxies with diamine hardeners.
Composite A was a unidirectional carbon fiber fabric and a two-component epoxy resin
with a mixture ratio 2.9:1.0 by mass. The system also included a sealant to provide a
protective coating. To assess the effectiveness of the coating, Group 1 specimens did not
apply the coating and Group 2 specimens used the coating.
102

Composite B consisted of unidirectional carbon fiber fabric, epoxy primer, epoxy putty,
epoxy saturant, and protective top coat.
Composite C consisted of a pre-cured unidirectional carbon fiber polymer laminate and
epoxy putty. Group 1 used a 25 mm (1-in.) wide strip and Group 2 used a 19 mm (-in.)
wide strip to assess the effects on concrete flexure-shear failure of the concrete specimen.

(a)
(b)
Fig. 1. Small beam bond test (a) test setup and (b) externally bonded composite
reinforcement
4. TIDAL CONDITIONING
4.1 Exposure Conditions
With the cooperation of the Florida Department of Transportation (FDOT), 29 specimens
were hung from the northeast fender of the SR 206 Bridge in Crescent Beach, Florida (Fig.
2). The bridge crosses the Matanzas River, which is a brackish water river on the Atlantic
Intracoastal Waterway.

SR206 Bridge

Orlando

Miami

Fig. 2. SR206 bridge in Crescent Beach, FL


Six concrete beams were hung from each fender beam using stainless steel cable and
clamps (Fig. 3). As indicated by the staining and barnacle growth on the fender system, the
beams were held at such elevation as to result in regular wetting and drying during natural

103

tidal fluctuations. Although the FRP composite reinforcement was oriented upward, the
specimens were protected from direct sunlight for most of the daylight hours by the fender
system. This prevented direct UV exposure and large temperature swings from solar
heating. Beam tests were conducted after 12 and 18 months continuous exposure.

Fig. 3. Typical real time beam installation


4.2 Results and Discussion
Twelve months after real time tidal exposure was initiated, 9 beam specimens (3 of
composite A, B, and C) were selected at random to be tested in flexure. All specimens
were retrieved by boat and upon arrival at fender system it was noted that the overall
location and orientation of each beam specimen remained relatively unchanged.
Additionally, it was observed that as a result of being exposed to tidal action all of the
beams had experienced significant barnacle growth (Fig. 4). The highest densities of
barnacle growth were seen on the bottom and side surfaces of each beam specimen. Despite
the large number of barnacles present along the perimeter of each CFRP strip, the strips
themselves experienced very little barnacle growth.

Fig. 4. Typical specimen layout after tidal exposure


After being removed from the bridge supports, the tidal specimens remained dry until
reaching the testing destination. At the testing location all remaining stainless steel support
strands were removed with bolt cutters and the specimens were submerged into a room
temperature, fresh water tank. After 24 hours of fresh water submersion, barnacles were
removed from the concrete substrate with a scraper blade. Care was taken to avoid
damaging the CFRP reinforcing strip. Immediately after removing the excess barnacles,
beam specimens were returned to the room temperature, fresh water tank. After an
104

additional 24 hours of fresh water submersion all nine specimens were tested in flexure to
failure.
After 18 months the remaining 20 beam specimens were removed from the tidal exposure
location to be tested in flexure. At this time there was no noticeable change in the density
of barnacles present on the surface of each concrete specimen compared to the 12 month
tidal exposure specimens. The same cleaning and testing procedures as for the 12-month
tests were used.
Test results after 12 and 18 months of tidal exposure are shown in Fig. 5. The plot presents
the average strength ratio for each parameter. The strength ratio is the ratio of the ultimate
strength of the conditioned specimens to that of the unconditioned specimens made with the
same materials and tested at the same age as the conditioned specimen. System A and B
both showed a decrease in strength after 12 months of 22% and 14%, respectively. System
C, however, actually showed a slight increase in strength after 12 months of exposure.
1.4

Strength Ratio

1.2
1.0
0.8
0.6
0.4

System A

0.2

System B
System C

0.0
0

100

200

300

400

500

600

Exposure Time (days)

Fig. 5. Strength ratios for tidal exposure


The initial reduction in strength of Systems A and B was followed by an increase in
strength over the following 6 months. In fact, System B increased to a strength ratio of
1.04 after 18 months of real-time exposure. After 18 months, the strength ratio of System
C increased further to 1.23. This increase might be due to the buildup of barnacles and
their associated marine adhesive, which protected specimens from deterioration and acted
as additional adhesive or sealant (El-Hawary et al. 2000). Furthermore, the relatively short
exposure time at these mild temperatures was insufficient to result in significant
degradation. The failure mode for all three systems was primarily adhesive with some
mixed-mode failure modes noted in System B.
Modulus of rupture beams were exposed with the FRP system beams for 18 months. The
strength ratio of these concrete beams without CFRP was an average of 1.01, indicating that
the concrete tension strength changed very little as a result of the tidal conditioning.
Consequently, the variations in strength ratios were all a result of changes in the CFRP
composites.

105

5. FREEZE-THAW AND UV CONDITIONING


5.1 Exposure Conditions
Sixty CFRP specimens (20 each of System A, B, and C) were placed on the roof of a six
story building in Laramie, Wyoming (Fig. 6). CFRP direct tension coupons were also
fabricated and exposed along with the beams. The local weather is harsh with high
ultraviolet radiation as well as cold and snow contributing to freeze-thaw and moisture
conditions in the winter. Generally, due to the high elevation, the winters are long with
relatively short spring, summer and fall seasons. No UV protection coating was used on
these specimens to evaluate the effect. Beam and tension tests were conducted after 6, 12,
and 18 months.

Fig. 6. Exposure in Wyoming


5.2 Results and Discussion
Freeze-thaw and UV conditioning was conducted in a real time environment that combined
UV, moisture variation and ambient freeze-thaw cycles present at an elevation of 7200 ft.
The flexural test results of CFRP Composite A, B and C Systems after 18 months are
shown in Fig. 7.
1.2

Strength Ratio

1
0.8
0.6
0.4

CFRP System A

0.2

CFRP System B
CFRP System C

0
0

100

200

300

400

500

Exposure Time (days)

Fig. 7. Strength ratios for solar exposure

106

600

The test results from CFRP System A, B and C solar exposure indicated the following:
System A flexural strength decreases sharply in the first 200 days. The specimens
failed by a mixed mode failure after 6 months and adhesive failure after 12 months.
System B flexural strength decreased sharply in the 200 days. The flexural bond
strength degraded up to 28% and specimens failed by a mixed mode failure after 6
months. At 12 months and beyond, the flexural strength tended to stabilize.
System C flexural test results showed linear strength reduction with time. The
specimens failed by mixed mode failure after 6 months. Adhesive failure mode
occurred after 12 months, and CFRP laminate slip was observed after 18 months.
6. COMPARISON WITH ACCELERATED CONDITIONING
Fig. 8 shows a comparison of the real time exposure results (Roof and Bridge) with results
from accelerated conditioning. The plot gives the average strength ratios for the three
tested CFRP systems. The accelerated conditioning consisted of submerging the small
beam specimens in 60 deg C heated water (60C-S) for the time indicated in the plot
followed by flexural testing to failure. In addition, specimens were subjected to the same
temperature at 100% relative humidity for 60 days (60C-RH). Group 1 indicates specimens
constructed at Univ. of Wyoming and Group 2 indicates specimens constructed at Univ. of
Florida. Details of the accelerated conditioning and testing can be found in Hamilton et al.
2011.
CFRP System

1.4

C
60 days
6 months
12 months
18 months

Strength Ratio

1.2
1
0.8
0.6
0.4
0.2

G2|Bridge

G1|Roof

G1|60C-RH

G2|60C-S

G1|60C-S

G2|Bridge

G1|Roof

G1|60C-RH

G2|60C-S

G1|60C-S

G2|Bridge

G1|Roof

G1|60C-RH

G2|60C-S

G1|60C-S

Conditioning

Fig. 8. Summary of average strength ratios. (G1Group 1, G2Group 2, 60C-S


submerged, 60C-RH100% relative humidity)
The drastic difference in behavior of System C between submerged accelerated
conditioning and wetting and drying is puzzling. The specimens degraded severely after
immersion in water at elevated temperatures. Further, the failure mode in the accelerated
conditions was that of delamination of the composite itself. Yet, relatively cool brackish
water (less than room temperature) did not cause degradation of System C. In fact, the
107

strength of the system increased over the real time exposure period. This appears to
indicate a significant influence of temperature on the precured laminate that was not seen in
the other wet layup systems.
Overall, the freeze-thaw/UV exposure results have consistently lower strength ratios than
those of the tidal exposure specimens. This may be an indication that the combination of
UV degradation along with the cyclic thermal loading due to differences in the coefficient
of thermal expansion (CTE) of the laminate, adhesive, and concrete may be a more
significant mode of deterioration than wetting and drying in brackish water under relatively
mild ambient temperatures. The diurnal temperature changes and the CTE differences
result in a residual stress between CFRP and the concrete that may be sufficient to affect the
bond strength.
The final recommendations of the research project were to submerge the specimens in 60
deg. C water for 60 days to evaluate externally bonded CFRP systems for durability. While
this approach does not necessarily cover all possible environments and conditions,
comparison of the results of the accelerated conditioning with that of the real time exposure
indicates that the recommended protocol will provide conservative results. Longer periods
of real-time exposure, however, are needed to verify that the accelerated conditioning
captures the long-term moisture degradation at ambient temperatures.
7. ACKNOWLEDGEMENTS
The research reported in this paper was sponsored by the American Association of State
Highway and Transportation Officials in cooperation with the Federal Highway
Administration and was conducted by the National Cooperative Highway Research
Program, which is administered by the Transportation Research Board of the National
Academies.
8. REFERENCES
1.
2.

3.

4.

Gartner, A., Douglas, E.P., Dolan, C.W., Hamilton, H.R. 2011. Small Beam Bond Test
Method for CFRP Composites Applied to Concrete. Journal of Composites for
Construction, ASCE, 15(1): 52-61.
Hamilton, H.R., Dolan, C.W., Tanner, J.E., and Douglas, E.P. 2011. Testing Protocol
for Bonded CFRP Durability. Proceedings 10th International Symposium on FiberReinforced Polymer Reinforcement for Concrete Structures, FRPRCS-10, eds: Sen, R.,
Seracino, R., Shield, C. & Gold, W., 2-4 April, Tampa, FL, 20 p.
Mirmiran, A., Shahawy, M., Nanni, A., and Karbhari, V. 2004. Bonded Repair and
Retrofit of Concrete Structures Using CFRP Composites, Recommended Construction
Specifications and Process Control Manual. National Cooperative Highway Research
Program Report 514, Transportation Research Board, Washington, DC.
El-Hawary, M., Al-Khaiat, H., and Fereig, S. 2000. Performance of epoxy-repaired
concrete in a marine environment. Cement and Concrete Research, edited by K.
Scrivener, Elsevier Science, 30: 259-266.

108

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

TENSILE CAPACITY OF STRESSED CFRP STRAND EXPOSED TO


EXTREME AGGRESSIVE GROUNDWATER ENVIRONMENTS
M. Sentry1, R. Al-Mahaidi2, A. Bouazza3and L. Carrigan1
1

Geotech Pty Ltd, Port Melbourne, Australia


Professor, Faculty of Eng. & Industrial Sciences, Swinburne University of Technology, Australia
3
Professor, Dept. of Civil Eng, Monash University, Australia
2

ABSTRACT
Current permanent ground anchor standards require the use of double protection systems
encapsulating the steel strands, to ensure a serviceable design life [1]. FRP reinforcement
has been successfully researched and used as a durable construction alternative to steel [2].
Minimal research has been devoted to developing FRP strands for use in high capacity
ground anchors under aggressive environments.
Finding from experiments investigating the effects of aggressive groundwater solutions on
various CFRP strands under stressed state were carried out on two different CFRP materials
under two varying ground conditions at elevated temperature of 60C under saturated
conditions. Tensile strength performance from this investigation indicated that under
extreme conditions certain CFRP strands are more reliable than others. This paper
summarises findings obtained from an extensive PhD research project.
1.

INTRODUCTION

To ensure adequate design life, conventional ground anchors use HDPE sheaths protecting
the steel strand to overcome environmental degradation [3]. Ground anchors in Australia
are exposed to acidic and alkaline ground conditions depending on grounds mineral
content, soil/rock type and age and groundwater flow path [4].
Minimal research has been conducted into the durability performance of CFPR materials
under conditions common to permanent ground anchor applications. Limited knowledge
on strand protection requirements for CFRP ground anchors is known. Greater
understanding into CFRP limitations when exposed to the aggressive ground environments
(groundwater and mineralogy) will provide sound insight into the protection levels required
when using CFRP material in permanent ground anchors.

109

Works presented investigated the tensile performance of two CFRP stands when exposed in
different aggressive environments under a stressed condition. Two different methods of
analysis were used to assess the material performance under extreme aggressive conditions
including tensile strength and scanning electron microscopy (SEM).
2.

PREVIOUS WORKS

Prior to investigating performance of CFRP strand under stressed conditioned when


exposed to aggressive environments, the authors previously presented findings from
research conducted investigating the effects of CFRP exposure to aggressive ground
environments under unstressed condition [4, 5]. Experiments were carried out under same
exposure conditions as works presented within this paper. Results from the unstressed
exposure samples showed minimal physical and tensile changes when compared to
controlled specimens.
3.

EXPERIMENT

3.1. Aggressive Ground Environment


Aggressive ground conditions generally relate to the aggressive nature of either the ground
(soil/rock/contaminated or hazardous waste fill) or the water within the ground
(groundwater) by means of increased levels of minerals, ions or contaminants.
Groundwater can be defined as the subsurface water located in soil pore spaces and in the
fractures of geological formations [6]. Three simulated ground environments were used to
asses CFRP performance (Table 1).
Table 1. Experimental simulated ground environments
Ground Environment Chemical Composition
pH Level
*
Acidic
NaHSO4 + MgCl2 + NaOH + NaCl
9.5 0.5
Alkaline
NaHCO3 + MgCl2 + NaOH + NaCl* 1.0 0.5
Neutral
H2 O
7.0 0.5
Note: * = added to increase salinity of solution to further accelerate aging of specimen.
3.2. Material Selection
Two separate materials from two different manufactures were used. A 15.2 mm diameter, 7
wire helically wound CFRP (CFRP-01) (this strand replicated the dimensions of a 15.2mm
diameter 7 wire low relaxation steel strand) and a 25 mm x 2 mm CFRP flat strand (CFRP02). Table 2 provides details of each material.

110

Table 2. Technical properties of CFRP test specimens


CFRP-01
CFRP-02
Property
Dimensions
Width (mm)
25
Diameter (mm)
15.2
Nominal Thickness (mm)
2
Effective Cross Sectional Area (mm2)
113.6
32
Linear Weight (g/m)
226
97
Carbon Fibre
Minimum fibre volume ratio
0.62
0.64
Density (g/cm3)
1.5
1.8
Tensile strength (MPa)
4200
4900
Elastic modulus (GPa)
240
230
Resin
Modified epoxy Vinylester
Type
Density (g/cm3)
1.6 - 2.0
1.15
Tensile strength (MPa)
80
>55
Product
Tensile strength (MPa)
2200
2500
Elastic modulus (GPa)
141
140
3.3.

Assessment Methodology

3.3.1. Tensile Assessment


Each sample was prestressed prior to curing in a unique stressing frame (Fig. 1Fig. 1a) to
50%Tult (Tult = ultimate breaking load). Once stressed, each sample was individually
exposed to the respective aggressive solutions. Samples were suspended within curing tank
for the required periods of time. Real time monitoring systems were applied to each
specimen to ensure applied stresses were maintained during curing.
Ultimate tensile capacity assessments were conducted in accordance with ASTM D3039M,
ASTM D3916 and ACI440.3R-04 [7-9]. Method of determining elastic modulus of the
cured specimens was in accordance with ACI440.3R-04. A range between 1000 and 3000
micro strain was used to establish the elastic modulus of each sample post curing as
recommended [9].
The results obtained for sustained loaded samples were compared against unstressed results
[4] to assess and compare tendon performance. Ultimate tensile testing was conducted
using a universal testing machine, calibrated and serviced by Monash University.
3.3.2. Scanning Electron Microscope (SEM) Analysis
SEM analyses micro structural changes to the CFRP tendon post specimen curing.
Assessment on the degeneration of the tendon can assist in understanding what effects this
111

may have on the specimens overall tensile performance. Once cured samples were
mounted in a neutral epoxy resin (Fig. 1Fig. 1b), examined surfaces were carbon coated
and analysed at Monash Universitys Centre for Electron Microscopy (MCEM) using the
JEOL JSM-840A scanning electron microscope. Electron back scatter diffraction was used
to visually assess physical alterations to the CFRP strands. Quantitative analytical
spectroscopy was implemented to assess the depth of penetration of chemicals present in
the aggressive ground solutions each specimen was exposed to during curing.
3.4. Test Program
Specimens were cured at temperatures of 60C to accelerate the curing process and attempt
to induce material deterioration. Temperature remained constant throughout each
experiment. Samples were cured for a period of one, three and six months prior to testing.
Details of the testing regime are summarised in Table 3.

Aggressive

Mounted

solution

Stressing

chamber

specimen

support

(b)

(a)

Fig. 1. (a) CFRP tendon stressing frame; (b) Mounted CFRP sample for SEM analysis
Table 3. Experimental test program
Number of Samples Tested
Exposed Curing
Temp
CFRP-01
CFRP-02
Total
Test
Time
Length
(C)
(mths) Alk Aci Neu Alk Aci Neu
(mm)
1
3
3
3
9
3
3
3
3
Tensile
9
300
60
6
3
3
3
3
3
3
18
Capacity
Total
9
9
9
3
3
3
36
6
1
1
1
1
1
1
6
60
300
SEM
Total
1
1
1
1
1
1
6
4.

RESULTS

4.1. SEM Analysis


Unfortunately due to the curing process, CFRP-02 samples taken for SEM analysis had
been damaged beyond use as a result of gripping issues. This indicates that curing the
CFRP-02 specimens in any solution type at elevated temperatures affects the bonding
112

between the fibres and the adjacent bond interface. Additional assessments needs to be
carried out on CFRP-02 specimens once sustained gripping issues are resolved.
4.1.1. Neutral Solution
CFRP-01 samples showed no signs of fibre fatigue, damage, fibre splitting or de-bonding
with the surrounding matrix after sample exposure to the neutral solution under stressed
conditions (Fig. 2a).
4.1.2. Acidic Groundwater
CFRP-01 cured samples showed no change in state from the controlled results (Fig. 2b).
The inner king wire was fully intact with no signs of deposits, micro-cracking or voids as
a result of the stress-cured curing process. Spectroscopy analysis conducted on the
localised outer fibre boundary layer showed high concentration levels of sodium and
chloride (Fig. 3). No signs of deposit penetration were observed from the visual SEM
analysis.

(a)
(b)
Fig. 2. Stressed CFRP-01 tendon cured for 6 months in (a) neutral solution (b) acidic
aggressive groundwater.

Fig. 3. SEM image and spectrum analysis for acidic groundwater cured CFRP-01 samples
after 6 months.

113

4.1.3. Alkaline Groundwater


CFRP-01 samples showed localised micro-cracks of the outer fibre layer had enabled the
penetration of deposits from the aggressive groundwater solution along a micro-crack
within close proximity to the outer boundary layer (Fig. 4a). Close inspection of the deposit
exposed some minor deterioration of the bond between the epoxy and fibres (Fig. 4b) and
concentrations of Sodium-Chloride crystals (Fig. 4c).
Spectroscopy analysis conducted on the localised outer fibre layer micro-cracks (Fig. 5)
showed high concentration levels of sodium and chloride and minor concentrations of
calcium. Although penetration of the aggressive alkaline solution occurred in the CFRP-01
sample, no sign of fibre deterioration was evident.

(a)

(b)

(c)

Fig. 4. Stressed CFRP-01 samples cured at 60C for 6 months (a) outer fibre layer
penetration; (b) deterioration of inner fibre region; (c) deposit within inner fibre region

Fig. 5. SEM image and Spectrum analysis of CFRP-01 sample after 6 months.
4.2. Tensile Capacity
Due to the flat profile of the CFRP-02 specimen, consistency issues with gripping during
testing resulted in inconsistent findings. Method of specimen anchorage during elevated
temperature curing was established as a reliable method of sustaining load. Further

114

investigations are being carried out to determine an effective gripping system to repeat
these tests.
Visual inspections of CFRP-01 tendon failure locations identified no consistent failure
pattern for all specimens tested over the six months in all curing environments. Failure was
by rupture, located within the gauge length. Failure strains for CFRP-01 specimens cured
in acidic, alkaline groundwater and neutral solutions indicated minimal variation compared
to control results (Fig. 6).
4.2.1. Neutral Solution
Results indicted a decrease in elasticity of 8.8% curing for six months (Fig. 6). Ultimate
strength increases of 14.5%, 15.5% and 16.3% were recorded for neutral stress-cured
specimens. Strength results observed confirms findings from SEM analysis.
4.2.2. Acidic Solution
Elasticity results were consistent, indicating that there is less impact on ductility variations.
Increases in strengths of 10.6%, 17.0% and 14.2% were recorded (Fig. 6). Compared to
unstressed results [4, 5], strength performance of CFRP-01 samples stress-cured out
performed unstressed-cured specimens. Recorded tensile capacity verifies findings from
SEM analysis. Minimal change in physical and engineering properties result from curing
of CFRP-01 samples under stressed conditions.
4.2.3. Alkaline Solution
Elasticity variations of 7.0%, 13.5% and 1.1% were recorded after one, three and six
months respectfully. Strength increases of 10.2%, 14.7% and 13.0% were recorded over
the six month assessment compared to controlled results (Fig. 6). These findings indicate
the stress-cured environment does alter the elasticity properties of the composite structure
in an alkaline solution. Although penetration of the aggressive solution occurred after six
months curing, this did not affect the tensile performance of the CFRP-01 sample.
5.

CONCLUSIONS

Tensile capacity and SEM assessment was carried out under stressed conditions at elevated
temperatures over a six month period on CFRP tendons exposed to aggressive acidic,
alkaline ground environments and neutral environments which can exhibit in ground
environments permanent ground anchors are installed in. CFPR-01 samples did not show
any sign of tensile strength decrease as result of the accelerated curing process even though
SEM analysis indicated penetration of aggressive solution had occurred after six months
curing. Due to gripping issues, no reliable results were obtained for the CFRP-02 samples.
Further research into a reliable gripping system for the flat tendon is required prior to reassessing the performance of this product under accelerated aggressive stressed conditions.

115

2500

2250

2000

1750

Stress(MPa)

1500

1250

1000

750

500

CFRP01 CONTROL

CFRP01 ACI 6mths

CFRP01 NEU 6mths

CFRP01 ALK 6mths

CFRP01 ACI 3mths

CFRP01 NEU 3mths

CFRP01 ALK 3mths

CFRP01 ACI 1mth

CFRP01 NEU 1mth

CFRP01 ALK 1mth

250

0
0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.1

1.2

1.3

1.4

1.5

Strain(%)

Fig. 6. Tensile test results of stressed CFRP-01 samples.


6.

ACKNOWLEDGEMENTS

The authors would like to express their gratitude to Monash University and the civil
engineering laboratory and research sponsor Geotechnical Engineering.
7.

REFERENCES

1.

Littlejohn, G.S. 1987. Ground Anchorages: corrosion performance. Proceedings Institute of Civil Engineers, Part 1 (82): 645-662.
Benmokrane, B., and Chennouf, A. 1997. Tensile and bond properties of AFRP and
CFRP tendons for ground anchor applications. Proceedings of the 3rd Symposium on
non-metallic reinforcements for concrete structures, FRPRCS-3, Sapporo, Japan, 14-16
October, 2: 727-734.
Sentry, M., Bouazza, A., Al-Mahaidi, R., Loidl, D., Bluff, C., and Carrigan, L. 2009.
Carbon Fibre Reinforced Polymer (CFRP): An Alternative Material in Permanent
Ground Anchors. Journal of Australian Geomechanics, 44(3): 47-56.
Sentry, M. 2010. An investigation into the durability and performance effects of CFRP
strand in permanent ground anchors, in Department of Civil Engineering, Monash
University, Clayton. 398 p.
Sentry, M., Bouazza, M., Al-Mahaidi, R., Collins, F., Carrigan, L., Loidl, D., Bluff, C.
2008. Durability of Carbon Fibre Reinforced Polymer (CFRP) strand for use in high
capacity permanent ground anchors. Proceedings of the Australian Structural
Engineering Conference (ASEC), Melbourne, 26-27 June, 8 p.
Sentry, M., Bouazza, M., Al-Mahaidi, R., Collins, F., Carrigan, L., Loidl, D., Bluff, C.
2007. Durability of Carbon Fibre Reinforced Polymer (CFRP) strand for use in high

2.

3.
4.
5.

6.

116

7.
8.
9.

capacity permanent ground anchors. Proceedings of the 14th international conference


on composite structures (ICCS14), Melbourne, Australia, 9 p.
ASTM. 2000. Standard test method for tensile properties of polymer matrix Composite
Materials. (D3039-00), West Conshohocken, Pa.
ASTM. 2002. Standard test method for tensile properties of pultruded glass-fibrereinforced plastic rod. (D3916-02), West Conshohocken, Pa.
ACI Committee 440. 2004. Guide Test Methods for Reinforcing or Strengthening
Concrete Structures. ACI 440.3R-04, American Concrete Institute, Farmington Hills,
MI, 40 p.

117

118

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

THE EFFECT OF LONG-TERM LOADING ON THE CFRPCONCRETE BONDING STRENGTH


C. Mazzotti1 and M. Savoia2
1
2

Associate Professor DICAM - Structural Engineering, University of Bologna, Italy


Professor DICAM - Structural Engineering, University of Bologna, Italy

ABSTRACT:
Results of an experimental campaign concerning the long-term behavior of FRP plates
bonded to concrete are presented. Four CFRP plates with three different bonded lengths
(from 100 400 mm) have been subject to long-term constant traction force. The tests have
been carried out in a climatic room with a constant temperature T = 20 C and humidity RH
= 60%, for about five years. The strain profile evolution with time along the bonded length
has been recorded by using a number of strain gauges along the plates. An appreciable
stress redistribution with time along the plate has been observed. At the end of the longterm test, all specimens have been subjected to failure load in order to verify if the interface
behaviour was affected by the long-term loading application. No remarkable degradation
has been found.
1. INTRODUCTION
Most of the studies carried out, in recent years, regarding the bond between the Fiber
Reinforced Polymer (FRP) and concrete concern the characterization of its performances
under short-term loadings. Nevertheless, it is very important to verify the durability with
time of the retrofit intervention. The effectiveness of the strengthening can in fact be in
principle reduced due to the rheological behaviour of the materials (concrete, adhesive,
FRP) in the case of long-term loadings.
As far as the behaviour of FRP-strengthened elements under sustained load is concerned,
only in the last few years the scientific contributions are closing the gap of knowledge,
starting from the early work of Plevris and Triantafillou [1]. Some contributions can be
found, i.e., experimental investigations [2], [3], [4] and numerical models [5], [6], [7]. A
common result of all research studies is that, under long-term loading, a shear stress
redistribution occurs, during time, along the FRP element due to the viscosity of both
concrete and adhesive. It has not been investigated up to now if the long-term loading could
negatively affect the bond strength of the strengthening system, eventually modifying the
interface law and the transmissible force.
119

In the present paper, the results of an experimental campaign concerning the long-term
behaviour of the bond between FRP plates and concrete prisms are presented. Three
different bonded lengths (100 - 400 mm) have been tested. The time duration of the tests
has been about 2000 days (more than five years of continuous loading). The strain
distribution along the bonded part has been monitored with time.
After the unloading, the specimens have been subjected to instantaneous bond failure test in
order to verify if the previous loading history affected the bond strength. Strains along the
reinforcement during the tests have been recorded and the corresponding shear stress
distributions evaluated, thus allowing for the calibration of shear stress-slip interface laws.
All the data have been compared with analogous information obtained, five years ago, by
tests on reference specimens made of the same constituents. The comparison shows that no
particular damage occurred to the interface behaviour due to the long-term loading.
2. LONG-TERM BOND TESTS
2.1 Specimen Preparation
The long-term behaviour of CFRP plates bonded to concrete has been investigated by
testing three different bonded lengths (B.L. = 100, 200 and 400 mm) and two concrete
specimens (two CFRP plates for each concrete specimen).
The concrete specimen dimensions were 150200600 mm (Fig. 1). Specimens were
fabricated using normal strength concrete (at the age of the long-term loading application,
compression strength was fcm = 52.6 MPa and tensile strength, fctm = 3.81 MPa, from
splitting tests). Further details can be found in [4].

200

CFRP plate
801 2 mm

Adhesive
801 5 mm

150
7 to 11 strain gauges

L9

L2

100 mm
L1

(b)
CONSTANT
LOADS

B.L
.
100
200
400

7 to 11 strain gauges

L1 L2 L3 L4 L5 L6 L7 L8 L9

L1
L11
0

10 10 10 15 15 15 15
10 10 10 20 20 20 30 30 40
10 10 20 20 40 40 50 50 50 50 50

(a)
(c)
Fig. 1: (a) Geometry of specimens and (b) experimental setup for the long term tests,
(c) spacing (mm) between strain gauges along the FRP plates with different B.L.
120

(b)

CFRP plates, 80 mm wide and 1.2 mm thick, have been used; according to technical data
provided by the producer, the mean elastic modulus is Ep = 165000 MPa.
Two opposite surfaces of each concrete block have been grinded with a stone wheel to
remove the top layer of mortar, until the aggregates were visible (approximately 1 mm).
The plates have been bonded to the concrete surfaces by using a 1.5 mm thick layer of a
two components epoxy adhesive, having mean tensile strength of 30.2 MPa and mean
elastic modulus Ea = 12840 MPa. No primer before bonding has been used. The plate
bonded length starts 100 mm far from the front face of the specimens (Fig. 1a). Two CFRP
plates have been bonded to the opposite faces of each concrete specimen, with different
bonded lengths (concrete Block P1: B.L. = 100, 200 mm; Block P2: B.L. = 200, 400 mm),
see the layout reported in Fig. 1a. The curing periods of the anchorages of blocks P1 and
P2 were 15 and 20 days prior to testing, respectively.
2.2 Experimental Setup and Instrumentation
The experimental setup is depicted in Fig. 1a. The concrete block was positioned on a rigid
frame with a steel reaction element at the front side (60 mm height), in order to prevent
longitudinal rigid displacements. The traction force was applied to each couple of CFRP
plates by using a mechanical frame allowing for the magnification of the force applied at
the opposite side of the horizontal arm (leverage system). The force was given as the
weight of a number of steel plates (Figure Error! Reference source not found.Error!
Reference source not found.1b); details on the experimental set-up can be found in [4].
The long-term tests were performed by prescribing a constant force during time.
Along the CFRP plates bonded to concrete, series of 7 11 strain gauges (depending on the
plate length) were placed, along the centerline. For each bonded length, the spacing
between the strain gauges is reported in Fig. 1c.
2.3 Loading Program
Two different loading histories have been prescribed. The FRP plates of specimen P1 have
been subject to a traction force of 12.30 kN (about 50 percent of the maximum
transmissible load, according to debonding tests on analogous specimens, see [8]), constant
in time for 1981 days. The plates bonded to specimen P2 have been subject to the same
sustained load (12.30 kN) for 510 days; after that, a second load step of 3.80 kN has been
applied and maintained for further 1454 days. The longitudinal strains along the plate at
different time intervals have been recorded by an automatic computer system. Further
details can be found in [4].
2.4 Long-term Results
Figures 2a-d show the time evolution of the longitudinal strains along the plates for 100
200 mm (specimen P1) and 200 400 mm (P2) bonded lengths, respectively. For each
strain gauge, the distance from the initial section of the anchorage is reported in
parenthesis. It can be observed that, at medium-low stress levels, plates with different
121

bonded length exhibit a similar behaviour: the rate of strain starts decreasing immediately
after the load application (as in classical creep tests on concrete specimens). The continuous
and uniform reduction of the strain rate for more than five years, as already shown in [4]
using the log-scale of time for a shorter time interval, suggest that after few months of
loading a steady-state creep condition was attained.

Longitudinal strain (

700

E1 (10)

E2 (20)

E3 (30)

E5 (60)

E6 (75)

E7 (90)

800

E4 (45)

700

(10)

Longitudinal strain (

800

600
(20)

500
400

(30)

300
(45)

200
(75)

100

(60)

0
730

1095

1460

E3 (30)

E4 (50)

E7 (120)

E8 (150)

E9 (190)

1825

(20)

400

(30)

300

(50)
(70)

200

(90)

2190

365

730

(a)

2190

(10)

E1 (10)

E2 (20)

E3 (30)

E4 (50)

E5 (70)

E6 (90)

E7 (120)

E8 (150)

E9 (190)

700

Axial strain ()

Longitudinal strain ()

1825

800

600

(20)

500

(30)

400
300

(50)

200

(120)

100

730

1095

1460

nd

Load step
(20)

600

E1 (10)
E5 (100)
E9 (290)

500
400

E2 (20)
E6 (140)
E10 (340)

E3 (40)
E7 (190)
E11 (390)

200

(150)

100

(100)

0
1825

2190

Time (days)

E4 (60)
E8 (240)

(40)

300

(90)

(190)

(60)

(70)

365

1460

900

Load step

700

1095

(b)

(10)
nd

(150)

(190)

(120)

Time (days)

900
2

(10)

500

Time (days)

800

E5 (70)

600

0
365

E2 (20)

E6 (90)

100

(90)

E1 (10)

365

730

1095

Time (days)

1460

1825

(140)
(190)
(240)
(290)
(340)
(390)

2190

(c)
(d)
Fig. 2: Time evolution of strains in FRP plates at different positions along the anchorage:
specimen P1 (a) B.L.=100 mm, (b) B.L.=200 mm; specimen P2 (c) B.L.=200 mm, (d)
B.L.=400 mm. The number in parenthesis indicates the distance (mm) from the initial
section of the anchorage.
As a confirmation, Fig. 4a shows the strain evolution with time of the strain gauges
belonging to the two plates with B.L.=200 mm. Thick lines refer to the specimen P2
subject to the second loading step p2 510 days after the first loading p1 (p2=0.35p1),
whereas thin lines refer to the specimen P1 subject to the initial loading step only (p1).
The evolution with time of the strain curves obtained from specimen P2, after the second
loading step, rapidly attains the same slope of the corresponding curves from specimen P1.
Finally, Figs. 3a-d show the strain profile along the plates at different times, for all the
specimens. The dashed lines join the first experimental strain value recorded at x=10 mm
with the theoretical value of the FRP strain at x=0, evaluated as 0(t)=F/ Ep(t) Ap, with t
being the time after the loading application and Ep(t) the CFRP elastic modulus, taking
into account the material viscosity (very small). As expected, at the beginning of the
bonded length the plates show increasing strain values with time, approaching the limit
value of 0 (no redistribution can occur in the plate free part). The strain profiles are
qualitatively similar, suggesting the increase of the delayed strain with time reduces
122

moving along the bonded length; on the contrary, at a distance greater than 150 180 mm
from the loaded end (a realistic value of transfer length), the delayed strain increase with
time is almost constant along the bonded length, so confirming the experimental results by
Wu and Diab [2]. After the second load increment (Figures 3c,d), the long-term strain
variation is less pronounced, due to the concrete and the interface ageing, and the strain
curves shifting along the bonded length with time is progressively reducing.
The creep strains measured on the CFRP plates suggest that the adhesive and the concrete
cover, constituting the interface where the load is transferred, are subject to creep strains,
thus increasing the compliance of the interface with time. A shear stress redistribution
along the interface then occurs: high shear stresses close to the loaded end decrease and a
longer portion of the bonded length is subject to the shear stress transfer, i.e. the transfer
length increases.
Due to the very high shear stresses at the adhesive and concrete cover level, close to the
loaded end the creep phenomenon may be highly non linear. Hence, the compliance
increase at the beginning of the anchorage can be much higher than far from it. The
evolution with time of the creep deformation at the interface level is then variable along the
anchorage, and may explain the shear stress redistribution along the interface. As a
consequence, the normal stresses in the plate increase far from the loaded section of the
plate, so explaining the large increase of the axial strains with respect to the values recorded
immediately after the load application.
800

800

End of Loading

End of loading
700

30 days

Longitudinal strain ()

Longitudinal strain ()

700

192 days

600

932 days
500

1981 days

400
300
200

30 days
192 days

600

932 days
500

1981 days

400
300
200
100

100

0
0

10

20

30

40

50

60

Distance (mm)

70

80

90

100

20

40

60

(a)

100

120

140

160

180

200

(b)
1000

1000

End of Loading

900

510 days

600

2nd Load step

Axial strain ()

175 days

700

676 days

500

930 days

400

End of Loading

900

30 days

800

Axial strain ()

80

Distance (mm)

1964 days

300

30 days

800

175 days

700

510 days
2nd Load step

600

676 days

500

930 days

400

1964 days

300

200

200

100

100
0

0
0

20

40

60

80

100

120

140

160

180

200

Distance (mm)

40

80

120

160

200

240

Distance (mm)

280

320

360

400

(a)
(b)
Fig. 3: Strain distributions along the bonded length, at different times. Specimen P1: (a)
B.L.=100 mm, (b) B.L.=200 mm and specimen P2: (c) B.L.=200 mm, (d) B.L.=400 mm.
123

3. FAILURE TESTS
After the conclusion of the long-term tests, all the specimens have been subject to failure
tests in order to verify the effect of the long-term loading on the FRP-concrete bond
strength and the type of failure. For all the specimens, failure was caused by shearing of a
layer of concrete 1-2 mm thick. A detailed description of the experimental set-up can be
found in [8]. Similar bond tests were also carried out at the beginning of the long-term
experimental campaign on specimens made with the same materials; a comparison between
results before and after the long-term loading can then be given.
Figure 4b shows the comparison between the bond forces obtained, for the different bond
lengths, before and after the creep tests. The tests clearly show that the long-term loading
did not cause damage of the interface transferring the shear force. A systematic increase of
the bond force can be in fact observed for all the bond length but bl=400 mm; in this case,
during the preparation of the failure test a partial debonding of the front side of the CFRP
plate occurred, thus producing a bond degradation and a reduction of the corresponding
force of about 15%. The force increase can be only partially explained by the strength
increase with time of concrete substrate due to aging; in particular, the compressive
strength measured on cylinders obtained from the concrete specimens is fcm(6 years)=54.2
MPa while at the beginning of the long-term tests was fcm(9 months)= 52.6 MPa. According
to the expression reported in [9], this strength variation should produce a bond force
increase of about 1 kN, much smaller than the measured increase. The increase of bond
strength can be also due to the ageing of the epoxy, for which, nevertheless, there are no
available data.
45

P2-200_E1
P1-200_E1

1000
900

P2-200_E3
P1-200_E3

P2-200_E5
P1-200_E5

P2-200_E7
P1-200_E7

700
600
500
E3

400
300
200
100

Before creep test


After creep test
After creep test

35

E1

Bond force (kN)

Axial strain

800

40

30
25
20
15

E5

10

E7

365

730

1095

Time (days)

1460

1825

2190

100 mm

200 mm

400 mm

Bond length (mm)

Fig. 4: (a) Strain evolution with time recorded by strain gauges E1, E3, E5, E7 (specimens
P1 and P2, B.L.=200 mm) and (b) bond strength measured before and after long-term tests.
From the strain-gauges measures along the FRP bonding, the strain distribution along the
reinforcement for increasing values of the applied force during failure tests have been
obtained; these distributions are reported in Figures 5a-d; points A to C refers to subsequent
steps of the debonding process. The shapes of the curves, both at low level of applied force
and at the onset of debonding or during it, are qualitatively similar to those obtained in
standard conditions (without application of a preceding long-term loading). Only in the
case of bl=400 mm, the debonding occurred before the test produced a strain distribution in
124

the first 40 mm of the reinforcement (yellow area of Figure 5d) almost constant also at the
beginning of the test. This portion of the strain curve has not been then considered in the
subsequent analyses. In order to verify if the strain distribution was affected by the long
term loading, the strain curves along the reinforcement from specimen P2-200 (solid
markers), for increasing values of the applied force, have been compared in Figure 6a with
the analogous results obtained from the reference specimen with the same bond length
(hollow markers). A very good matching can be observed at both low level of applied force
(linear-elastic behaviour) and during debonding.
3000

3000

4 KN

4 KN

12 KN

2500

12 KN

2500

20 KN

20 KN

28 KN

28 KN

Axial strain ()

Axial strain ()

2000

2000

34 kN
35 kN

1500

36 kN
40 kN
point A

1500

point A
point B

point B
point C

1000

1000

point D

500

500

0
0

10

20

30

40

50

60

70

80

90

100

20

40

60

80

100

120

140

160

180

200

x (mm)

x (mm)

(a)

(b)

3000

3000
4 KN

4 KN

12 KN

2500

12 KN

2500

20 KN

20 KN

FRP Debonded part

28 KN

28 KN
2000

Axial strain ()

Axial strain ()

2000

36 kN
37.5 kN
point A

1500

point A
point B

1500

point B
point C

1000

30 KN

point C

1000

point D

500

500

0
0

20

40

60

80

100

120

140

160

180

200

50

100

150

200

x (mm)

x (mm)

(c)

(d)

250

300

350

400

Fig. 5: Strain distribution along the bonded part during the failure tests Specimen P1 (a)
B.L.=100 mm, (b) B.L.=200 mm, Specimen P2 (c) B.L.=200 mm, (d) B.L.=400 mm.
Following a standard procedure [8], the average shear stress between two subsequent strain
gauges and the corresponding CFRP-concrete slip have been evaluated as a function of the
measured strains, thus obtaining a set of couples of average shear stress and slip, used to
calibrate a local interface law [10]. Figure 6b show the interface law calibrated adopting the
data obtained from the reference specimens and from those subject to the creep tests,
together with all the shear stress slip data. The initial branch of the curves has not been
affected by the long-term loading (same linear elastic stiffness); the interface law of the
specimens subjected to the creep test exhibits a higher maximum shear stress and a slightly
steeper softening branch. The fracture energy in the second case is greater, due to the
greater value of the debonding force.

125

Reference test

Axial strain ()

2000

1500

16

4 KN
12 KN
20 KN
28 KN
point A

4 kN-Ref
12 kN-Ref
20 kN-Ref
28 kN-Ref
point A-Ref

point B
point D

point B-Ref
Point C-Ref

Referencebonddata
ReferenceCurve
Bonddataaftercreeptest
Curveaftercreeptest

14
12

Test after
creep load

1000

Shearstress(MPa)

2500

10
8
6
4

500
2
0

0
0

20

40

60

80

100

120

140

160

180

200

0.05

0.1

0.15

x (mm)

Slip(mm)

(a)

(b)

0.2

0.25

0.3

Fig. 6: Reference case (B.L.= 200 mm) and specimen P2 after the long-term test (a)
Strains along the bonded part and (b) shear stress-slip curve.
4. CONCLUSIONS
The results of an experimental campaign concerning the long-term behaviour (about five
years of loading) of bond between concrete prisms and FRP plates are presented.
Bonding with plate with different lengths exhibit a similar behaviour with time,
characterized by a steady-state delayed strains increase reached after about three months of
loading. After more than five years of testing, no remarkable change of quality of the curve
has been observed (steady interface viscosity).
The application of subsequent load steps at different times shows the long-term interface
behaviour is mainly governed by the aging of the concrete (enhancement of the mechanical
properties due to the evolution with time of the hydration level of the cement paste).
Moreover, the post-processing of the experimental results shows a significant redistribution
of the shear stresses along the anchorage due to the creep deformation at the interface level.
The failure tests carried out at the end of the long-term test show that no appreciable
damage occurred due to creep of the interface.
5.

ACKNOWLEDGEMENTS

The financial supports of the (Italian) Dept. of Civil Protection (RELUIS 2009 Grant
Task 3.1: New materials for the seismic retrofit) is gratefully acknowledged. The authors
would like to thank also the Sika Italia S.p.A.
6.

REFERENCE

1.

Plevris, N. and Triantafillou, T.C. 1994. Time-dependent behaviour of RC members


strengthened with FRP laminates. Journal of Structural Engineering, ASCE, 120(3):
1016-1042.
Wu, Z. and Diab, H. 2007. Constitutive model for time-dependent behaviour of FRPconcrete interface. Journal of Composites for Construction, ASCE, 11(5): 477-486.

2.

126

3.

Choi, K.-K., Meshgin, P. and Reda Taha, M.M. 2007. Shear creep of epoxy at the
concrete-FRP interfaces. Composites Part B: Engineering, 38, (5-6): 772-780.
4. Mazzotti C., Savoia M. 2009. Stress redistribution along the interface between concrete
and FRP subject to long-term loading. Advances in Structural Engineering, 12(5): 648658.
5. Savoia, M., Ferracuti B. and Mazzotti, C. 2005. Long-term creep deformation of FRPplated r/c tensile members. Journal of Composites for Construction, ASCE, 9(1): 6372.
6. Diab, H. and Wu, Z. 2007. Nonlinear constitutive model for time-dependent behaviour
of FRP-concrete interface. Composites Science and Technology, 67(11-12): 23232333.
7. Benyoucef, S., Tounsi, A., Adda Bedia, E.A. and Meftah, S.A. 2007. Creep and
shrinkage effect on adhesive stresses in RC beams strengthened with composite
laminates. Journal of Composites Science and Technology, 67(6): 933-942.
8. Mazzotti, C., Savoia, M. and Ferracuti, B. 2008. An experimental study on
delamination of FRP plates bonded to concrete. Construction and Building Materials,
22(7): 1409-1421.
9. Research National Council CNR Committee 2004. Guidelines for Design,
Construction of Externally Bonded FRP Systems for Strengthening Existing Structures.
CNR DT 200/2004, Rome, Italy, 144 p.
10. Ferracuti, B., Savoia, M. and Mazzotti, C. 2007. Interface law for FRPconcrete
delamination. Composite Structures, 80(4): 523-531.

127

128

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

DESIGN OPTIMIZATION OF COMPOSITE STRUCTURES


CONSIDERING BONDED JOINS DURABILITY
P. Hamelin1, A. Si Larbi2 and E. Ferrier3
1

Professor, University of Lyon, Laboratory LGCIE, Lyon, France


Associate Professor, University of Lyon, Laboratory LGCIE, Lyon, France
3
Professor, University of Lyon, Laboratory LGCIE, Lyon, France
2

ABSTRACT
First we introduce the notion of critical length for load transfer between substrates and
adhesive. Then we present viscoelastic measurements which permit to evaluate the
rheological behaviour of the epoxy and to identify its creeping function. For a final
application which concerns mixed steel-concrete structure connected by adhesive join we
are able to design an innovative technological process which consider discontinuous epoxy
blobs with a length defined from durability analysis.
1. INTRODUCTION
In recent years, structural bonding has been much used for the repair of damaged structures
in reinforced concrete [1], [2], considering reinforcement by the addition of composite
pultruded plates, or by the introduction of additional composite rebars (NSM process) [3] in
zones subjected to tension or shear.
The performance and durability of epoxy adhesives has been established and the principal
characteristics considered in the dimensioning rules are the properties in instantaneous and
delayed tensile and shear resistance. In the case of bonding on concrete, the modes of
rupture often depend on the surface treatment of the concrete and of the tensile and shear
resistance properties. We are particularly interested, in this research, in optimizing the
design of bonded joints to ensure load transfer by shear between the concrete and the steel
and also in optimizing the dimensioning of composite concrete-steel beams in order to
achieve a non-fragile ultimate behaviour controlled by the plastic deformation of the steel
and by the absence of any initiation of cracking in the bonded joints following the HSF
methods proposed by P. Hamelin, 2009 [4].

129

2. OPTIMIZATION OF MECHANICAL PERFORMANCE RELATIVE TO


SHEARING BONDED JOINTS: Notion of a critical length in function of glue
nature
2.1. The Theory: Volkersen Expressions
We first consider the standard Volkersen expressions [5]. The expression for the variation
of shear stress as a function of abscissa x of a point on the bonding surface is given by:

x moyen

Where
F = external force
l = length of overlap between the two substrates
h = width of overlap
G = shear modulus of the adhesive
E1 and E2 = longitudinal elasticity
(1)modulus of
substrates 1 and 2
s1 and s2 = thickness if substrates 1 and 2
e = thickness of the layer of adhesive
adh shear resistance of the adhesive

l chx E 1 s1 E 2 s 2 shx

.
l
2 l E1 s1 E 2 s 2
sh
ch

2
2

With: moyen

1/ 2
F
et G E1 s1 E 2 s 2
lh
E1 s1 E 2 s 2 e

(1)

(2)

Fig. 1. Description of the bonded joints [12]


Considering: E2

kE1

, s 2 ps1 ,

1 kp
1 kp

we simplify the expression to: x moy

and

1 kp
kp

(3)

l cosh x
sinh x
with
M.
l
l
2
sinh
2

cosh

GN
E1 s1e

(4)

Since stress is maximal at the extremities, shear failure of the bonded joint takes place for
/ 2 adh we can establish a relationship between adh and moy
Calculation of the force supported by the joint Fr moy .bl
Frd l

adh
A

1
with A
coth Al M tanh Al

GN
(Hamelin, 2009)
4 E1es1

(5)
(6)

2.2. Influence of the Nature of the Adhesive on the Mechanical Behaviour of a SteelConcrete Assembly
With substrate n1 being concrete and substrate n 2 being steel, following the data in
Table 1, we used the expressions above to study, for two types of adhesive (hard epoxy
130

[Sikadur 31] and flexible polyurethane [Sikaflex]), the variation of the load supported for
different lengths of bonded joint (Fig. 2).
Table 1. Data for two different adhesives with two substrates (concrete and steel)
E1 (MPa)

s1 (mm)

E2 (MPa)

s2 (mm)

e (mm)

b (mm)

30 000

50

210 000

5,7

7,000

0,1140

0,1123

2,2531

3,000

55

Load supported by the bonded join

Polyurthane

30,0000

Epoxy

out

25,0000

G (MPa)

29

adh (MPa)

9,5

A(polyurthane)

0,003392

kN )
Load ((N)

20,0000

15,0000

Epoxy

10,0000

Fcritical
5,0000

200

polyurethane

critical

critical

0,0000

400

600

800

1000

Lenght

1200

1400

1600

1800

G (MPa)

4 590

adh (MPa)

19,5

A(epoxy)

0,042678

2000

(mm)

Fig.2. Theoretical force supported by the bonded joint for two types of adhesive
There is a critical, or efficient, bond length above which any increase in length provides no
increase in the force supported by the join (Fig. 2). In fact, the maximum shear stress in the
adhesive is not reached because the failure appears in the concrete with an average
maximum shear stress equal to 2MPa. The critical length which depends on the mechanical
performance of the adhesive can be evaluated considering the same calculation with two
metal substrates (Table 2) which present higher resistance in comparison with concrete.
Table 2. Data for the calculation of the joints bonded load for two metal substrates
E1
(MPa)

s1
(mm)

E2
(MPa)

s2
(mm)

210000

3,25

210000

3,25

1,0000

1,0000

0,0000

2,0000

131

kp

G
(MPa)

e
(mm)

(MPa)

4590

19,5

b
(mm)

0,0739

0,041004

10

Epoxy

Fr (kN)
(N)

criticalsteel

l (mm)

Optimum shear length = 60mm

Fig. 3. Shear resistance of the bonded joint as a function of the length of the joint for two identical
metal substrates and a hard epoxy adhesive (Sikadur 31)
For a width of 10mm, the critical length for epoxy is equal to 60mm (Fig. 3). Consequently,
for a width of 55mm (Table 1) the critical length will be determined by the maximum load
supported by the concrete ( Fcritical concrete * b * criticalsteel 2MPa * 55mm * 60mm ). We can
evaluate on the Fig. 2 the corresponding critical length for load transfer equal to
criticalconcrete 150mm .As the assumption of perfect adhesion is considered the average stress
level for the epoxy is equal to 0.75MPa in this case.
Consequently, if we want to increase the shear load supported by connection, we made an
assembly using discontinuous blobs of adhesive, limiting the length of the join to the
efficient length and increasing the number of blobs and their width to obtain the global and
local shear resistance. The validation of this principle has been demonstrated by Hamelin et
al. [4].
3. DURABILITY PREDICTION OF ADHESIVE JOIN
3.1. Determination of the Creep Behaviour Law of the Adhesive by Mechanical
Spectrometry
As the average shear stress level in the adhesive join is low (1MPa) we will consider a
linear viscoelastic behaviour [7]. We used a METRAVIB type visco-elasticimeter which
allowed the harmonic periodic application of shear stress with a frequency ranging between
1Hz and 100Hz and temperature scanning from -20C to 120C (Fig. 4).

132

Tension test

Shear test

Flexural test

Fig. 4. METRAVIB Viscoelasticimeter


3.2. Results and Interpretation
The behaviour of the polymer becomes extremely visco-elastic as the vitreous transition
temperature is approached. Therefore, as illustrated in Fig. 5 for epoxy resin (Sikadur 31),
the passage of Tg is accompanied by a considerable drop in the mechanical properties. In
our case, the glass transition temperature is 42C.
Epoxy
1,00E+10

1,00E+00

G'
G"
tan delta

tan delta

G', G" (Pa)

1,00E+09

1,00E+08

1,00E-01

1,00E+07

1,00E+06
-20

20

40
60
Temprature (C)

80

100

1,00E-02
120

Fig. 5. Variation of G, G and Tg in function of temperature (epoxy resin)


Presentation of the results (Fig. 6) in the complex plane/ surface (Cole-Cole diagram)
shows that: The curve tends towards an asymptotic point G0 corresponding to the
instantaneous modulus. The curve also tends towards an asymptotic point Ginf
corresponding to the long-term modulus. This representation helps to establish a
rheological model describing the visco-elastic behaviour of the resin over time, frequency
and temperature.

133

3,00E+08
2,50E+08

G' (Pa)

2,00E+08
1,50E+08
1,00E+08
5,00E+07

Ginf

G0

0,00E+00
0,0E+00

5,0E+08

1,0E+09

1,5E+09

2,0E+09

2,5E+09

3,0E+09

G'' (Pa)

Fig. 6. ColeCole diagram (epoxy resin)


To account for the dissymmetrical shape of the curve experimentally obtained in the
complex plane (Fig. 7), the Zener biparabolic model seems to be the most appropriate, and
G 0 G inf
(7)
so the complex modulus is written as G * G
1 i h i k
is a parameter which takes into account the two periods of relaxation.

G0 - Ginf

G0
Ph
Pk

Fig. 7. Zener biparabolic model


The parameters , h, and k of this biparabolic model in parallel are determined by following
the process described by Huet [8] and Yong-Sok [10].
3.3. Experimentally Obtained Parameters
From the previously described tests on the viscoelasticimeter, the following values are
obtained for the epoxy resin, Sikadur 31:

h
0.11

k
0.47

Table 3. Constants of the rheological model

G0
1.3
8.58e7

134

Ginf
2.86e9

After integration, the analytical expression of the biparabolic model then becomes:
G 0 G
As a function of relaxation time of the resin
(8)
G t / G
1 t / h t / k
We can then draw the curve of the shear modulus over time for two constant temperatures:
+20C and +30C (Fig. 8).

Module
de modulus
cisaillement
(MPa)
(MPa)
Shear

3500

Rheological model 20C

3000

Rheological model 30C

2500
2000
1500
1000
500
0
1

10

100

1000

10000
100000
Time (s)
Temps (s)

1000000

1E+07

1E+08

1E+09

Fig. 8. The shear modulus over time for two temperatures for epoxy adhesive (creeping function)
3.4. Synthesis: Notion of Critical Shear Strength and Critical Transfer Length after
Adhesive Creep
Considering the results established above, we can recalculate the load supported by the
adhesive and the critical length for life time prediction of 50 years (Fig. 9). We note that the
maximum shear stress is limited at 6 MPa and the length increases from 60 to 100 mm.
5000
4500

S1: Es=210 000MPa ; Gadh=4 500MPa

4000

Load (kN)

3500
3000
2500

S2: Es=210 000MPa ; Gadh=1 500MPa

2000
1500
1000
500
0
0

50

100

150

200

250

Length (mm)

Fig. 9. Influence of adhesive creeping on load transfer for a life time prediction of 50 years
Now, if we consider the load transfer between concrete and adhesive taking into account
the variation of mechanical properties for the two materials during 50 years, Fig. 10

135

confirms that the maximum stress is limited to 1.0 MPa in the concrete and the critical
length to 250 mm.
45000
40000
C1: Ec=3 000MPa ; Gadh=4 500MPa ; adh=19MPa

35000
30000

C2: Ec=10 000MPa ; Gadh=1 500MPa ; adh=6MPa

Load (kN)

25000
20000
15000
10000

C3: Ec=30 000MPa ; Gadh=4 500MPa ; adh=6MPa

5000
0
0

50

100

150

200

250

Length(mm)

Fig. 10. Load transfer between concrete and steel for different durations
4. CONCLUSION: DESIGN DATA FOLLOWING EUROCODE
RECOMMENDATION
In as far as the failure of the adhesive joint occurs in the concrete or the adhesive, the shear
stress at the ULS of the interface is equal to
f tj

adhu ,d min adh * adh ,e ; with adh 0,8 if TG 50C et adh 0,4 if TG 50C (9)
adh td

Considering durability conditions the shear characteristics of the adhesive at the SLS:

adh, d min adh. f *

adh,e
adh

f ij
with adhf 0,4 at 20C if Tg > 50C
td
adhf 0,1 at 20C if Tg < 50C

(10)
(11)

In the case of load transfer by bonding joins between concrete and steel, we recommend a
discontinuous connection by adhesives blobs with a critical length equal at:
critical 150mm for t=0
and
critical 250mm for t=50years

136

5. CONCLUSION
We have established how it is possible to determine the resistance of an adhesive joint
according to the calculation of its efficient length in function of a design shear stress adu, d .
The evaluation of the weighting coefficient for the stress level of the adhesive and the
variation of the shear modulus have been determined by viscoelastic properties
measurements.
From a technological point of view we recommend gluing by blobs incorporating zones of
hard adhesive (epoxy) and zones of flexible adhesive (polyurethane mastic). Following this
principle of discontinuous assembly, the proposed dimensioning method aims to verify the
equality between the maximum shear flux induced either by the resistance of the concrete
in compression or the steel in tension and the sum of the shear forces supported by each
blob.
6. REFERENCES
1.

Meier, U. 1995. Strengthening of structures with carbon fiber epoxy laminates.


Construction and Building Materials, 9(6): 341-351.
2. Varastehpour H. and Hamelin P. 1996. Experimental study of RC beams strengthened
with FRP plates, Proceeding of ACMBS-2, Canadian Society for Civil Engineering,
Montreal, Quebec, Canada, pp. 527-536.
3. Taljsten, B., Carolin, A., and Nordin, H., 2003. Concrete structures strengthened with
near surface mounted reinforcement of CFRP, Journal of Advanced Structural
Engineering, 6(3): 201-213.
4. Hamelin, P. 2010. La connexion par collage p227-254, Pont mixtes Acier Bton, Un
guide pour des ouvrages innovants, Projet National MIKTI Presse des ponts ISBN
978-2-85978-449-2, 457 p.
5. Volkersen O. 1965. Recherche sur la thorie des assemblages colls, Construction
mtallique (No. 4), pp. 3-13.
6. Schapery, R.A. 1996 Characterization of non linear time dependant polymers and
polymeric composites for durability analysis, Progress in durability analysis of
composite Systems, Edited by Cardon-Fukuda, pp. 21-36.
7. Brinson, L.C., Gates, T.S. 1995. Effects of physical aging on ling term creep behaviour
of polymers and polymers matrix composites, International Journal of Solid and
Structures, 32: 827-846
8. Huet, C. 1963. Etude par une mthode dimpdance du comportement viscolastique
des matriaux hydrocarbures, Thse Doct. Ing. Facult des sciences de lUniversit de
Paris, 210 p.
9. Hamelin, P. 1979. Contribution ltude du comportement rhologique de liants
viscolastiques en vue de lanalyse du fluage des matriaux composites utiliss sous
forme de structures du Gnie Civil, Thse de doctorat es Sciences, Universit Claude
Bernard, 161 p.
10. Yong, S. 1984. Modelisation of viscoelastic behaviour of carbon-epoxy unidirectional
composite. European Mechanics Colloquium (182), Brussel.

137

138

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

DURABILITY STUDY OF GLASS FRP BARS SUBJECTED TO


EXTREME ENVIRONMENTAL AND LOADING CONDITIONS
M. Robert1, A. Fam2 and B. Benmokrane3
1

Postdoctoral Fellow, Department of Civil Eng., University of Sherbrooke, Quebec, Canada


Canada Research Chair Professor, Dept. of Civil Eng., Queens University, Kingston, Canada
3
Canada & NSERC Research Chairs Professor, Department of Civil Engineering, University of
Sherbrooke, Quebec, Canada
2

ABSTRACT
Fiber-reinforced polymer (FRP) materials have emerged as a practical alternative material
for producing reinforcing bars for concrete structures. This is due to their relatively low
cost-to-performance ratio and noncorrosive nature compared to traditionnal steel
reinforcing bars. In addition, FRP materials exhibit properties, such as high tensile strength,
that make them suitable for use as structural reinforcement. However, their durability in an
alkaline environment is still of concern and factors that can affect the long-term behavior of
GFRP materials have to be investigated. Moisture, high pH of surrounding environment,
extreme temperature and the presence of microcracks can affect the long-term properties of
FRP reinforcing bars. This paper summarizes the long-term durability study of glass fibrereinforced polymer (GFRP) reinforcing bars subjected to different conditionnings
simulating the real environment of application. In particular, the behavior of GFRP bars
subjected to moisture, high pH of moist concrete, extreme temperatures, and to tensile
prestress was investigated showing the great durability of GFRP reinforcing bars.
1. INTRODUCTION
Many steel reinforced concrete infrastructures exposed to harsh environments present
durability issues and are likely not to reach the lifetime expected or have already attained
the service limit. Steel corrosion induced by chlorides and/or carbonation are the main
causes of deterioration of steel reinforced concrete structures. The amplitude of the
phenomenon and economic factors have pushed research engineers, industries and public
institutions to develop new technologies to better protect reinforcing steel or providing non
corrosive materials in replacement of steel [1]. In anticipation of offering an alternative
material, research works have been conducted to develop and use fibre reinforced polymer
(FRP) composites in civil engineering applications.
One of the biggest challenges for the acceptance of FRP in civil engineering applications
concerns their durability and more specifically their capacity to maintain their structural
139

performance in severe and changing environmental conditions of service [2]. The


degradation process of GFRP reinforcing bars may be caused by the deterioration of one (or
more) of its phase: fibre, resin matrix or interface [3]. The degradation mechanisms are
more complex compared to those occurring in a monophasic material, such as steel. The
diffusion of moisture and alkalis from the concrete into the polymer matrix can lead to
changes in the mechanical, thermomechanical and physico-chemical properties of GFRP
reinforcing bars [4].
This paper presents an overview of several aspects affecting the long-term durability of
GFRP reinforcing bars subjected to laboratory accelerated aging simulating the genuine
environment of application. The durability of GFRP bars submitted to various
environmental conditions such as wet concrete, extreme temperatures and microcracksinducing pre-loading will be evaluated though the residual mechanical properties.
2. EXPERIMENTAL PROGRAM
2.1 Material
Experimental results presented in this paper were obtained during research work conducted
in the laboratory of durability of FRP composite materials of the department of civil
engineering of the University of Sherbrooke, Sherbrooke, Canada. Sand-coated glass FRP
reinforcing bars manufactured by a Canadian company (Pultrall inc., Thetford Mines,
Quebec, Canada) are used in this study. The bars are made of continuous E-glass fibres
embedded in vinylester matrix using a pultrusion process. The GFRP bars used have a
nominal diameter of 12 mm.
2.2 Test Plan
2.2.1 Aging in Moist Concrete
The aging was performed by immersing the mortar-wrapped GFRP bars in a wood
container specially manufactured for the study. The water level was kept constant
throughout the study to avoid a pH increase which could be due to a water level decrease
and a significant increase of the concentration of the alkaline ions in the solution. The
specimens were completely immersed at three different temperatures (23, 40 and 50oC) and
were removed from the water after four different periods of time (60, 120, 180 and 240
days). After each period, usually six GFRP bars were removed from the water, tested under
tension to compare their tensile strength retention values to those of the reference
specimens. Microstructural and physical analyses were also performed after immersion.
2.2.2 Extreme Temperatures
Properties of GFRP bars submitted to extreme temperatures were measured. All GFRP bar
specimens were tested under extreme temperature conditions using a MTS environmental
chamber cooled by liquid nitrogen. Sub-zero temperatures equal to -100o, -80o, -60o, -40o, 20o and also 0oC were chosen to simulate the effect of Nordic climate on GFRP bars. The
temperatures below -60 oC were used to amplify the effects of temperature on mechanical
140

and microstructural behaviour of the FRP materials. Elevated temperatures equal to 50,
100, 150, 200, 250 and 325oC were chosen to simulate short-term environmental conditions
in the case of fire. All the tests were conducted using a steady-state temperature (heat then
load) regime.
2.2.3 Pre-Loading of GFRP Bars Samples
The effect of cracking and microcracking on the long-term behaviour of GFRP bars was
evaluated. GFRP bars were preloaded at 80 percent of their theoretical ultimate tensile
strength (854 MPa) (UTS) to initiate cracks and microcracks in polymer and glass fibres.
All bars were preloaded under tension according to ASTM D 7205 standard. The rate of
loading ranged between 250 and 500 MPa/min and the maximum load was maintained for
10 minutes, and then was reduced at the rate of 250 to 500 MPa/min. The study of the longterm behaviour of pre-loaded GFRP bars was performed by using accelerated aging in
moist moist concrete, as described above.
2.3 Mechanical Characterization
Tensile tests were performed on bars aged in solution, bars subjected to controlled
temperatures and bars pre-loaded at 80 percent of their UTS. The measured tensile
strengths of the bars before and after exposure were considered as a measure of the
durability performance of the specimens. All bars were tested under tension according to
ASTM D 7205 standard. Each specimen was instrumented with a Linear Variable
Differential Transformer (LVDT) to capture the elongation during testing. The rate of
loading ranged between 250 and 500 MPa/min. The applied load and bar elongation were
recorded during the test using a data acquisition system monitored by a computer. Due to
the brittle nature of GFRP, no yielding occurs and the stress-strain behaviour was linear.
2.4 Physico-chemical Characterization
Scanning Electron Microscopy (SEM) observations and image analysis were performed to
evaluate the microstructure of aged specimens and the integrity of the GFRP material after
different conditionings. These observations were conducted to determine the potential
degradation of the polymer matrix, possible glass fibers and interface.
Glass transition temperature was determined by Differential Scanning Calorimetry (DSC)
in accordance with ASTM E 1356 standard for both the reference specimens and specimens
exposed to conditionings. If decrease in Tg was observed for conditioned samples, it was an
indication of plasticizing effect or chemical degradation. Aged sample maintaining a lower
Tg than for the reference showed an irreversible chemical degradation.
Thermal stability of specimens was investigated by using Thermogravimetric Analysis
(TGA), and the weight loss traces were recorded as function of temperature in the range of
20800 C according to ASTM E 1868 standard. The mass variations recorded could give
indications concerning the different phenomena of degradation occurring in the matrix and
also explain the loss of mechanical properties observed at very high temperatures.

141

Chemical degradation of polymer matrix was characterized by using Fourrier Transform


Infrared Spectroscopy (FTIR). The most interesting region of the FTIR spectra is located
between 3300 cm-1 and 3600 cm-1, which corresponds to the stretching mode of the
hydroxyl groups of the vinylester resin. When a irreversible chemical degradation occurs,
such as polymer hydrolysis reaction, new hydroxyl groups are formed and the
corresponding infrared band increases. Changes in the peak intensity were quantified by
determining the ratio of the OH peak to the carbon-hydrogen stretching peak of the resin,
which is not affected by the conditioning.
2.5 Arrhenius Relation
Equation (1) expresses the Arrhenius relation, in terms of the degradation rate [5]
Ea
k A exp
(1)

RT
where k = degradation rate (1/time); A = constant relative to the material and degradation
process; Ea = activation energy of the reaction; R = universal gas constant; and T =
temperature in Kelvin. The primary assumption of this model is that only one dominant
degradation mechanism of the material operates during the reaction and that this
mechanism will not change with time and temperature during the exposure [6]. Only the
rate of degradation will be accelerated with the temperature increase. Equation (1) can be
transformed into:
1 1
E
exp a
k A
RT
1 E 1
ln a ln A
k R T

(2)
(3)

From Equation (2), the degradation rate k can be expressed as the inverse of time needed
for a material property to reach a given value [6]. From equation (3) one can further
observe that the logarithm of time needed for a material property to reach a given value is a
linear function of 1/T with the slope of Ea /R [6]. Ea and A can be easily calculated by using
the slope of the regression and the point of intersection between the regression and the Y
axis respectively.
3. EXPERIMENTAL RESULTS
3.1 Effect of Immersion
3.1.1 Tensile Strength Retention
Table 1 shows the experimental results obtained during the tensile tests concerning the
ultimate strength of reference and pre-loaded samples aged in moist concrete and tested
after immersion. Figure 1 shows the retention of the ultimate strength of reference and preloaded bars aged in moist concrete according to the duration of immersion of embedded
bars at various temperatures. Table 1 shows a slight decrease of the ultimate tensile strength
142

with immersion duration. The recorded results show that the longer the time of immersion,
the larger the loss of resistance. Furthermore, it is clear that the temperature of immersion
affects the loss of resistance. It can be seen that for duration of immersion of 8 months, the
loss of resistance for aged reference bars is equal to 16, 10 and 9 % at 50o, 40o and 23oC,
respectively. Results concerning aged pre-loaded bars presented in Table 1 show slight
decrease (6 to 11%) of the ultimate tensile strength after 240 days of immersion of preloaded bars embedded in moist concrete. This decrease was similar to the loss of tensile
strength measured on same bar subjected to same conditionings but without preloading [7].
This phenomenon was due to the increasing of diffusion rate of the solution into the sample
due to the presence of cracks and microcracks and to the acceleration of chemical reactions
of degradation with the temperature of immersion, leading to a larger absorption rate of the
solution for the same time of immersion. It was also noted that the loss of elastic modulus
was negligible for both, reference and pre-loaded bars aged in moist concrete.
Table 1. Experimental tensile strength of reference and pre-loaded specimens aged in
moist concrete
Time of immersion
(days)
0
60
120
180
240

Temperature
(oC)
23
23
40
50
23
40
50
23
40
50
23
40
50

Reference Bars
Mean Tensile
COV
Strength (MPa)
(%)
788
6,9
753
8,2
755
3,7
767
3,8
702
2,0
666
7,8
720
3,2
717
2,6
708
4,8
711
2,5
714
3,5
708
7,1
665
9,3

Pre-Loaded Bars
Mean Tensile
COV
Strength (MPa)
(%)
854
2.1
846
5.6
847
6.4
838
4.2
849
2.5
832
8.1
837
3.9
836
3.3
823
5.7
808
3.1
810
4.2
784
7.9
768
9.0

a)
b)
Fig. 1. Tensile strength retention for: a) reference bars aged in moist concrete, and b) preloaded bars aged in moist concrete

143

The tensile test of unconditioned specimens, specimens aged in moist concrete at 50oC
during 240 days and pre-loaded specimens aged in moist concrete at 50oC during 240 days
showed an approximately linear behaviour up to failure. Specimens failed through the
rupture of fibres. Micelli and Nanni [8] also observed similar tensile failure modes of
GFRP bars. The failure was accompanied by the delamination of fibres and resin, as shown
in Figure 2. From this observation, it could be concluded that the aging in moist concrete or
the presence of pre-existing cracks and microcracks due to the pre-loading of bars, have no
significant effects on the failure mode occurring during tensile tests.

Reference

Aged in moist concrete

Fig. 2. Typical failure mode for GFRP bars


3.1.2 Microstructural Effect of Pre-Loading
The micrographs of Figure 3 show the longitudinal bar surface of reference and GFRP bar
pre-loaded at 80 percent of the UTS. Observations of fibres/matrix interfaces and of the
microstructure in general demonstrated that the pre-loading affected the microstructure of
GFRP bar. The only visible damage occurred at the fibre level since the elongation at the
rupture was lower for the fibres compared to the polymer matrix. No damage occurred at
the matrix and fibre/matrix interface.

a)
b)
Fig. 3. Micrograph of longitudinal GFRP bar: a) Reference GFRP bar at low magnification,
and b) GFRP bar loaded at 80% of the UTS at high magnification
3.1.6 Microstructural Effect of Aging in Moist Concrete
The micrographs of Figure 4 show the fibres/matrix interface for reference and pre-loaded
bars aged in moist concrete at 50oC during 240 days. Observation of these interfaces, and of
144

the microstructure, in general, demonstrate that the conditionings of mortar-wrapped bars in


water do not affect the microstructural properties of the GFRP bars, even if the bars show
pre-existing cracks or microcracks.

(a)
(b)
Fig. 4. Micrographs of fibre/matrix interface (X3000): a) Unconditioned bars, b) Preloaded GFRP bar specimen aged in moist concrete at 50oC during 240 days.
3.1.7 Effect on Polymer Matrix
The change in the Tg of GFRP bars embedded in concrete after aging in water was also
investigated. No significant change of the Tg was measured after the aging of reference and
pre-loaded bars in moist concrete during 240 days at 50oC. In fact, the Tg of reference
specimen was 127oC, the Tg of reference specimens aged in moist concrete was 129oC, and
the Tg of pre-loaded specimens aged in moist concrete was 125oC. These results clearly
show that even after 240 days in salt solution at 50oC, the polymer matrix is not affected.
A FTIR analysis of unconditioned bar specimen, and preloaded (80% of the UTS) mortarwrapped bars and aged in moist concrete at 50oC for 240 days was conducted (Figure 5).
The experimental ratio of the OH peak to the carbon-hydrogen stretching peak of the resin
for the 12.7 mm diameter preloaded concrete-wrapped samples and immersed in water for
240 days at 50oC was 0.53 compared to 0.51 for unconditioned samples. So, the hydroxyl
peak did not show any significant changes. This observation led to the conclusion that no
chemical degradation of the polymer occurred during the immersion of preloaded concretewrapped bars.

Fig. 5. FTIR spectra of unconditioned specimens and pre-loaded specimens aged in moist
concrete
145

3.1.8 Long-Term Predictions


Following the procedure proposed by Bank et al. [9], the natural logarithm of time to reach
a set of levels of normalized performances versus 1/T, expressed as the inverse of absolute
temperature (1000/K), was used to predict the service life at the Mean Annual Temperature
(6oC) in Montral, Qubec, Canada. A coefficient of determination (R2) value close to 1 is
desired. However, the ASTM procedures recommend a minimum value of 0.80 for
acceptability and the obtained R2 values are between 0.94 and 0.99. From the Arrhenius
plot, the service life time necessary to reach the established tensile strength retention levels
(PR) can be extrapolated for any temperature. Consequently, predictions are made for
tensile strength retention as a function of time for an immersion at 6C and the general
relation between the PR and the predicted service life at the average temperature of 6oC are
drawn (Figure 6). The predicted time to reach the determined strength property retention
level (PR) for GFRP bars embedded in moist concrete without any preloading at an
isotherm temperature of 6oC is approximately 1 and 210 years for a PR of 90 and 75%,
respectively. Also, the predicted service life of GFRP bars embedded in moist concrete at
an isotherm temperature of 6oC to reach a PR less than 75% can be estimated to be infinite.
These predictions show that the GFRP bars are durable face to the concrete environment,
which is supposed to be well simulated by the immersion of embedded GFRP bars in water.

Fig. 6. General relation between the PR and the predicted service life at a mean annual
temperature of 6oC (Montreal, Quebec)
3.2 Effect of Extreme Temperatures
3.2.1 Effect of Extreme Temperatures on Mechanical Properties
Figure 7 shows the relation between the tensile strength of GFRP bars and the test
temperature respectively. Three different zones were identified in Figure 7: 1) the reference
plateau between -40o and 50oC; 2) the zone where the stiffness increasing at temperatures
lower than -50oC; and 3) the mechanical strength decreasing zone for temperatures near and
above the glass transition temperature (Tg) of the polymer. The mechanical strengths
between -40o and 50oC were stable. In this range of temperatures, the molecular chain
mobility of the polymer did not change since the temperature was below Tg. This result
146

showed that for standard environmental conditions of Nordic countries as Canada and north
of USA (temperature ranging from -40 to 50 oC), the mechanical properties of GFRP bars
were not changed. For temperatures lower than -50oC, the molecular chains mobility of the
polymer decreased, leading to an increase of the mechanical stresses needed to rupture the
material. Below Tg (around 120oC), the matrix was in a glassy state. When increasing the
temperature and reaching the decomposition region, the breaking of molecular bonds
started and the ductility of the material increased, leading to a decrease of mechanical
strengths and stiffness of the material. At very high temperatures (greater than 300oC),
strong degradation of the polymer occurred (combustion, oxidation) and load transfer
provided by the matrix was severely reduced.

Fig. 7. Tensile strength of GFRP bars tested at different temperatures


3.2.2 Effect on Polymer Matrix
Figure 8 shows the mass variation of specimens as a function of temperature measured by
TGA.

Fig. 8. Mass variation of GFRP bar specimen when heated between 20o and 800oC.
147

It was seen that major weight loss occurred between 300 and 450oC. This important drop,
up to 18%, was due to the thermal degradation of the polymer. At these temperatures,
thermal degradation occurred and the molecular chains of the polymer broke, leading to the
formation of microcracks both at the fibre/matrix interface and in the matrix phase (Figure
6). The weight losses measured by TGA showed a major degradation of the resin from
300oC to 450oC. Below this range of temperatures of degradation, the properties were
recovered when the GFRP specimens were heated to the desired temperature and then
tested at the ambient temperature. At this point, the experimental results suggest that
irreversible loss of mechanical properties occurred above 300oC.
The Tg was measured by DSC for reference and specimen heated at 350oC for 2 hours. The
Tg of the reference specimen was equal to 113oC, whereas the specimens heated at 350oC
for 2 hours was equal to 67oC. This showed that the matrix weakened when heated at
elevated temperatures, which explained the causes of the observed large decrease of
mechanical properties.
3.2.3 Microstructural Effect of Extreme Temperatures
The polymer matrix degradation was confirmed by microstructural analysis. The
micrographs presented in Figure 9 show the polymer, the fibre and the interface between
the fibres and matrix for reference specimen and specimen conditioned at 350oC for two
hours. The comparison of these micrographs showed that significant damage occurred for
GFRP bar conditioned at 350oC (Figure 9b) as compared to reference sample (Figure 9a).
The presence of microcracks in specimen tested at 350oC confirmed the degradation of the
polymeric matrix and explained the loss of mechanical properties.

(a)
(b)
Fig. 9. Micrographs of longitudinal fibre/ matrix interface of: a) unconditioned GFRP bar
specimen, b) Conditioned GFRP bar specimen aged in air at 350oC for 2 hours.
4. CONCLUSIONS
Based on the results of this study the following conclusions may be drawn:
1- Even at high temperature (50oC), where the environment is the more aggressive, the
change in tensile strength of GFRP bars embedded in moist concrete is minor.

148

2- No significant microstructural changes were observed after 240 days immersion of


reference and pre-loaded GFRP bars embedded in concrete in tap water at 50oC.
3- The polymer matrix is not affected by moisture absorption and high temperatures: no
changes of the glass transition temperature occur as observed by DSC and no
irreversible degradation occurred to the polymer matrix as observed by FTIR. The
preloading of GFRP bars didnt lead to any further degradation.
4- After 240 days water immersion of pre-loaded bar embedded in mortar, the retention
rate of tensile strength are 95, 96 and 98% at 50, 40, and 23oC, respectively.
5- According the to long-term predictions, the tensile strength retention of GFRP bars
embedded in moist concrete will decreased by 20 and 25% after 100 and 200 years
respectively. It was shown that the service life time allowed to reach tensile strength
retention of less than 50% should be infinite.
6- Tensile properties of GFRP bars increased when the temperature decreased. This
phenomenon was due to the increase of stiffness of the amorphous polymer matrix due
to low temperatures.
7- At severe temperatures experienced in Northern regions such as Canada (temperature
ranging from -40o to 50oC), the tensile strength appeared to be stable, which showed that
the mechanical behavior of GFRP bars was not affected by temperature in this range of
temperature
5. REFERENCES
1.
2.
3.
4.
5.
6.
7.
8.
9.

Nkurunziza, G., Benmokane, B., and Debaiky, A.S., Masmoudi, R. 2005. Effect of
sustained load and environment on long-term tensile properties of glass FRP
reinforcing bars. ACI Structural Journal, 102(4): 615-621.
Robert, M., Cousin, P., Wang, P., Benmokrane, B. 2010. Temperature as an
Accelerating Factor for Long-Term Durability Testing of FRPs: Should there be any
limitations? Journal of Composites for Construction, 14(4): 361-367.
Benmokrane B, Wang P, Ton-That T, Rahman H, Robert J. 2002. Durability of glass
fibre reinforced polymer reinforcing bars in concrete environment. Journal of
Composite for Construction, 6(2): 143153.
Benmokrane, B., Wang, P., Pavate, T., and Robert, M. 2006. Durability of FRP
Composites for Cvil Infrastructure Applications, Chapter 12. Whittles Publishing,
Scotland, pp. 300-343.
Nelson, W. 1990. Accelerated testingStatistical models, test plans, and data analyses.
Wiley, New York, 601 p.
Chen, Y., Davalos, J. F., Ray, I. 2006. Durability Prediction for GFRP Bars Using
Short-Term Data of Accelerated Aging Tests. Journal of Composites for Construction,
10(4): 279-286.
Robert, M., Cousin, P., and Benmokrane, B. 2009. Durability of GFRP Bars Embedded
in Moist Concrete. Journal of Composites for Construction, 13(2): 66-73.
Micelli, F., and Nanni, A. 2004. Durability of FRP rods for concrete structures.
Construction Building Materials, 18(7): 491503.
Bank, L.C, Gentry, T.R, Thompson, B.P, Russel, J.S. 2003. A Model Specification for
Composites for Civil Engineering Structures. Construction and Building Materials,
17(6-7): 405-437.

149

150

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

FIRE TESTING OF FRP STRENGTHENED REINFORCED


CONCRETE COLUMNS
N. Bnichou1, D. Cree2, E.U. Chowdhury2, M.F. Green2 and L.A. Bisby3
1

Institute for Research in Construction, National Research Council of Canada, Ottawa, Canada
Dept. of Civil Engineering, Queens University Kingston, Ontario, Canada
3
Institute for Infrastructure and Environment, University of Edinburgh, Scotland, UK
2

ABSTRACT
This paper presents the results of full-scale fire tests on two concrete columns strengthened
with fibre reinforced polymer sheets and protected with a new insulation system. One
column was circular while the other was square. The columns were subjected to the ULCS101 standard fire. Both columns obtained fire endurance ratings of over 4 hours. At the
end of the fire test, the columns were loaded to failure. The temperatures in the columns
during the fire test and the load capacities of the columns will be presented. Additional
work is required to compare the results against thermal and structural numerical models
developed specifically for circular and square columns.
1. INTRODUCTION
Fibre reinforced polymers (FRPs) are now widely applied for repairing concrete structures
[1, 2]. FRP are known to have excellent strength, stiffness, and corrosion resistance, but
their mechanical properties are significantly reduced at elevated temperatures due to the
properties of the matrix resin. As such, concerns regarding fire resistance have limited FRP
applications in buildings. To address this concern, a new fire insulation material is applied
on the surface of FRP wraps to obtain appropriate fire resistance. Novel materials for
potential use inside a building, including FRPs and insulation materials must pass certain
fire exposure limits as prescribed in North America by CAN/ULC S101 [3] or ASTM E119
[4]. The objective is for the structure protected with the new insulation material to
withstand a fire resistance (fire endurance) test by being exposed to a standard fire for a
specific duration, under strict conditions. Until the fire endurance property of this unique
supplemental insulating fire protection system is known, it will hinder its extensive use as a
reinforcement coating on the interior of buildings.
2. MOTIVATION
To gain insightful information on fire endurance, there has been an on-going collaborative
research program in the development of various insulating materials for protecting FRP
151

wrapped reinforced concrete columns in fire conditions. The initial tests consisted of two
circular columns (Column 1 and Column 2) [5], followed by another three circular columns
(Column 3, Column 4 and Column 5) [6]. The current study reports the work done on one
circular column (Column 5) [6] and one recently tested square column (Column 6). A team
comprising of the National Research Council of Canada (NRC), Queens University,
Canada and industry partner Sika Canada are collaborating on the current research project.
This project involves the full-scale fire test of two concrete columns strengthened with FRP
wrap and insulated with a supplemental fire protective coating. The column fabrication,
FRP strengthening, the insulation fire protection material and fire test results are discussed.
The objectives for these fire tests are to investigate the behaviour of FRP strengthened
reinforced concrete columns protected with an insulation layer and subjected to elevated
temperatures. The strengthened and insulated columns were subjected to the ULC-S101
standard fire, under sustained service load, which simulates the load that the columns
would be expected to experience in an actual fire situation. The results of both experiments
(Columns 5 and 6) are presented in this paper.
3. EXPERIMENTAL PROCEDURE
3.1 Column Test Specimens
The experimental program consisted of two fire tests on one circular column and one
square reinforced concrete column strengthened with the SikaWrap Hex 103C FRP
externally-bonded strengthened system and insulated with Sikacrete-213F, a sprayapplied fire protection mortar. The columns were designated as Column 5 and Column 6,
respectively and photos of the columns before fire testing are shown in Figure 1. Column 5
was 400 mm in diameter while the cross-section of Column 6 was 305 mm by 305 mm.
Both columns had a height of 3810 mm and were designed for 28 MPa concrete strength
using Type 10 Portland cement with crushed carbonate aggregate, maximum aggregate size
of 14 mm, and natural sand as the fine aggregate.
The longitudinal steel reinforcement in Column 5 consisted of eight 20M (19.5 mm
diameter) deformed bars, symmetrically placed with 40 mm clear cover to the spiral
reinforcement. The lateral reinforcement for the columns consisted of 10M (11.3 mm
diameter) deformed steel spiral with a centre-to-centre pitch of 50 mm. The longitudinal
reinforcing bars and the steel spiral had average yield strengths of 456 MPa and 396 MPa,
respectively. The primary longitudinal reinforcement bars for Column 6 were four 25M
deformed steel bars. The lateral reinforcement for the column consisted of 10M deformed
steel bars spaced at 305 mm. The main reinforcing bars and ties all had average yield
strengths of 477 MPa. The steel reinforcement had a clear cover of 40 mm from the exterior
surface of the concrete to the steel ties and a 50 mm concrete cover to the principal
reinforcement. Column 5 was poured vertically into a 400 mm inside-diameter Sonotube,
while Column 6 was cast vertically in a plywood formwork. The columns were
instrumented with Chromel-alumel (Type-K) thermocouples to record the internal
temperatures within the concrete, the reinforcing steel bars, the concrete-FRP interface, the
FRP-insulation interface and the outer surface of the insulation.

152

3.2 FRP Strengthening and Fire Protection


Both columns were strengthened with a SikaWrap Hex 103C unidirectional carbon/epoxy
FRP strengthening system with a Sikadur 300 resin epoxy saturant/adhesive. Column 5 was
strengthened with two layers of FRP wrap while Column 6 was strengthened with three
layers. The carbon FRP has an ultimate tensile strength of 849 MPa and tensile elastic
modulus of 70.5 GPa. To provide fire protection, Sikacrete-213F, a supplemental fire
insulation material, was spray-applied over the surface of the FRP wrapped columns. Prior
to the spray application for Column 5, a 1/8 th inch diameter steel mesh, having 50 mm by
50 mm openings, was attached to the surface of the column to provide reinforcement for the
protection system. Column 5 had an average insulation thickness of 44 mm, while Column
6 had an average insulation thickness on flat surfaces of 40 mm and an average corner
insulation thickness of 51 mm.
3.3 Fire Endurance Test
The fire endurance experiments were conducted by exposing the columns to heat in a
furnace specially built for fire testing loaded columns. The columns were installed in the
furnace by bolting the end plates to the test frame loading head at the top and a hydraulic
jack at the bottom. This resulted in a fixed-fixed end condition for the column, which
simulates the end conditions to be expected in an actual building. Column 5 was loaded in
compression until the applied load on the circular column was 3031 kN, and Column 6 until
the applied load on the square column was 1717 kN. The respective loads were maintained
at a constant value throughout the fire endurance test. Both columns did not fail under their
respective sustained service loads, therefore after 4 hours of fire exposure; the loads were
increased until failure was observed.

(a)
(b)
Fig. 1. Images of the FRP-wrapped and insulated concrete columns (a) circular and (b)
square, prior to fire test.
153

4. RESULTS AND DISCUSSION


4.1 Temperature Behaviour
The fire behaviour of the insulation, FRP wrap, and crack propagation in the insulation in
Column 5 and Column 6 were monitored during the test through small observation
windows (view ports) in the walls of the furnace. Prior to commencing the fire tests, it was
observed on both columns that minor cracks developed on the surface of the dried
insulation, which were due to shrinkage. The initial cracks were approximately less than 5
mm wide but widened to about 10 mm as the test progressed. For both columns, flames
were seen emanating through the small cracks on the surface of the insulation at
approximately 2 hours into the fire test. Cracks on the surface of the insulation, prior and
during the fire endurance test, are depicted in Figure 2. The steel mesh reinforcement
within the insulation of Column 5 may have prevented the spalling of the insulation from
the column surface, however, no spalling of the insulation system was observed from
Column 6 which did not contain a mesh. Prior to the column failures, the fire protection
insulation system remained intact with minor external cracks throughout the length of the
column for the duration of the fire test.
The temperature results for Column 5 and Column 6 at mid-height for horizontally
positioned thermocouples are given in Figure 3 and Figure 4, respectively. The temperature
measurement readings are for the insulation surface, FRP surface, concrete surface and
internal to the concrete. Both Figures 3 and 4 also include the standard fire curve. During
the test on Column 5, a dip in the average furnace temperature was a result of an abrupt
furnace stop at 3 hours into the fire test and required 12 minutes to bring the temperature
inside the chamber back to the prescribed temperature-time curve.

(a)
(b)
Fig. 2. (a) Typical cracks on the surface of the fire insulation system prior to the fire test,
and (b) flame from the cracks on the surface of the fire insulation system during the fire
test.

154

Insulation surface
Concrete surface
Cover 150 mm

1000

Temperature (C)

Temperature (C)

ASTM E-119
FRP surface
Cover 50 mm

ASTM E-119
Concrete surface corner
Concrete surface
Cover 153 mm

800
600
400
200

Insulation surface
FRP surface
Cover 63 mm

1000
800
600
400
200
0

0
-1

2
3
4
Time (hours)

-1

Time (hours)

Fig. 3. Column 5 temperature of thermocouples


at mid-height as a function of time.

Fig. 4. Column 6 temperature of thermocouples


at mid-height as a function of time.

The polymer resin (Sikadur 300) has a glass transition temperature (Tg) of 60C if cured at
ambient temperature. The concrete surface temperature for Column 5 (Figure 3) was shown
to have an average temperature of 60.5C after 29 minutes of fire exposure. This
demonstrated that the insulation system limited the time at which the temperature was
below the Tg of the polymer resin, which is important for the bond performance. The
temperatures recorded at the concrete surface increased at a steady rate until they reached a
plateau at 100C. This plateau is due to moisture evaporation within the insulation. After
approximately 1 h 30 min, the concrete surface temperature began to increase at a higher
rate up to 266C after 4 hours. The temperature at the concrete surface remained below the
manufacturer stated Tg for about 29 minutes, by providing a thickness of approximately 44
mm of insulation. After 4 hours, the highest temperatures recorded at the FRP surface and
concrete surface were 408C and 266C, respectively.
Column 6 had an average concrete surface temperature of 60.1C (Figure 4) after 33 minutes
of fire exposure. Similarly, the corner thermocouples at mid-height detected an average
concrete surface corner temperature of 60.2C (Figure 4) after 22 minutes of fire exposure.
This showed that the corner FRP increased in temperature quicker than the sides. The data
demonstrated that it is possible to maintain the temperature of the concrete surface below
the Tg of the FRP material for about 0.5 hrs by providing an insulation system of
approximately 40 mm thick. After 4 hours, the highest temperatures recorded from the
horizontally positioned thermocouples at the FRP surface and concrete surface were 406C
and 250C, respectively.
For both columns, the test data indicated the insulation system would be unable to maintain
the FRP temperature below the Tg of the polymer resin for prolonged periods of time during
fire. However, the insulation system was able to maintain the temperatures within the
concrete and reinforcing steel below 200C. For example, Figure 3 shows that after 4 hours,
the temperatures in Column 5 on the concrete surface, in the concrete at 50 mm from
surface (at the location of the reinforcement), and in the concrete at 150 mm from surface
are 266C, 148C and 98C, respectively. Similarly, Figure 4 shows that, for Column 6
after 4 hours, the concrete surface, the reinforcement temperatures, the concrete at 63 mm
from the surface, and concrete at 153 mm from the surface are 250C, 193C 164C and
155

136C, respectively. The insulation system was able to maintain the temperatures within
the concrete and reinforcing steel of both columns below 200C. By maintaining this
criteria, no significant deterioration in the mechanical properties of concrete and steel
occurred. Hence, it can be stated that the columns maintained their full unconfined axial
load carrying capacity for greater than 4 hours of exposure to the ASTM E119 standard
fire.
4.2 Structural Behaviour

Column 5-Axial displacement

Column 6-Axial displacement

Column 5-Axial load

Column 6-Axial load

5
0
-5
-10
-15
-20
-25
-30
-35
-40
-45
-50

0
-1000
-2000
-3000
Preload
phase

Fire test
phase

-4000

Applied load (kN)

Axial deformation (mm)

Figure 5 shows the axial deformations and load applied as a function of time for Columns 5
and 6 during the preload and fire test phase. A negative axial deformation or applied load
value indicates compression of the column. The columns experienced an initial load of
approximately 350 kN due to the weight of the hydraulic jack prior to the start of the
preload phase. Column 5 resisted the sustained concentric load of 3054 kN for 4.5 hours.
Therefore, after 4.5 hours, the load was steadily increased until the column reached failure
at a load of 4984 kN. The column failed by apparent crushing of the concrete core.
Similarly, Column 6 resisted the sustained fire load of 1717 kN for 4 hours. Since it did not
fail, the load level was increased to 2641 kN, at which point the column failed in a nonviolent fashion by crushing of the concrete core, which resulted in the insulation
delaminating at the centre of the column, while the insulation remained largely intact at the
extremities of the column. The temperature data obtained during the fire endurance test of
Column 5 and Column 6 show that the FRP wrap had been rendered ineffective even with
an insulation layer due to increased temperatures in excess of its glass transition
temperature. However, due to the superior thermal insulation system, the concrete and
reinforcing steel retained their room-temperature strength.

-5000
-2

-1

2
3
Time (hours)

Fig. 5. Applied load and axial deformation as a function of time for Columns 5 and 6.
The axial deformations of the columns were the result of a combination of load effects and
thermal expansion. During the preload phase of the fire test, Column 5 had an initial 7.3
mm axial deformation due to the applied value of 3054 kN. After the preload phase, the
axial load (3054 kN) was kept constant for 4.5 hours of fire exposure. No major expansion
was observed during heating up to 4.5 hours due to the relatively low temperatures in the
156

concrete. For instance, after 3 hours of fire exposure, the temperature of the rebar for
Column 5 was 105C and the concrete surface temperature was 225C. However, a severe
drop in axial deformation after 4.5 hours occurred from the increasing applied load up to
failure. At failure, Column 5 had an axial deformation of 32 mm.
Similarly, during the preload phase of the fire test, Column 6 had a 6 mm axial deformation
as a result of the applied load of 1717 kN. After the preload phase, the applied axial load
(1717 kN) was kept constant for 4 hours of fire exposure. After 1 hour of fire exposure, the
column began to experience a gradual expansion of approximately 2 mm for up to 3 hours
or until the end of the test. The source of expansion in Column 6 may be from the thermal
expansion of the concrete and a lack of spiral reinforcement as contained in Column 5. For
instance, the temperature of the rebar for Column 6 was 150C and the concrete surface
temperature was 200C after 3 hours. The drop in the deformation curve after 4 hours was
due to the increase in applied load to failure. Column 6 experienced a final axial
deformation of 6.8 mm. In general, the initial degradation for the compressive strength of
concrete is experienced between 200 and 250C, while at 300C strength reduction is in the
range of 15-40%. [7]. In addition, the thermal expansion of concrete is reduced when a
compressive strength of 0.45fc or higher is applied to the specimen during heating [8].
Based on the strength of Column 6 concrete, (36 MPa at 28 days, but tested after one year)
applied load and cross-sectional area, a small expansion from the concrete may have
occurred. The temperatures in the steel reinforcement remained low enough to prevent any
material property degradation. Normally, at 350C or higher, the yield strength of
conventional steel is reduced by 66% of its room temperature yield strength [9]. According
to the temperature of the reinforcement in both columns, the ductility did not increase nor
did the yield strength and ultimate tensile strength reduce.
5. CONCLUSIONS
From results of these full-scale fire endurance tests on one circular and one square FRP
wrapped reinforced concrete column, the following conclusions can be drawn:
1.

The insulation system was effective in protecting the FRP wrapped columns such that
they were able to achieve four hour fire endurance ratings according to ULC S101 and
ASTM E119.

2.

The insulation fire protection material was not able to maintain the temperature of the
FRP below its glass transition temperature for the duration of the fire endurance test.
With an average insulation thickness of 44 mm, the temperature of the concrete surface
for Column 5 reached its glass transition temperature of 60C at about 29 minutes into
the fire test, while Column 6 with 40 mm of insulation material had an FRP
temperature of 60C at 33 minutes into the fire exposure. The difference may be
partially attributed to the mechanism of heat transfer through the insulation in a
circular column versus a rectangular column.

3.

The supplementary insulation fire protection material used in this fire endurance test is
an effective fire protection system. However, the shrinkage cracks in the cured
insulation material should be investigated to further improve the system. Although
157

flames were observed emanating from cracks formed in the insulation of both columns,
the overall insulation remained intact for more than 4 hours of exposure to the ULC
S101 or ASTM E119 standard fire.
6. REFERENCES
1.
2.
3.
4.
5.
6.
7.
8.
9.

Neale, K.W. 2000. FRPs for structural rehabilitation: A survey of recent progress.
Progress in Structural Engineering and Materials, 2: 133-138.
Hollaway L.C. 2010. A review of the present and future utilisation of FRP composites
in the civil infrastructure with reference to their important in-service properties.
Construction and Building Materials, 24: 2419-2445.
ULC. 2007. Standard Methods of Fire Endurance Tests of Building Construction and
Materials. CAN/ULC-S101-07, Underwriters Laboratories of Canada, Scarborough,
ON, 70 p.
ASTM, Test Method E119-10b. 2010. Standard Methods of Fire Test of Building
Construction and Materials. American Society for Testing and Materials, West
Conshohocken, PA, USA, 33 p.
Bisby, L.A., Kodur, V.R., and Green, M.F. 2005. Fire endurance of fiber-reinforced
polymer-confined concrete columns. ACI Structural Journal, 102 (6): 883-891.
Chowdhury, E. 2009. Behaviour of fibre reinforced polymer confined reinforced
concrete column under fire condition. PhD thesis, Department of Civil Engineering
Queens University, Kingston, Ontario, Canada.
Georgali, B., and Tsakiridis, P.E. 2005. Microstructure of fire-damaged concrete-a case
study. Cement & Concrete Composites, 27: 255-259.
Thelandersson, S. 1982. On the multiaxial behaviour of concrete exposed to high
temperature, Nuclear Engineer Design, 75: 271-282.
Chijiiwa, R., Tamehiro, H., Funato, K., Yoshida, Y., Horii, Y., and Uemori, R. 1993.
Development and Practical Application of Fire-Resistant Steel for Buildings, Nippon
Steel Technical Report 58, pp. 47-55.

158

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

AN INVESTIGATION INTO THE SUSTAINABILITY OF FRP


REINFORCEMENT BARS
M.C.K Pearson, T. Donchev and M. Limbachiya
Kingston University, London, UK

ABSTRACT
This paper explores the sustainability of FRP internal reinforcement in the construction
industry and comments on the uses currently employed. Sustainability has many focuses in
industry, but principally it is to ensure that the generations of the future are not negatively
affected by the impact of current developments. FRP has already proven itself to be a
contender in construction against steel with its continually lowering costs of manufacture.
The lower carbon emissions, longer life and correspondingly lower life cycle cost for FRP
reinforced RC elements and the relatively environmentally safe waste are between other
advantages of using those materials. This paper also looks at the manufacture of FRP
materials and how they can be used in industry for FRPs to be a material of the future. This
work expresses not only the idea of resources being renewable or reusable, but also if they
are economically effective in order to give a rounded approach to Sustainability.
1. INTRODUCTION
As we are aware, the challenge facing the world is lowering its carbon emissions. This was
first internationally recognized at the Kyoto Protocol of 1997 which states that countries
must lower their GHG (green house gases) emissions or penalties will be incurred.
The construction industry makes 22.3% of carbon emissions according to the European
Commission for Emissions document, with the energy sector still emitting more (38.2%)
and transport at 23.1% (European Commission, 2010). There are options to make this better
for construction even in times of financial difficulties, but this must be done as a personal
responsibility in order for change to occur globally. America is the largest carbon
contributor; emitting 25.2% of all emissions, China is the second with 15.2% then Russia
with 6.7% and other industrial countries such as Japan, India, Germany and the UK not
producing more than 5% (Nation Master 2010). However, this is justified by their size and
populations as well as their contribution to the global economy.
Steel is widely used in the construction industry and this year (2011) it is predicted that
1306 mmt (million metric tons) will be used: a record level (Recycling Portal 2010). Steel
159

has a high carbon footprint, in particular high yield steel which is preferred in construction,
due to more carbon being used to produce a high yield strength and FRPs carbon emissions
are less than that of steel. Therefore, the continued trend of FRP rebar being used in
preference to steel will lower our carbon emissions over time in construction.
The life cycle costing must be equal if not less than that of steel for FRP to seriously make
a difference to industry. A negative aspect of some FRPs production in terms of materials is
the polymer manufacture, which can be made from fossil fuels. However, in BFRP the
polymer is polysiloxane which is vastly abundant and easily synthesized from siloxanes.
Although this is greatly negative for some FRPs, we must not only look at the long-term
benefits of FRPs with respect to carbon emission.
As well as the direct benefits of FRP, there are indirect benefits with construction, such as
constructing FRPs with concrete. The lightweight rebars and a higher tensile capacity
allows for lighter weight structures. This in turn provides for less concrete and less cover is
necessary to protect the rebars from aggressive environments.
2. ASPECTS OF IMPORTANCE FOR SUSTAINABILITY
Sustainability has evolved in meaning over the years. The original definition was coined as
development that meets the needs of the present without compromising the ability of
future generations to meet their own needs, stated in the Bruntland report in 1987 (World
Commission on Environment and Development (WCED, 1987). In essence, sustainability
must protect the present and the future; therefore the following aspects have been identified
as important in this paper:
2.1 Economy and Life Cycle Costing
Construction accounts for 30.6% of GDP worldwide (CIA, 2010) and therefore has a
massive influence on the world economy. FRP continues to get cheaper and with greater
developments being continually made to FRPs to make them better in construction. That
said it is still more expensive than Steel (Burgoyne and Balafas, 2007). With that in mind,
can FRP be economically more viable that steel? We should not disregard FRPs because
the cost of manufacture is greater than that of steel, as the life cycle costing and other
hidden costs, (such as for repairs) are not included. FRP also gives the opportunity to
design for specialized buildings with longer life spans, and many research papers have
concluded that the life cycle costing for FRP is more beneficial for FRPs as seen in
Nishizakis research. Nishizaki concluded that although the initial costs are high, when
longer life spans are necessary FRP is an efficient choice of material (Nishizaki et al.,
2006).
2.2 Properties
The very properties of FRPs make them sustainable. Polysiloxane, the polymer commonly
used for FRPs, is an inorganic polymer which is already oxidized due to the Si-O bond and
consequently does not suffer from oxidation effects, unlike steel. This property in turn
160

allows for a much greater service life and lower maintenance with fewer necessary repairs.
Moreover, Polysiloxanes have high resistance to UV light and temperature (Mowrer, 2003).
CFRP, GFRP and BFRP all have the same if not greater tensile capacity in comparison to
steel. They are also lightweight (lighter than steel) and have specific stiffness; these
properties add to the repertoire for FRPs being good to construct with. The recent studies
into chemical testing with BFRP have shown great resistance to acid and alkali resistance.
This inertness is a great property to ensure longer life spans of structures. The high
resistance to temperature has also made BFRP a choice in structures for enchanting the fire
resistance of structures as commented by Palmieri et al. (2009).
2.3 Construction Phases
There are definitive phases that a construction material must go through in its life. These
are: manufacture, construction, mid-life and life expansion and end of life as commented by
Burgan and Sansom (2006); these are discussed in relation to FRP usage.
2.3.1 Manufacture
FRP rebar is undoubtedly more difficult to manufacture than steel at present. However, the
carbon emissions and the impact to the environment are less with FRP. Steel contributes
greatly to carbon emissions and temperatures of a little under 1400C are required
constantly, therefore large amounts of energy are used. In order to manufacture Glass,
Carbon and Basalt fibres, temperatures of up to 2400C are used (Lee and Jain, 2009),
which is greater than steel, but currently less FRP is being made and the process of
manufacture is continually being refined. Both FRP and steel rely on fossil fuels to be burnt
(Gerdeen et al., 2006), which are not in infinite supply and this is of concern as Lee and
Jain (2009) point out in their paper outlining the role of FRP. The abundance of Siloxane
(which is any compound containing R2SiO) and Basalt (volcanic rock) for BFRP, make
BFRP a particularly sustainable choice of FRP for the future as well as the products
necessary for CFRP and GFRP.
FRPs contain a greater amount of chemicals in order for the composition to be made as
seen in the Table 1.
Some of these chemicals are rare (Ti for example) and can be hazardous, but for BFRP and
GFRP they are naturally found in the production of the compounds and therefore not added.
Also when controlled they are no greater a threat to the environment than the by-products
of steel (carbon as for CO2 a GHG and unwanted products from the slag).
As for the abundance of materials glass fibres (as SiO2) for FRPs (in particular GFRP) are
in great supply and is easily manufactured using thermosetting plastics such as unsaturated
polyester. Carbon for CFRP too is sustainable to reuse in construction and is greatly
available from graphite. Basalt Fibre is made by the direct extraction of basalt rock
(volcanic) from selected quarries. The rock is then crushed into an extremely fine dust. This
basalt dust is then attached to the polysiloxane polymer, thus reinforcing the polymer. With
161

this argument of materials being in abundance the same can be argued for steel as iron is
the 4th most abundant element of the planet, although the manufacture of steel could be
seen to be more harmful to the environment, as the materials required for FRPs are more
readily available than smelting steel from the extraction of iron ore.
Material Fe
Mild

steel

Table 1. Elemental composition of materials


O
C
Si
Mg Al
Ca
Na K

High
Yield
Steel

GFRP

BFRP

Ti

Br

Cl

2.3.2 Construction
The single greatest disadvantage for FRP has been the lack of design codes. Burgoyne and
Balafas stated in 2007 that this was a main contributor to why FRP is not a financial
success. Uomoto (2007) also agreed that efforts needed to be made for design codes for
FRP and this would mean greater ease and use of FRP in buildings and not just as an
alternative material. Indeed great efforts have now been made and there are now design
codes in many countries such as America (although a Eurocode is now been developed) and
this is due to the continual interest in FRP as seen in all research and specifically the
document ACI 440.1R-06 (Ospina and Gross, 2006).
The lack of sufficiently developed design codes has not stopped engineers building with
FRPs, however, and indeed the properties of FRPs compliment the building industry well.
Their lighter weight makes the construction process safer, boosting the construction
industrys poor health and safety record. This also boosts the construction time of buildings
as seen by (Lee and Jain, 2009). The UK has seen this directly with domes and mosques
been made in cheaper and faster fashion (Kendall, 2007).
2.4 Midlife and life expansion
Life expansion is easy with the rebars as other external FRP can be retrofitted which
mirrors the structural behavior of the internal rebars. The properties of FRPs compliment
strengthening to buildings in need of repair and this has been a market that has greatly
boosted the use of FRP. It is always of benefit to extend the service life of a structure as far
fewer resources are required than to demolish and re-build. This compliments the ideology
of sustainability with reuse of buildings. FRPs also do not require methods like cathode
protection due to their inertness to corrosion and would only require standard monitoring
techniques with sensors and monitoring the deflection of the beam in order to ensure that
the beam is not failing.
162

2.5 End of Life and Waste


With rebars there is very little waste as the design is established well before the build so the
amount used is the amount ordered (waste minimalisation). That said, the waste of FRP
becomes an issue in demolition. Steel offers benefits that FRP in general does not; steel is
readily reusable and recyclable and there are large businesses that specialize in this. Steel is
easily processed from the waste concrete and then recycled, but the reprocessing of steel
also produces CO2 emissions. FRP can also be reused in other aspects such as crushing it
down into aggregate (dispersed reinforcement). Patents have now been passed for the
recycling of waste FRP, as seen by (Kamite et al., 2008) and this is a major step of many to
come for the recyclability of FRPs. There are other methods of using chemical substances
to break down the resin from the polymers, however this is an expensive and hazardous
solution, but with waste of plastics being a growing problem there are many initiatives in to
process FRPs.
The main issue with recycling and therefore reusing FRPs is the way in which they are
manufactured as identified by Bartholomew (2004). FRPs are highly engineered plastics
designed for long lifespan. Therefore they demand high amounts of energy (as they have
high calorific values) making incineration with energy recovery troublesome to the
operators as well (Halliwell, 2009).
FRP waste is in general non-hazardous, however only the dust particles (which are only
produced in cuttings during the manufacture stage) have been studied in laboratory
conditions and have been seen to be harmful, although much less in comparison to other
construction materials (Hesterberg et al. 1998; U.S. Department of Health and Human
Services, Public Health Services, Agency for Toxic Substances and Disease Registry,
2004).
FRP waste according to the European waste list 2000/532/EC, is classified under several
codes (CEFIC Technical Bulletin, 2006) and these must be known in order to organize
waste correctly and legally in Europe. Possibly the most commonly used for construction
purposes at the demolition stage would be CEFIC - 17 02 03. However, this would depend
on the state of the FRP; as stated before it can be classified into Hazardous waste within
the H1 to H14 list of Annex III of EC Hazardous Waste Directive (CEFIC Technical
Bulletin, 2006).
Halliwells paper on Recycling FRP materials, comments on FRPs being difficult to reuse,
as it is difficult to ascertain the loading properties of pre-used FRP beams, and this can be a
lengthy process. FRP waste is also predicted to double, from some 156000t being used in
2000 to a predicted 304000t in 2015 (Harvers, 2002).

163

3. CONCLUSIONS
FRP is a material that has the capacity to allow for greater service life and greater life cycle
costing for construction of structures. FRPs are cost effective in many aspects, have higher
strengths, greater resistance to aggressive environments and are sustainable in our current
global environment. Certainly FRPs have their disadvantages in comparison to steel; they
are not readily reusable and there is no well established design code in the EU. The
finalization of the design code worldwide and standardization of FRP materials will ensure
the future of FRPs. In terms of manufacture, FRP offers great sustainability and great
promise to the world of construction.
4. REFERENCES
1.
2.
3.
4.
5.
6.
7.
8.
9.

10.
11.
12.
13.
14.

Agency for Toxic Substances and Disease Registry (ATSDR). 2004. Toxicological
Profile for Synthetic Vitreous Fibers. Atlanta, GA: U.S. Department of Health and
Human Services, Public Health Service.
Bartholomew, K. 2004. Fibreglass Reinforced Plastics Recycling. Minnesota
Technical Assistance Program.
Burgan, B.A. and Sansom, M.R. 2006. Sustainable Steel Construction. Journal of
Constructional Steel Research, 62(11): 1178-1183.
Burgoyne, C. and Balafas, I. 2007. Why is FRP not a Financial Success? Proceedings
of the 8th International Symposium on Fiber-Reinforced (FRP) Polymer
Reinforcement for Concrete Structures (FRPRCS-8), Patras, Greece, 16-18 July, 10 p.
European Commission. 2010. EU Energy in Figures 2010. Directorate-General for
Energy and Transport (DG TREN).
Gerdeen, J.C.U., Lord, H.W. and Rorrer, R.A.L. 2006. Engineering Design with
Polymers and Composites CRC Press, Taylor and Francis Group.
Halliwell, S. 2009. Recycling of FRP Materials in Construction. ICE Manual of
Construction Materials, pp. 695 706.
Harvers, F. 2002. Green Recycling Label. The Step Towards Closing The
Loop, Proceedings of the 3rd Automotive Seminar, Cambrai/ Lieu Saint Amand,
France.
Hesterberg, T.W., Chase, G., Axten, C., Miller, W.C., Musselman, R.P., Kamstrup,
O., Hadley, J., Morscheidt, C., Bernstein, D.M., and Thevenaz, P. 1998.
Biopersistence of Synthetic Vitreous Fibers and Amosite Asbestos in the Rat Lung
Following Inhalation. Toxicology and Applied Pharmacology, 151(2): 262-275.
Kamite, M. and Kato, M. 2008. Process for Recycling Waste FRP. United States
Patent 7407688.
Kendall, D. 2007. Building the Future with FRP Composites. Reinforced Plastics,
51(5): 26-33.
Lee, L.S. and Jain, R. 2009. The role of FRP composites in a sustainable world. Clean
Technologies and Environmental Policy, 11(3): 247-249.
Mowrer, N.R. 2003. Polysiloxanes. in Ameron International - Performance
Coatings & Finishes.
Nation Master. 2010. Environment Statistics - CO2 Emissions (most recent) by
country.
(http://www.nationmaster.com/red/pie/env_co2_emi-environment-co2164

emissions2010), May.
15. Nishizaki, I., Takeda, N., Ishizuka, Y. and Shimomura, T. 2006. A Case Study of Life
Cycle Cost based on a Real FRP Bridge. Proceedings of the Third International
Conference on FRP Composites in Civil Engineering (CICE2006), eds. Mirmiran, A.
& Nanni, A., Miami, Florida, USA, 13-15 December, pp. 99-102.
16. Ospina, C.E. and Gross, S.P. 2005. Rationale for the ACI 440.1R-06 Indirect
Deflection Control Design Provisions. Proceedings of the 7th International
Symposium on Fiber-Reinforced (FRP) Polymer Reinforcement for Concrete
Structures (FRPRCS-7), eds: Shield, C.K., Walkup, S.L, Busel, J,P. & Gremel, D.D.
Kansas City, USA, pp. 651-670.
17. Palmieri, A., Matthys, S. and Tierens, M. 2009. Basalt Fibres: Mechanical Properties
and Applications for Concrete Structures. Proceedings of the International conference
on Concrete Solutions, Padua, Italy, CRC Press, Balkema, pp. 165-169.
18. Uomoto, T. 2007. Test Methods for FRP Materials. Proceedings of the 8th
International Symposium on Fiber-Reinforced (FRP) Polymer Reinforcement for
Concrete Structures (FRPRCS-8), Patras, Greece, 16-18 July, 9 p.
19. UP resin group of CEFIC. 2006. Classification and Handling of FRP Waste within the
Current EC Legislation.
20. World Commission on Environment and Development (Brundtland Commission).
1987. Our Common Future, 43 p.
21. Zurck. 2011. Worldsteel: Steel Use Will Increase by 10.7 Percent.
(http://www.recyclingportal.eu/artikel/24078.shtml), May.

165

166

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

METHODS FOR EVALUATING LONG-TERM PERFORMANCE OF


FRP-STRUCTURAL SYSTEMS APPLIED TO CONCRETE BRIDGES
K. Crawford
Lead Project Engineer, Directorate of Civil Engineering, 95th Air Base Wing, Edwards AFB, CA,
USA

ABSTRACT
A challenge confronting transportation infrastructure owners using carbon fiber-reinforcedpolymer (CFRP)-structural systems to repair and strengthen reinforced-concrete (RC)
bridges is having necessary protocols to determine and predict the long-term performance
of the CFRP-concrete bridge systems. Adequate procedures to measure CFRP-RC bridge
performance over time and models to predict performance are not currently available to
owners. While a range of environmental and mechanical deterioration factors potentially
impact the CFRP-concrete (CFC) bonding system affecting the long-term durability and
performance of the bridge structural system, the sustained performance of CFRPstrengthened bridges is a function of three factors: (1) design of the CFRP-concrete
structural system, (2) the CFRP material application process to the concrete, and (3)
maintenance of the CFRP-bridge and structural components. Whereas the CFRP design
and application determine the baseline conditions for bridge performance upon construction
completion, the maintenance process sustains bridge performance over time. This paper
considers methods to evaluate the performance of CRP-strengthened concrete bridges by
using non-destructive testing in the CFC bridge maintenance process to examine the
condition (strength) of the CFRP-concrete bonding structure and to monitor changes in the
bonding system. Maintenance procedures that control and mitigate deterioration factors
impacting CFRP-concrete bonding will contribute to long-term performance of the CFCbridge structural system. The 19 CFRP-strengthened bridges on European Corridor 8 in the
Republic of Macedonia are used as a model for the evaluation.
1. INTRODUCTION
The purpose of this paper is to present methods to test the performance characteristics of
large areas of CFRP composite material applied to a number of concrete bridges by
looking at the condition of the bond between the CFRP plates and the concrete structural
member. Because load performance of CFRP-strengthened (CFS) RC bridges is dependent
on the condition and integrity of the CFC bond the objective of this approach is to identify
areas of CFRP material de-bonding from concrete structural members, and thus be able to
167

determine the degree of degradation in structural performance. Ideally the applied CFRP
plates have 100% bonding to the concrete structural members and the designed level of
performance is maintained. However, over time under influence of a range of
environmental and mechanical deterioration factors the bond between the CFRP material
and the concrete will degrade. Therefore to maintain bridge performance de-bonded areas
must be identified and repaired. To examine the integrity of the CFRP-concrete bonding
structure this paper describes a non-destructive testing (NDT) method for testing the bond
structure using a mobile impact-echo (hammer) device. This device generates low
frequency acoustic pulses to detect areas of CFRP plate de-bonded from the concrete
structure through changes in frequency. The NDT concept presented in this paper is
designed to test large areas of CFRP bonded plate on multiple bridges. The device is
currently under development and will undergo field testing in late 2011, in the Republic of
Macedonia, on the highway bridges strengthened in 2001 with CFRP-structural systems.
In RC bridge maintenance programs, a desired objective of bridge owners is to maximize
bridge load performance for its designed service life with a reasonable investment in
maintenance and rehabilitation funds. However, for owners of CFRP-strengthened bridges
a fundamental issue in bridge maintenance protocols concerns how to measure, maintain,
and predict the performance of the CFRP-strengthened bridges at a reasonable cost. Nondestructive testing of the CFC bond with an impact-echo device is an effective method for
testing the bridge CFC bonding structure and evaluate bridge load performance. For CFS
bridge maintenance programs non-destructive testing should meet the following criteria:
(1) NDT testing is easy to apply; (2) the NDT device and testing methods are inexpensive;
(3) the NDT procedure is capable of testing large areas of CFRP plates in reasonable time;
and (4) NDT testing provides sufficient data to quantify the structural condition of the
bond. Testing the condition of the CFC bonding structure is a necessary element in a
comprehensive CFS bridge maintenance program to maintain bridge load performance over
time.
2. BACKGROUND
In 1999, in support of the NATO peacekeeping operations in Kosovo, it became necessary
to move heavy military equipment by road from Bulgaria through the Republic of
Macedonia into the theater of operations, as reported in earlier papers [1,2]. Certain
bridges on the movement route were overloaded above designed safety limits and required
strengthening. The decision was made to use a CFRP-structural system to strengthen 19 of
48 slab and girder bridges (over 90 km) on the highways M2 and M1 (E-870) in
northeastern Macedonia to allow the efficient movement of military heavy equipment
transports (HETS). The 17 bridges on the M2 were strengthened in 2001 and the two M1
bridges strengthened in 2002. A total of 14,600 linear meters of CFRP plates in widths of 5
to 15 cm were applied to bridge girders and decks, a total area of 1,318 m2. 19 bridges,
originally designed with the 1949 M-25 loading scheme, were strengthened with CFRP
using a V600 load scheme, increasing the bending moment at main girder 10-meter midpoint from 75 ton-meter(tm) to 160 tm(bridge B7) and from 50 tm to 110 tm at the main
girder 7-meter midpoint (bridge B18). The application of the FRP structural system
produced a significant increase in the structural member bending moment and bridge
loading capacity. As a consequence the CFC bonding structure becomes a critical element
168

in bridge load performance and must be tested and maintained to sustain the V600 designed
loading. The challenge for the owner becomes how to test and evaluate 14,000 meters of
bonded CFRP plates and how to determine the extent of de-bonding, if any, and its impact on
bridge performance.
Figure 1 shows examples of the numbers of CFRP plates applied to three (of 19) bridges,
with their bonded areas shown in Table I. It has been ten years since the 17 bridges on the
M2 and two bridges on the M1 were strengthened. Testing of CFRP bond on the bridges has
not been performed.
The issue is this: How are the strengthened bridges currently
performing, i.e. is the designed V600 load scheme still sustained by the bonded CFRP plates
as originally applied to the bridges in 2001? Table I shows the CFRP plate area for 10 of
the 19 CFS bridges. The issue for bridge engineers becomes how to evaluate the entire CFC
area on 19 bridges in a timely and cost effective manner. The NDT impact-echo method
proposed in this paper provides an effective solution to evaluate the large bond area on the 19
CFS bridges.

Fig. 1. M1- Bridge B2-N


96 Plates (12cm)

M2 Bridge B36
80 Plates (15 cm)

M2 - Bridge B22
60 Plates (15 cm)

4. FACTORS CONTROLLING CFS BRIDGE PERFORMANCE


Three factors define performance of RC bridges strengthened with CFRP-structural
systems:
(1) Design of CFRP-concrete structural strengthening system
(2) Application of the CFRP material to the concrete structural members
(3) Maintenance of the CFRP-bridge structural components and system
4.1 CFRP Design
When the M2 project was designed in 2000, national codes and ACI-440(in draft) were used.
Per 440 [3] the design for CFRP structural systems required the flexural strength Mn of
the concrete member to exceed the required loading factorial moment Mu such that:
Mn Mu where Mu = [(d a/2) fy As ] i Mi

(1)

where d=depth of section, a=depth of concrete, fy=steel yield strength, and As=cross
section area of flexural steel. is the strength reduction factor and i Mi is the sum of
the factored loading moments on the section. The FRP strengthening limits of the concrete
169

flexural members were controlled by the failure mode and the ductility and serviceability
limits of the structural system.
Table 1. Area of applied CFRP plates for selected bridges
Typical CFRP-Strengthened Bridges
CFRP Plate
(10 of 19 CFS bridges)
Length (m) and Area (m2)
Bridge
No.

Location
On M2

Bridge
Type

Length
(m)

B7
B11
B18
B22
B28
B35
B36
B37
B39
B2-N

14+027
21+876
38+444
41+786
49+631
66+058
67+409
68+452
71+211
M1-Kum

Girder
Slab
Slab
Slab
Girder
Girder
Slab
Girder
Girder
Slab

120
10
36
30
50
52
21
17
85
46

No.
Spa
n
6
1
1
1
3
3
2
1
4
4

Plate
Length

Number
Plates

Plate
Area

1478
198
218
1308
1210
346
1032
415
1778
938

72
26
54
60
60
20
80
36
48
96

116.1
29.6
32.6
196.3
100.6
41.2
82.8
34.4
146.2
112.5

Total CFRP Plate Area for 19 bridges: 1,318 m2


While a strength reduction factor of 0.80 is recommended by ACI 440, because of
uncertainty in the condition of the reinforcing steel a reduction factor of 0.70 was used.
This strength reduction factor improved the reliability of the CFRP-strengthening design
for the M2 bridges.
4.2 CFRP Application
The FRP structural system application to RC bridge members is a process to apply an FRP
composite material with an epoxy compound to a concrete substrate, to form a high-quality
tri-layer FRP-epoxy-concrete structural system. The goal is to achieve 100% bonding of
the FRP material to the concrete substrate. The quality of the bond is critical for achieving
the designed load performance. These conditions are normally specified by the FRP
material manufacturer. Training and certification by the FRP manufacturer of the
contractor, and inspection personnel. The actual application process will require substrate
repair and surface preparation, correction of corrosion related deterioration, and repair and
injection of cracks, normally greater than 0.3 mm. The concrete surface must be prepared
to achieve a specified surface profile. The mixing and application of the bi-component resin
in the correct temperature range is necessary to achieve a long-lasting bond. The CFRP
application process requires controlling key parameters, such as temperature, humidity, and
moisture in the substrate. A good quality control and quality assurance system by the
contractor and the inspectors is necessary to ensure a lasting FRP-concrete bond. In
practice, with no de-bonding factors affecting the bonding structure, the CFRP-epoxyconcrete tri-layer system should last indefinitely and sustain its original designed load
performance. In reality some de-bonding will occur
170

4.3 CFRP-Bridge Maintenance


The third factor necessary for sustaining CFC bridge performance is maintenance of the
bridge, its structural components, and the applied CFRP strengthening system, to include
the composite material and the bonding epoxy. Good maintenance procedures by the owner
are necessary to sustain optimum designed performance of the FRP-strengthened bridges.
The FRP bridge maintenance process consists of three key components: inspection of the
bridge and its FRP system, testing of the FRP-epoxy-concrete bonding structure, and
maintenance of the FRP composite and epoxy bonding material and the bridge concrete and
steel structural components.
NDT evaluation with the described impact-echo device is key to obtaining bond condition
data.
The maintenance program establishes inspection criteria and procedures, sets maintenance
standards and procedures to maintain bridge performance, and sets conditions to maximize
the service life of the bridge structures. Control and mitigation of deteriorating factors and
de-bonding mechanisms is a key component in a bridge maintenance program.
5. METHODS TO EVALUATE CFRP-CONCRETE BOND CONDTION
Of the ten NDT methods described in ACI 228 the best method for detecting flaws in
bonded overlays on concrete slab structures is the impact-echo method [4] .
5.1 Principles of Impact-Echo NDT
The principle of the impact-echo technique generates a transient pulse into the CFRPepoxy-concrete tri-layered structure such that a wave is reflected by the boundaries of the
structure interfaces. The pulse produces P and S waves that propagate into the layered
structure in hemispherical waveforms. The objective of using this NDT method on the
CFRP plate is to accurately identify areas of de-bonding between the plate and concrete.
Using the acoustic signal generated by an impact device applied to a layered bonding
structure provides an effective means to detect and identify areas of voids and delaminated
materials in the bonding system.
5.2 NDT Device and Concept of Operation

Fig. 2. (a) Impact-echo device

(b) Plate impact-signal pattern

171

Figure 2 shows (a) the mobile impact-echo device developed to test CFRP plate bond and
(b) the resulting impact-signal pattern. The four-wheel mechanical impact device consists
of two 15cm cam-actuated levers with hammer pins, aligned 5 cm apart on axes A and B.
Each pin is driven by an adjustable compression spring to obtain an equal impact force on
the CFRP plate to generate optimum acoustic signal intensity. The impact of the alternating
striking pins generates acoustic waves penetrating the CFRP-epoxy-concrete tri-layer
bonding which in turn reflects an acoustic wave to a signal transducer (receiver). The pins
are designed to strike the CFRP plates at four cm intervals alternately on the A and B axis.
A 15cm wide CFRP plate will have 25 impacts per meter, with one impact producing a
bond condition response for 600 mm2 of CFRP plate area. The impact device is designed to
strike the plate with equal force over the length of the CFRP plate generating an acoustic
signal with consistent intensity for a fully bonded plate. If the CFRP plate is bonded for the
length of the concrete structural member the signal frequency will be consistent, i.e.
without variation over the entire length of the plate.
The frequency of the generated pulse is a function of the impact duration. A shorter
duration, < 80 s, generates a higher frequency and is suitable for detecting flaws in
concrete structural members less than one meter in thickness. For the M1 and M2 bridges
the desired depth of flaw detection, from the CFRP plate surface to the closest steel
reinforcing in the concrete is less than 30 cm. The depth of the CFRP plate surface to the
epoxy-concrete interface is less than 1.5 cm and is the focus of interest in the impact-echo
NDT method proposed. Each impact signal received is logged on a data recorder. A
bonded plate will generate a specific frequency, and a de-bonded plate will generate a
lower frequency. The variation in received acoustic frequency will indicate a change in the
CFC bond structure, i.e. potential de-bonding. As testing of the plate bond structure
proceeds and different frequencies are recorded, frequency analysis of the displacement
waveform (reflected pulse) is used to obtain an amplitude spectrum from the time-domain
signal. The frequency of the reflected pulse is a function of wave speed and distance
between the plate test surface and the internal reflecting surface, i.e. a function of the CFC
bonding/de-bonding structure. De-bonded CFRP plate areas are are marked for repair.
6. CFC BRIDGE DETERIORATING FACTORS
There are a number of failure modes and de-bonding mechanisms, e.g. moisture, affecting
bonded CFRP plate. In work by Kotnia[5], two primary modes of bond failure occur in the
plate anchorage areas and the maximum bending moment in the mid-beam region, a focus
for this NDT method. In one theoretical de-bonding load model by Nigro and Savoia [6],
CFRP plate width, length, thickness, and axial stiffness define the de-bonding load for a
structural member. Thus a decrease in plate bonded width and length proportionally
decreases CFS bridge load capacity. Control and mitigation of CFRP plate de-bonding
mechanisms, especially moisture, in a CFC bridge maintenance program will contribute to
extended service life for the bridges.
7. CONCLUSION
This paper is written from the perspective of a transportation infrastructure owner
confronting the issue on how to inspect, test, and maintain a large number of RC bridges
172

strengthened CFRP systems. Methods to efficiently test the bond condition of large areas
of CFRP plates applied to RC bridges are not readily available. The proposed impact-echo
NDT method will identify de-bonded CFC areas on concrete structural members will
provide bridge engineers qualitative data on CFRP-concrete bond condition and the
necessary input for bridge maintenance protocols.

Mobile impact-echo device is a means to rapidly test and evaluate large areas (long
lengths) of CFRP plates on multiple bridges.
Recording the frequency data for the bond establishes a signature for CFRP plate
bond condition at a point in time and becomes a historical record for the bridge.

CFC bond field measurements on the 19 E-870 bridges with the impact-echo device
described in this paper are not yet available. Field testing on a number of the bridges on the
E-870 is scheduled in the autumn of 2011.
8. REFERENCES
1.

2.

3.
4.
5.

6.

Crawford, K. 2007. A Model for the Inspection and Maintenance of FRP-Bridge


Strengthening Systems. Proceedings of the Third International Conference on
Durability and Field Applications of Fibre Reinforced Polymer (FRP) Composites for
Construction (CDCC2007), eds: Benmokrane, B. & El-Salakawy, E., Quebec City,
Quebec, Canada, 22-24 May, pp. 269-276.
Crawford, K., Nikolovski, T. 2006. The Design and Application of CFRP Composite
Material to Strengthen 19 Concrete Highway Bridges in the FY Republic of
Macedonia. Proceedings of the 2nd International fib Congress, Naples, Italy, 5-8 June,
12 p.
ACI Committee 440. 2002. Guide for the Design and Construction of Externally
Bonded FRP Systems for Strengthening Concrete Structures. ACI 440.2R-02,
American Concrete Institute, Michigan, USA, 45 p.
ACI Committee 440. 1998 (Reapproved 2004). Nondestructive Test Methods for
Evaluation of Concrete Structures. ACI 228.2R, American Concrete Institute,
Michigan, USA, 62 p.
Kotynia, R. 2005. De-bonding Failures of RC Beams Strengthened with Externally
Bonded Stripes. Proceedings of International Symposium on Bond behavior of FRP in
Structures (BBFS 2005), eds: Chen and Teng, Hong Kong, China, 7-9 December, pp.
247-252.
Ferracuti B., Martinelli E., Nigro E., Savoia M. 2007. Fracture Energy and Design
Rules against FRP-Concrete Debonding. Proceedings of the 8th International
Symposium on Fiber Reinforced Polymer Reinforcement for Reinforced Concrete
Structures (FRPRCS-8), Patras, Greece, 16-18 July, 10 p.

173

174

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

CREEP BEHAVIOUR OF GFRP CABLES, SHORT AND LONG TERM


TEST
E. Ferrier, A. Si-Larbi and P. Hamelin
Laboratoire de gnie civil et d'ingnierie environnementale, universit claude bernard, lyon i, 82 bd
niels bohr, domaine scientifique de la doua, 69622 villeurbanne cedex, France.

ABSTRACT
This paper deal with the relaxation of GFRP cable used for linear RC unbonded posttensioning method. In 1990, GFRP cables have been used on one span of a prestress
concrete bridge of French Alps Highway. This bridge has been built with a usual steel
cables post-tensionning method. Some GFRP cables have been add for an experimental
purpose. The level of prestress applied on cables was 60 % of the ultimate cable strength.
Some cables have been kept in the same environmental conditions without loadings. After
this period in 2006, on site test are done in order to evaluate the loss in properties of the
GFRP cable, but also the stress cables loss. The result of this experiment allows obtaining
the durability of the cable and the effect of permanent loading on GFRP. Relaxation
functions are obtained in laboratory using time-temperature principle with short time period
relaxation test. Time temperature principles have been applied in order to estimate the loss
in stress of GFRP cables and compare then with the result obtained directly on site after
more than 15 years of relaxation. The paper present the methodology to evaluate the
relaxation of the cable with laboratories testing and compare successfully the relaxation
function to the long term relaxation test obtained on site.
1.

INTRODUCTION

1.1 Bridge Description


The bridge is a 373 m continuous span box girder located in Bardonnex (74, France). The
bridge was made with prestress concrete using bonded conduits post-tensionned method.
Structural reliability is insuring with this steel prestress method but some GFRP cables
have been had for experimental purpose. The use of a composite reinforcing bar for
concrete as an alternative to traditional steel reinforcing bar has many potential advantages.
The material is relatively light and corrosion resistant. However it is important to obtain
representative data on the long-term durability of concrete reinforced with composite rebar
in order to be able to asses the performance of the structures under non-standard conditions
which have to be considered in the design and regulation process.

175

In 1990, an experimental investigation of the reliability of prestress GFPR cables has been
done on this bridge. Few partners (GTM construction, ATMB, LCPC) have decided to
investigate the performance of such prestressing technique. In this purpose GFRP rods have
been prestress using post tensioning method. 5 GFRP bars have been prestress in two
unbonded conduit on each side of the box girder (Fig. 1). At each end of tendon specific
anchorage system have been developed made of steel wire strands placed around GFRP
tendons.

Fig. 1. Cross section on Bridge box girder

(a) Picture of the bridge


(b) Bridge box girder
Fig. 2. Bridge of Bardonnex (74 France)
1.2 Experimental Investigation
In order to investigate the long term behaviour of the GFRP cables and the residual stress in
the tendon after 16 years of loading, several tests are done. The cable stress loss is
essentially due to:
-

An initial anchorage or frictional loss during the prestressed process located to


the tendon anchorage.
Aging of GFRP and Young modulus diminution
Time dependent losses due to relaxation of the cable depending on the
rheological behaviour of GFRP tendon.
176

The total tendon stress loss can be sum-up by relation 1.


f GFRP (t , t 0 ) f GFRP , E (t , D ) f GFRP , R (t , t 0 ) f GFRP , A (t )

(1)

with f GFRP , E (t , D ) : loss due to aging of GFRP (E)


and f GFRP , R (t , t 0 ) : loss due to relaxation of GFRP (R)
and f GFRP , A (t ) : anchorage loss (A)
The loss in stress is measured directly on the bridge tendon thanks to a specific device. The
residual stresses are obtained and compare with the initial one. The next stage is to evaluate
the evolution of mechanical properties of GFRP (tensile strength and Young modulus).
Parts of the free tendon (Fig. 3) are taken on site and tested in lab. This value are compare
to the one obtain 16 years ago on the tendon. In order to study the relaxation behaviour of
the cable, some samples of GFRP tendon are also test using time-temperature principle in
order to obtain the material relaxation function. It is then possible to explain the loss
measured on site thanks to the different laboratory testing.

Fig. 3. Picture of GFRP tendon end anchor


2.

STRESS LOSS OF GFRP CABLES

2.1 Measurement of Stress Loss On-Site


The aim of this part is to evaluate the relaxation of GFRP tendons after 16 years of loading
on the bridge.
In order to obtain this value, a specific device was developed (Fig. 4); the device was made
of a bearing plate placed over the dead end anchorage. A threading steel tube allows to
apply a force on a anchoring cone. The wedge transfers this force to the tendon. It is then
possible thanks to a load cell to measure the load which is necessary to apply to obtain an
additional strain to the tendons. On each GFRP bars, in the mid part, strain gauges were
bonded.
During the loading process, strain gauges and the load cell were connected to a data
acquisition system. Figure 6 illustrate the value of load and strain in function of time. It is

177

clear that when the wedge became free (Fig. 5), the composite strain increases suddenly,
this allows evaluating the value of the residual prestress force (Fig. 6).
Wedge
32

Load cell
20
32

Threading steel tube

40

Bearing plate

32

53

32

20

160
20

20

32

20

32
160

Fig. 4. On site tensile force measurement device

Fig. 5. Photos of wedge before and under maximal loading value

-1000

20

Load (daN)

-1500
-2000

15

-2500
-3000

10

-3500
-4000

GFRP Strain (mm/m)

-500

25

90
12
0
15
0
18
0
21
0
24
0
27
0
30
0
33
0
36
0
39
0
42
0
45
0
48
0
90
0
93
0
96
0
10
00

0
30
60

-4500
-5000

0
Load cell
GFRP strain

Time (s)

Fig. 6. Load and composite strain in function of time


For each cable the value obtains is given by table 1.
Table 4. On site GFRP cable residual prestress force
Cable 1
Cable 2
Cable 3
Cable 4
Force (kN)

40

44

41
178

43

Average
42

An average value of 42 kN is then measure. These values correspond to a 31 % force loss


compare to the initial 61 kN.
2.2 Evolution of Tensile Mechanical Properties
Tensile test are done with a displacement control rate of 1 mm/min. Initial 17 years ago
tensile strength of GFRP was 1330 MPa 150 MPa for a Young modulus of
45900 MPa 2500 MPa where as these value has respectively decrease to
800 MPa 180 MPa and 36500 MPa 2000 MPa. It is then obvious to conclude on a
decrase of 40 % of tensile strength and a 20 % Young modulus decrease (Fig. 7). The loss
in force in GFRP tendon due to the loss in Young modulus properties is given by relation 2.

t 0 , t k f , D E f ,0 (t 0 )

(2)

f GFRP, E t 0 , t A f k f , D E f ,0 (t 0 )
With consideration of a linear damaging process of GFRP
k f ,D 1 a t

(3)

With a: aging factor (0.0117); t : aging time in years; Ef,0 : initial Young modulus; (t0) :
initial prestress strain; Af : composite area
1600
1400

Strength (MPa)

1200
1000
800
600
400

Tensile test after 17 years


Initial tensile test

200
0
0

5000

10000

15000

20000

25000

30000

35000

Strain (m/m)

Fig. 7. Average curve of Tensile test


This result are comparable to the one obtained by Zermeno and all [1] and Abbasi and all
[2].
3.

RELAXATION FUNCTION: SHORT TERM TEST

3.1 Time-Temperature Principle


The construction of the master curve can be done step by step by using the time
temperature superposition WLF principle (William Landel and Ferry [3]). The time
superposition of the short-term creep compliance test data provided the means for the
sequenced, short term data to be collapsed into a single master curve (MC). This principle
is used for long-term creep and fatigue polymer study. It allows a time test reduction.
Brinson and Gates [4] have extended the use of the time-temperature superposition
179

principal to composite. It allows predicting the long-term creep compliance or relaxation


using as input the material properties developed from short-term tests.
It means that a test made with several temperatures on a short period allows assessing the
long-term behavior for a reference temperature by the master curve construction method
using a shift factor at(T). This shift factor is simply defined as horizontal distance required
to shift a relaxation curve to coincide with a reference relaxation curve. The test procedure
is as follows: the test specimens (three for each test) are placed in a thermo-regulated room
at different temperatures ranging from 20 to 80C. The loss is composite force is measured
during a five-hour period, up to 80C. From each single test, we obtain a short-term
relaxation function which is usually referred to as a momentary curve. By repeating this
process at different temperatures, a family of relaxation functions can be obtained. All
those individual relaxation curves can be superposed onto a single master curve. For each
test, we have made 5 short-term relaxation tests (5 hours) with an initial constant strain of
6000 m/m and an initial tensile force of 12.4 kN. In the application of this approach to
obtain data from a short-term relaxation test, it is important that the data should not bias
against any region of the desired curve. Since the application is to be a curve in log time,
the data should approximate to an equal spacing in log time. This has been accomplished
for this work through adjustment of the data sampling rate during the experimental study.
We repeat that the goal of shifting momentary curves to a single reference (master) curve is
to predict response of the GFRP stress.
3.2 Results
Single curve obtain between 20C to 80 C are plot on figure 8 a. Each curve may be shift
in order to obtain a master curve at 20C.
Load cell

Anchoring cone
Relaxation frame

Loading frame

GFRP tendon

1.00 m

Fig. 8. Experimental relaxation frame


14000

14000

12000

12000
10000

Force (N)

20C

8000

40C
60C
70C
80C

6000

Force (N)

10000

8000
6000

4000

4000

2000

2000

20C
40C
60C
70C
80C
Master curve 20C

0
0

5000

10000

15000

20000
Time (s)

100

10000

1000000

(a) Single force-time curve


(b) Master curve construction
Fig. 9. Relaxation result

180

100000000
Time (s)

This master curve is build thanks to the six relaxation curve measured experimentally. The
reliability of this method may be enhance by plotting the logarithm of the shift factor (at) in
function of T-To, with To reference temperature of the master curve (20C) and T
corresponding shift factor temperature. The linear variation of log(at) in function of T-To
generally correspond to a reliability of time-temperature principle.
3.3 Relaxation Function
For considering viscoelasticity, the three-parameter Zener model (or standard linear model)
shown in Fig. 9 can be applied to describe the viscoelastic relaxation of GFRP tendons.
Based on this model, the differential equation of the constitutive relationship is given as
equation 4 which is the relaxation function of three-parameter Zener model shown in Fig.
10. We may investigate the asymptotic behavior of the relaxation by applying Laplace
transform with the limit theorems to Eq. (4):

E2

( E1 E2 )

( E1 E2 )

(4)

The elastic modulus E1 used in the three parameters model is corresponding to the elastic
modulus E obtained from the Eq. (4), which was derived based on elastic theory. Based on
Eqs. (6), the elastic constant E2 can then be determined from the experimental data and the
results are shown in Figs. 7 and 8. It is notable that these elastic constants decrease with the
increasing temperature. The viscosity parameter can be simply calculated from Eq. (13),
combined with the experimental data, and the results are plotted in Fig. 9. Using the master
curve relaxation function obtained, a Zener rheological model is retained to fit with
relaxation GFRP behaviour.

Fig. 10. Zener model


f GFRP , R A f 0 E (t , t0 )

E E
With E (t , t0 ) 1 2
E E2
1

(5)

( E1 E 2 )

1 e
E1 e ( E1 E 2 )

181

(6)

With

1
1
1

K E1 E2

(7)
t

Then E (t , t0 ) K ( E1 K ) e e
With e

(8)

(9)
E1 E2
Viscosity factor (4.109), e : relaxation time, E2 (88 000 MPa) and E1 (36000 MPa) :
Elastic constant
14000
12000

Force (N)

10000
8000
6000
4000
Master curve 20C
Zener Model

2000
0
1

100

10000

1000000

100000000
Time (s)

Fig. 11. Comparison between experimental and modelling


The model allows obtaining, with a good accuracy, the relaxation function and after 16
years of loading the expected residual load obtain with the modelling is 44.5 kN. Finally
the loss due to relaxation is given by:
t


fGFRP, R (t, t0 ) Af E0 K (E1 K) e e 0

4.

(10)

CONCLUSION

This experimental investigation allows to compare real loss of stress obtain on site and
different laboratory testing. The total loss of the GFRP bars can be given by relation 11:
t

f GFRP (t ) Af k f , D E f , 0 (t0 ) Af E0 K ( E1 K ) e e

0 ) f GFRP , A (t )

(11)

After 17 years of loading the residual prestress force measure on site is equal to 42 kN for a
value of 41 kN using the model develop. In this case the loss due to anchorage loss may be
considered to be equal to zero. We have demonstrated how a simple experimental technique
and associated analyses can be applied to characterize the viscoelastic properties of GFRP
bars under aging and relaxation. The well known viscoelastic theory, Zener model, can well
182

describe the time-dependent relaxation behaviors and therefore facilitate the determination
of the parameters, such as elastic moduli and viscosity, from fitting the experimental data.
5.

REFERENCES

1.

Zermeno De Leon, M.E. 1988. Contribution ltude du comportement mcanique de


cbles composites. these INSA LYON 1, France.
Abbassi A., Hogg P. 2005. Temperature and environmental effects on glass fibre rebar:
modulus, strength and interfacial bond strength with concrete. Composites Part B, 36:
394-404.
Williams, M.L., Landel, R.F. and Ferry, J.D. 1955. The Temperature Dependence of
Relaxation Mechanisms in Amorphous Polymers and Other Glass-forming Liquids.
Journal of the American Chemical Society, 77: 3701-3715.
Brinson, L.C. and Gates, T.S. 2003. Viscoelasticity and Aging of Polymer Matrix,
Composites. Comprehensive Composite Materials, Chapter 2.10, pp. 333-368.

2.
3.
4.

183

184

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

DURABILITY OF GFRP BARS UNDER LOW TEMPERATURE


ENVIRONMENT AND HIGH TEMPERATURE ALKALI EXPOSURE
S.A. Sheikh, D.T.C. Johnson and A.D. Caspary
Department of Civil Engineering, University of Toronto, Toronto, Canada

ABSTRACT
An investigation was conducted to evaluate the performance of glass fibre reinforced
polymer (GFRP) reinforcing bars under extreme environmental conditions. These
conditions included 1) low temperature exposure around -40oC and 2) accelerated high
temperature alkali exposure (AAE). The objective of this study is to contribute to the
understanding of the life cycle performance and cost of a typical concrete flexural member
reinforced with internal GFRP bars.
Forty-five tests spread over three bar sizes were conducted to investigate the effects of cold
temperatures on GFRP bars. Results from the extreme cold tests showed minimal decrease
in the material properties. The maximum reduction in tensile strength was of the order of
2%. The modulus of elasticity differed by up to 10% when compared against the reference
samples. The accelerated alkali exposure consisted of 2000 hours of exposure to an alkaline
solution maintained at 60oC, while the bars were subjected to a uniform axial tensile strain
of 3x10-6 (3000 ). The study was conducted using specially constructed frames, as well as
a specifically designed system used to circulate alkali solution from a concentrated
reservoir to individual chambers surrounding the reinforcing bars.
Results from the tests on alkali exposure showed the tested bars consistently retained
tensile strength in excess of 91% of their original strength with only minimal changes to
stiffness. This study shows excellent performance of the tested GFRP bars exposed to
extreme environments indicating the likelihood of high durability during a lifespan of
normal service loads.
1. INTRODUCTION
Specifications for use of FRP bars in concrete structures in Canada (CSA S807-10, ISIS 06)
require, among other details, testing and evaluation of the cold temperature resistance as it
is theorized that extreme cold temperatures can adversely affect the mechanical properties
of FRP resulting in reduced performance of FRP-reinforced structures. Since FRP bars used
as internal reinforcement are subjected to highly alkaline environment, the Standards also
require retention of a certain minimum strength of bars after exposure to a prescribed alkali
185

solution for approximately three months. This requirement includes testing of bars without
load as well as under sustained loads. This paper reports the results of the tests conducted
on GFRP bars for these two conditions and the likely effect of reduced properties on
structural performance of a typical flexural member.
2. COLD TEMPERATURE TESTING
In this test series, bars were preconditioned down to -35oC and then tested in direct tension
in a -40oC environment. Changes in the ultimate strength, modulus of elasticity and
ultimate elongation of the FRP bars were investigated (Johnson, 2009).
2.1. Test Program for Cold Temperature
A total of 45 GFRP bar specimens were tested, the results of which were compared against
a range of control specimens. Three bar sizes were tested (8, 12 and 16 mm diameter bars)
with each size having 15 cold exposed samples. The bars were mounted with metal
couplers used for gripping the samples during testing. Couplers were mounted using a
readily available two part epoxy resin.
Because of the low thermal conductivity of GFRP, preconditioning was done in a low
temperature thermal chamber (Figure 1). The bars remained in the chamber for 96 hours
after which it could be concluded that the core temperature was sufficiently low. The
preconditioning chamber was maintained at -35oC throughout the entire process.

Fig. 1. Preconditioning chamber for cold temperature testing


After preconditioning, the bars were individually transported to a specialized test frame
which used an attached environmental chamber to maintain a -40oC environment
throughout the entire test. Samples were mounted with the couplers outside of the chamber
to keep the epoxy connecting the specimens to the couplers at room temperature during
tensile testing. The frame and chamber along with a mounted specimen are shown in Figure
2. Independent thermocouples were used to verify the internal test conditions of the
chamber throughout the test. The temperature was adjusted and corrected periodically
throughout testing to maintain the -40oC environment from start to finish.

186

Fig. 2. Test chamber and mounted 16mm specimen


Low temperature in the range of -40oC limited the use of any LVDTs, as such the
elongation and modulus of elasticity were determined using surface mounted strain gauges.
Special strain gauges for cryogenic environments and low temperature epoxies were used in
this program.
2.2. Results of Cold Temperature Testing
In general, the results of the cold exposed specimens were similar to those of the control set
regardless of size. The failure stress was the most consistent of the three mechanical
properties compared. The other two (elongation and modulus of elasticity) changed slightly
under the exposure however these changes can be primarily attributed to issues related to
measuring the elongation/strain at low temperatures. The detailed results for each size are
summarized in the sections to follow.
2.2.1 8 mm Test Results
Of the 15 tests conducted at low temperatures, a total of eleven specimens failed by rupture
of the bar, the other four failed by de-bonding of the couplers from the sample. The
mechanical characteristics of exposed specimens are compared against those from a set of
control specimens which were not exposed or tested in a cold environment. The results of
the testing are summarized in Table 1.
Table 1. Summary of results for 8mm tests (Standard Deviation in Parenthesis)

Average Reference
Samples
Average of Cold Exposed
Samples that Ruptured
Percent of Reference
Value
Average of Cold Exposed
Samples that De-bonded

Ultimate Strength
(MPa)

Modulus of
Elasticity (MPa)

Ultimate Elongation
(%)

1374 (30)

59990 (3000)

2.30 (0.17)

1370 (35)

63540 (5800)

2.10 (0.20)

99%

106%

91 %

1141 (32)

60830 (4100)

1.91 (0.18)

187

The results from the 8mm tests show a minimal change in mechanical properties, a small
increase in modulus of elasticity and resulting drop in elongation were expected. For
samples that failed by de-bonding of bars at couplers, the loads were consistently lower
than the value for those specimens that ruptured, however the modulus of elasticity was
similar regardless of failure mode. A typical de-bonding failure and a typical rupture failure
are shown in Figure 3.

Fig. 3. Failure modes for cold temperature samples


2.2.2 12-mm Test Results
For the 15 tensile tests conducted on the 12mm sized bars, a significantly higher frequency
of de-bonding failures was noticed. This was in large part due to the higher tensile force
required to fail the bar. Of the 15 samples tested, 3 failed by rupture and 12 failed by debonding, however for this size the de-bonding loads were similar and in some cases higher
than the rupture loads which indicate that the strength of the coupler for this size is similar
to the strength of the bar itself. The results of the testing are summarized in Table 2.
Table 2. Summarized 12mm Test Results (Standard Deviation in Parenthesis)
Ultimate Strength
(MPa)

Modulus of Elasticity
(MPa)

Ultimate Elongation
(%)

1128 (28)

60190 (3600)

1.88 (0.16)

1164 (26)

55165 (2800)

2.09 (0.06)

Percent of Reference Value

103 %

92%

111%

Average of Cold Exposed


Samples that De-bonded

1109 (42)

56160 (2900)

1.98 (0.10)

Average of Reference
Samples
Average of Cold Exposed
Samples that Ruptured

As was the case for the 8mm bars, again only a small difference was noted between the
control specimens and the exposed ones. The differences can be primarily attributed to
instrumentation rather than the effect of environmental conditions.
2.2.3 16 mm Test Results
The tests on 16mm bars were conducted initially with LVDTs measuring elongation and
then changed to use strain gauge measurements. Of the 15 cold exposed specimens; 8 failed
188

by rupture of the bar and 7 by de-bonding. The 8 specimens that ruptured are further
subdivided into two groups based on their instrumentation method. Specimens which used
LVDT measurements to determine the elongation are labelled as group A of which there
are 6 specimens. The remaining pair of rupture specimens used strain gauges and is labelled
as group B. As was the case with the 12mm samples, the de-bonding failure loads were
similar and in some cases higher than the rupture loads of some of the samples.
Table 3. Summarized 16mm Test Results (Standard Deviation in Parenthesis)
Reference Average
Average of Ruptured Specimens
(Group A)
Average of Ruptured Specimens
(Group B)
Percent of Reference Value
(Group A)
Percent of Reference Value
(Group B)
Average of Cold-Exposed Samples
that De-Bonded

Ultimate Strength
(MPa)
1208 (48)

Modulus of
Elasticity (MPa)
62360 (1800)

Ultimate
Elongation (%)
2.10 (0.05)

1192 (52)

53070 (2500)

2.25 (0.08)

1186

56990

2.09

99%

85%

107%

98%

91%

100%

1221 (60)

55260 (4400)

2.22 (0.22)

For the 16mm group again the ultimate strength remains unchanged and differences in the
modulus of elasticity and elongation were noted and can primarily be attributed to issues
related to instrumentation at cold temperatures.
In general for all sizes strength was not affected by the cold temperature. A variation in
rupture strain and the modulus elasticity measured between 6% and 15% was attributed to
the difficulties in instrumentation under extreme cold temperatures.
3. HIGH TEMPERATURE ALKALI EXPOSURE
CSA-S807 requires that FRP bars in high durability category should retain a minimum
strength of 80% after about 3 months (2000 hours) of exposure to alkali solution without
load and 70% after 3 months of exposure to alkali solution with sustained load causing a
strain of 3000 micro-strain. If the bars can maintain 80% strength retention under load, the
tests without load are not required.
3.1. Test Program for Alkali Exposure under Sustained Load
In this test program, 8 mm diameter bar specimens were subjected to sustained load and
alkali solution at 60oC in especially designed and built test frames as shown in Figure 4
(Caspary 2011). Each frame can accommodate up to 8 specimens under independent
sustained load measured by the built-in load cell. With an anchoring steel coupler and a
hinge at each end of the specimen and the load cell, the total specimen length in the 990
mm high frame was 560 mm. Three frames containing 24 specimens were stacked together
and enclosed in an insulated space as shown in the figure.

189

A small water-tight enclosure around the test zone of each bar specimens housed the alkali
solution at 60oC. The alkaline solution consisted of 118.5g Ca(OH)2, 0.92.4g NaOH and
4.2g KOH per one litre of deionized water providing a pH value of 12 to 13. The solution
was replaced every 24 hours using a pump and a reservoir to maintain its temperature at
60oC. The strain in each bar was measured with the help of strain gauges installed on all the
specimens. The load was adjusted at regular intervals to maintain the tensile strain at 3000
. The specimens remained under sustained load while exposed to 60oC alkali solution for
over 2000 hours when they were removed from the frames and tested for residual
mechanical properties. Nine control specimens were also stored in the laboratory at room
temperature without any exposure to alkali or load to serve as the standard against which
properties of the exposed specimens were to be compared.

Fig. 4. Test frames with specimens subjected to sustained load and alkali exposure
3.2. Test Results of Alkali Exposed specimens under Sustained Loads
All the specimens in this series were tested under tensile load in a Satec High-Capacity
universal testing machine. A total of 21 exposed specimens and nine control specimens
were tested. Additionally, mass change tests were performed on five exposed samples. A
summary of the results is given in Table 4. Exposure to alkali solution resulted in a crystal

190

built-up on the bars due to ion transfer from the alkali solution. Figure 5 shows a
comparison of the physical appearance of the exposed and the original bars.
It is clear from Table 4 that the GFRP bars displayed only a small change in the mechanical
properties tested during this program. The reduction in strength under severe exposure
conditions is about 8% while the allowable limit under sustained load and alkali exposure is
30%. There is practically no change in bar stiffness. A slight increase in the measured value
of the modulus of elasticity of the exposed specimens is attributed to the experimental
scatter and difficulty in measuring deformations. It should be noted that while load
measurements in such tests are reasonably certain, there is a large variation in strain
measurements which in most cases is due to the environmental condition to which the
gauges are exposed. The change in mass as a result of exposure to alkali solution for 2000
hours at high temperature was also found to be minimal.
Table 4. Summary of test results from alkali exposed specimens under sustained loads.
(Standard Deviation in Parenthesis)
Control Sample

Ultimate Strength
(MPa)
1439 (39)

Modulus of Elasticity
(MPa)
57307 (2800)

Alkali-Exposure Average

1324 (89)

57853 (3900)

0.85

Percent of Reference
Value

92.0%

100.9%

100.8%

Mass Change (%)


N/A

Fig. 5. Comparing appearance of exposed samples to control samples


4. EFFECT OF REDUCED GFRP PROPERTIES ON THE BEHAVIOUR OF
FLEXURAL MEMBERS
Various codes and design guidelines limit the adverse effects of environmental exposure by
providing the maximum allowable strength reduction. In the case of alkali exposure under
sustained load that limit is 30%. Figure 6 shows responses of a slab (S1) with tensile

191

strength of GFRP bars at 1200 MPa and 850 MPa which reflects a strength reduction of
29%. The slab is 220 mm thick and originally designed to fail due to

Fig. 6. Moment Curvature relations for sample slabs S1 and S2


concrete crushing with 16 mm bars at 125 mm spacing. Concrete strength is assumed to be
40 MPa. A 29% reduction in the tensile strength of bars still causes the slab to fail at the
same moment due to concrete crushing and results in the same behavior. Behaviour of
another slab (S2) with 16 mm bars at 250 mm spacing is also shown in Figure 6. Due to the
reduction in FRP strength of 29%, the failure mode of the slab changed from concrete
crushing to FRP rupture although the moment capacity is reduced only by about 10%. It
should, however, be noted that the responses shown in Figure 6 do not include any capacity
reduction factors. Inclusion of these factors in design would most likely result in no change
in mode of failure of the structural components because the capacity reduction factor for
GFRP is 0.5 and for concrete 0.75 (CSA S606).
5. CONCLUDING REMARKS
GFRP bars from one manufacturer were subjected to severe environmental treatments as
recommended by various guidelines in North American codes and standards. Results from
the tests on bars subjected to cold temperatures of -40oC and sustained loads combined with
alkali solutions at 60o C are reported here. The sustained loads produced tensile strain of
0.003 in the bars.
Results of the tests show excellent resistance of GFRP bars under cold conditions and alkali
exposure. There was no significant change in the tensile properties of GFRP bars as a result
of exposure to cold temperatures. The alkali exposure under sustained load produced a drop
in tensile strength of about 9% but no changes in bar stiffness were observed. A simple
analysis of slabs reinforced with GFRP bars shows that the changes observed in bar

192

properties as a result of tested exposures will result in no significant changes in the


behavior of the slab.
6. ACKNOWLEDGEMENT
The research reported here is funded by a CRD grant from NSERC and industry partners,
Facca Inc., Schoeck Canada Inc, Pultrall Inc. and Vector Construction Ltd. The authors
would like to thank the technical staff in the Structures Laboratories at the University of
Toronto and at Kinectrics Inc. for their assistance.
7. REFERENCES
1.
2.
3.
4.
5.

Canadian Standards Association (CSA). 2010. Specifications for Fibre-Reinforced


Polymers. CAN/CSA S807-10, Mississauga, Ontario, Canada, 35 p.
ISIS Canada, 2006. Specifications for Product Certification of FRPs as Internal
Reinforcement in Concrete Structures. Winnipeg, Manitoba, Canada, 27 p.
Johnson, D.T.C. 2009. Investigation of Glass Fibre Reinforced Polymer reinforcing
bars as reinforcement for concrete structures. MSc Thesis, University of Toronto. ON,
Canada.
Caspary, A. 2011. Investigation of short-term and long-term behavior of GFRP bars.
MSc Thesis, University of Toronto, ON, Canada. (In Progress)
Canadian Standards Association (CSA). 2006. Canadian Highway Bridge Design
Code. CAN/CSA S6-06, Mississauga, Ontario, Canada, 733 p.

193

194

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

DURABILITY OF FRP-CONCRETE INTERFACE UNDER FATIGUE


LOADING
C. Carloni1, K.V. Subramaniam2, M. Savoia3 and C. Mazzotti3
1

Dept. of Architecture, University of Hartford, West Hartford, USA


Dept. Civil & Env. Enrg., City College of New York (CUNY), New York, USA
3
Dept. DICAM, University of Bologna, Bologna, Italy
2

ABSTRACT
The topic of this research work is the fatigue behavior of externally strengthened concrete
beams with fiber-reinforced polymers. An increase of the fatigue life after strengthening is
generally reported in literature and is attributed to the reduced stress level in the steel
reinforcement due to the presence of FRP. This paper presents an experimental analysis
devoted to investigate the fatigue-induced properties of the FRP-concrete interface. Strain
patterns along the bonded length and the surrounding concrete were measured using digital
image correlation. A decrease of the effective length during cycling is observed, indicating
a different debonding mechanism during fatigue. Post-fatigue response of the FRP-concrete
interface appears to be qualitatively similar to the monotonic one.
1. INTRODUCTION
Modern transportation infrastructures are in serious need of rehabilitation as a result of
material decay, corrosion, impact damage or increase in service load. During the last two
decades, externally bonded uni-directional fiber-reinforced polymers (FRP) composite
sheets have attracted significant attention from the civil engineering community. Most of
the strengthening and retrofit schemes take advantage of the high tensile capacity of the
FRP composite. The strengthening is a result of load-sharing between the concrete structure
and the FRP composite, which is achieved through stress-transfer between the concrete
substrate and the FRP composite. Bond quality directly contributes to the effectiveness of
the stress transfer between FRP and concrete. Debonding of the FRP composite
reinforcement is the most critical concern in this type of applications. Bond characteristics
between FRP sheets and concrete have been extensively studied [1-4]. However, the
majority of these studies deal with monotonic loading or few-cycle loading at very high
force level [5]. Kim and Hefferman [6] noticed that further research is needed to address
limitations in the current knowledge of fatigue response and durability of FRP retrofitting
systems. Most of the recent studies focus their attention on the global response of FRPstrengthened beams subjected to fatigue loading [7-9]. An increase of the fatigue life after
195

strengthening is reported in literature [6] and is attributed to the reduced stress level in the
steel reinforcement due to the presence of FRP. When early debonding of the FRP sheets
occurs, the stress carried by the FRP is redistributed back to the internal reinforcing steel.
Improvement in fatigue performance can be achieved only if adequate bond between FRP
and concrete substrate is assured. In this paper, fatigue direct-shear tests of FRP strips
attached on the surface of concrete blocks are used to investigate the extent of the
interfacial debonding.
2. MATERIAL AND EXPERIMENTAL SET-UP
Direct shear tests were performed under fatigue and monotonic quasi-static loading
conditions adopting the classical pull-push configuration. The dimensions of the concrete
block and the FRP sheet are reported in Fig. 1. The 28-day compressive strength of
concrete was equal to 35 MPa. The FRP composite comprised of continuous uni-directional
carbon fiber sheet in a two-part thermosetting epoxy matrix. The tensile strength and the
Youngs modulus were equal to 3.83 and 230 GPa. The FRP composite sheet was bonded
along the centerline on one side of the concrete block using the wet-layup procedure.
y
FRP

b1

FRP

width of the FRP:


b1=25 mm

Restraint

35

Adhesive

Steel
I-Beam

Steel
I-Beam

Restraint

L=330

152

Concrete
Block

143

L=330

LVDT

nominal thickness of the


FRP: t=0.167 mm

b =125

h=125

Fig. 1. Set-up of the monotonic and fatigue tests.


The surface deformation in the FRP and surrounding concrete during the tests has been
determined from the displacement field measurements using a full-field optical technique
known as digital image correlation (DIC) [10]. A description of DIC is reported in [3,4,10].
3. MONOTONIC QUASI-STATIC TESTS
Three specimens (Test-S1, Test-S2 and Test-S3) were tested under monotonic loading
condition following the test procedure previously described [3,4]. During each test, the
global slip, which is defined as the relative displacement between points on FRP and
concrete initially located at the beginning of the bonded area, was increased at a constant
rate equal to 0.0004 mm/sec, up to failure. Global slip was measured using two LVDTs.
The typical applied load-global slip response (Test-S2) is reported in Fig. 2.

196

P crit = 7.74 kN

Applied Load (kN)

C'
A [0.040 mm, 4.00 kN]
B [0.091 mm, 5.76kN ]
C [0.163 mm, 7.23 kN ]
C' [0.163 mm, 6.64 kN ]
D [0.545 mm, 7.69 kN ]

5
4

3
2
1
0
0.00

0.20

0.40

0.60

0.80

1.00

Global Slip (mm)

Fig. 2. Typical applied load-global slip response for monotonic quasi-static tests (Test-S2).
The initial monotonic linear increase of load is followed by a non-linear relationship
between load and slip. The end of the monotonic load increase with slip is marked by a
sudden load drop (Point C in Fig. 2). The portions of the load response before and after
Point C are referred to as pre- and post-peak responses, respectively. In the post-peak part
of the load response, the load levels off at a nominal constant value equal to 7.74 kN that is
identified as Pcrit. Pcrit was determined as the mean value of the load when the global slip
varied between 0.4 and 0.8 mm. The average Pcrit of the three static tests was 7.69 kN.
The axial strain distribution along the centerline of CFRP composite during the post-peak
response corresponding to point D is shown by the dotted curve in Fig. 3. [1, 3, 4]. The
approximated strain distribution using the function reported in [1] is also shown in Fig. 3.
0.008

yy

LSTZ

0.007
0.006

yy

0.005
0.004
0.003
0.002

Point D (figure 2) - Experiental Data

0.001

Point D (figure 2) - Fitting Curve

0
-0.001
0

20

40

60

80

100

120

140

160

y (mm)

Fig. 3. Axial strain along the FRP composite at point D of the load response in Fig. 2.
During debonding, the strain levels off at a value yy approximately equal to 6700 . The
average value of for the three monotonic quasi-static tests is 6500 . The strain
distribution can be divided into three main regions: (a) the unstressed region; (b) the stress
transfer zone (STZ); and (c) the fully debonded zone. Analysis of the strain data revealed
197

that the stress transfer zone was fully established when the load response attains Pcrit. Once
the STZ was fully established, as the global slip increased, there was a translation of the
STZ further along the length of the FRP sheet while its shape remained constant [1-4].
Within the range of the global slip 0.4mm 0.8mm used to compute Pcrit, the stress transfer
zone was fully established. The average value of the estimated length of STZ for the three
monotonic quasi-static tests is 80 mm.
The cohesive law, which relates the interface shear stress with the relative slip at each point
of the bonded area, can be obtained using the procedure outlined by Taljsten [1-4,11]. From
the measured strain yy along the FRP, the interface shear stress zy was calculated as:

zy Et

d yy

(1)

dy

where E and t are the Young modulus and the thickness of the FRP strip, respectively. The
relative slip, s(y), between FRP and concrete at a given location along the FRP is obtained
by integrating the axial strain in the FRP from the unloaded end up to that point, and
assuming it is zero at the unloaded end. The average values of the maximum shear stress
(max) and the corresponding interfacial slip (so) for the three monotonic quasi-static tests
are found to be equal to 6.43 MPa and 0.043 mm, respectively. The interfacial fracture
energy GF is obtained from the area under the entire zy s curve:
sf

GF zy ds

(2)

The average value of GF for the three monotonic quasi-static tests is 0.80 MPamm.
4. FATIGUE TESTS
Three direct shear tests (Test-F1, Test-F2 and Test-F3) were performed under fatigue
loading. The maximum (Pmax) and minimum (Pmin) applied loads are defined as a
percentage of the ultimate load Pcrit determined in the quasi static tests. If Pmax is less than
50% of the failure load, the test is said to be performed under a low-level cycling
conditions; otherwise the test is said to be performed under high-level cycling conditions.
High-level cycling has been investigated in this paper. The load range was 15%-80%, 15%
-70%, and 15%-60% for Test-F1, Test-F2 and Test-F3, respectively. Fatigue cycling was
performed in load control at 1 Hz. The load-slip response data were collected for ten cycles
at the prescribed intervals of 0.05 mm threshold value in the global slip. Failure of all tested
specimens was associated with progressive debonding of FRP up to a complete separation.
The collected load cycles of Test-F1 specimen are plotted versus the global slip in Fig. 4.
The hysteresis curves are qualitatively similar. The area enclosed by the loop represents the
energy dissipated in the cycle. In Test-F1, this area is nominally constant in the first half of
the test whereas it decreases in the second half while the loading-unloading behavior
becomes approximately parallel. Yun and Wu [12] observed that the area enclosed by the
loops does not change significantly when externally-bonded systems are tested between
198

Load (kN)

10% and 45% of the ultimate load. It should be noted that the total number of cycles at
failure for Test-F1 was 1290, which is significantly smaller than the number of cycles
reported in literature [7-9,12,13]. A smaller number of cycles can be attributed to the range
of fatigue load and the frequency. Bizindavyi et al. [13] tested all specimens at 1Hz.
Gheorghiu et al [8,9] tested at 2Hz or 3Hz depending on the load range and the maximum
number of cycles. Yun and Wu [12] tested externally-bonded systems at 5Hz.
7
6.5
6
5.5
5
4.5
4
3.5
3
2.5
2
1.5
1
0.5
0

Cycle# Cycle# Cycle# Cycle# Cycle# Cycle# Cycle# Cycle# Cycle# Cycle#
210
243
558
956
1055 1246
375 394
676
795

Pmax = 6.00 kN
Pmin = 1.25 kN

0.1

0.2

0.3

0.4

0.5

Global Slip (mm)

0.6

0.7

0.8

0.9

Fig. 4. Applied load vs. global slip response of Test-F1.


The total number of cycles for Test-F2 and Test-F3 was 13192 and 116995, respectively.
The number of cycles dramatically increases if the fatigue-load range decreases. The
enclosed loop area increases in Test-F2 and Test-F3. The ratio of the area of the first block
of cycles over the area of the last block of cycles was calculated to be 1.2 for Test-F2 and
1.38 for Test-F3, indicating that as the load range decreases the dissipated energy increases
within the test as the number of cycles increases.
The slopes of the loops in Test-F1, Test-F2 and Test-F3 are found to be gradually
decreasing with the number of cycles, indicating a progressive loss in the stiffness of the
interfacial bond between CFRP and concrete substrate. Yun and Wu [12] reported that
under low-level fatigue load, the slopes of the hysteresis loops remain unchanged, except
for early loops which show a slightly decrease of the slope. Bizindavyi et al. [13] reported a
decrease of the slope of the hysteresis loops. The rate of change of slope was observed to
increase with increasing load amplitude. Gheorghiu et al. [8,9] reported a gradual decrease
of the slope for low-level cycling and a constant slope for high-level cycling. The average
slope for each block of ten cycles was calculated from the load response as the load-toglobal slip ratio and is plotted in Fig. 5 as a function of the fatigue life for the three tests. N
is current number of cycles and Nf the total number of cycles at failure. At initial load
cycles, a sharp reduction in slope is observed, which can be attributed to crack initiation at
the notch close to the free end of the concrete specimen. As the interfacial crack
propagates, the decrease in slope becomes smaller and levels off at a constant value
k f 7.5 kN/mm . It is worthy noticing that the decrease in the slope can be partially
attributed to the fact that during fatigue loading the FRP strip debonds from the substrate,
therefore at any value of the global slip the contribution of the axial deformation of the FRP
199

strip between the remaining bonded area and the position of the LVDTs should be taken
into account in reading the global slip.

Fig. 5. Average cycle slope kf vs. fatigue cycles for Test-F1, Test-F2 and Test-F3.
5. POST-FATIGUE QUASI-STATIC LOAD RESPONSE
A quasi-static test was performed on a further specimen (Test-F4), previously subjected to
fatigue loading between 15% and 80% of the average quasi-static peak load (Pcrit). The
fatigue cycling was performed in load control at 1 Hz until a prescribed threshold of the
global slip equal to 0.4 mm was obtained. The specimen was then unloaded to a load equal
to 0.5 kN and finally monotonically loaded at a rate equal to 0.0004 mm/sec until failure.
The post-fatigue monotonic load response is shown in Fig. 6, together with the response of
Test-S2.
9
8

P crit = 6.80 kN

Applied Load (kN)

3
A

1
0
0.00

0.20

0.40

F
A [0.264 mm, 2.71 kN]
B [0.416 mm, 5.53 kN]
C [0.469 mm, 6.02 kN]
D [0.508 mm, 6.10 kN]
D' [0.514 mm, 5.75 kN]
E [0.617 mm, 6.51 kN]
F [0.701 mm, 6.75 kN]
G [0.864 mm, 6.94 kN]
H [0.991 mm, 7.06 kN]

0.60

0.80

1.00

1.20

Global Slip (mm)

Fig. 6. Fatigue and post-fatigue response of Test-F4, and quasi-static response of Test-S2.
From the post-fatigue response curve, it is observed that as the average slip is increased, the
load response is nominally linear up to point B. A sudden drop in the load response at D is
seen before the response levels off and essentially remains constant at Pcrit equal to 6.80
kN. Pcrit was calculated within the range of the global slip 0.6 mm 1.0 mm. The strain
200

distributions of the post-fatigue response using the procedure reported in [1] are shown in
Fig. 7. It can be seen that the strain distribution in post-fatigue post-peak monotonic test
response is qualitatively similar to that in the monotonic test [1-4]. The length of the stress
transfer zone from the post-fatigue monotonic test is 80 mm.
0.008

0.005
0.004

Fully debonded during fatigue

0.006

yy

Stress transfer
zone at point C
(Figure 6)

Point A
Point B
Point C
Point D
Point E
Point F
Point G
Point H

0.007

0.003
0.002
0.001
0
0

20

40

60

80

100

120

140

160

y (mm)

Fig. 7. Strain distribution along CFRP sheet in the post-fatigue response of Test-F4.
The length of the stress transfer zone in the pre-peak region of the post-fatigue response is
smaller than that observed in the pre-peak region of the quasi-static monotonic tests. This
fact could imply that a different transfer mechanism at the interface occurs during fatigue.
The shear stress-relative slip curves [1-4,11] obtained from different points (F, G, H) on the
post-peak response are plotted in Fig. 8.
8
7

zy (MPa)

6
Point F

Point G

Point H

Point D (Test-F1)

2
1
0
0

0.1

0.2

0.3

0.4

0.5

0.6

s (mm)

Fig. 8. Cohesive material law (at points F,G,H of the post-fatigue response of Test-F4)
compared with the result from Test-F1.
The curves closely agree, confirming the self-similarity of the crack propagation process in
the post-peak response. The average maximum interface shear stress (max) and
corresponding interfacial slip (so) are 6.46 MPa and 0.035 mm, respectively. Comparison
with the material law obtained from point D of Fig 2 indicates that prior fatigue loading has
no significant effect on the cohesive stress response during debonding process. The
interfacial fracture energy (GF) was calculated to be 0.64 MPa*mm.
201

6. CONCLUSIONS
The observations made in this paper can be summarized as follows:

The critical loads in the post-fatigue quasi-static monotonic response are similar to
those in the monotonic load response.

A significant decrease in the interfacial bond stiffness is observed during fatigue tests.

The strain analysis in the pre-peak response of the post-fatigue loading revealed that a
different debonding mechanism might occur during fatigue.

The STZ length between the FRP sheet and concrete in the post-peak region of the
post-fatigue monotonic loading is identical to that in the monotonic loading.
REFERENCES
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.

Ali-Ahmad, M., Subramaniam, K.V., and Ghosn, M. 2006. Experimental Investigation


and Fracture Analysis of Debonding between Concrete and FRP. Journal of
Engineering Mechanics, ASCE, 132 (9): 914-923.
Ferracuti, B., Savoia, M., and Mazzotti, C. 2007. Interface law for FRP-concrete
delamination. Composite Structures, 80 (4): 523 531
Subramaniam K.V., Carloni C., and Nobile L. 2007. Width effect in the interface
fracture during debonding of FRP from concrete, Engineering Fracture Mechanics,
74(4): 578-594.
Carloni, C., and Subramaniam K.V. 2010. Direct determination of cohesive stress
transfer during debonding of FRP from concrete. Composite Structures, 93: 184-192.
Mazzotti, C., Savoia M. 2009. FRP-concrete bond behaviour under cyclic debonding
force. Advances in Structural Engineering, 12: 771 - 780.
Kim, Y.J. and Heffernan, P.J. 2008. Fatigue Behavior of Externally Strengthened
Concrete Beams with Fiber-Reinforced Polymers: State of the Art. Journal of
Composites for Construction, 12 (3): 246-256.
Aidoo, J., Harries, K. A., and Petrou, M. F. 2004. Fatigue behavior of carbon fiber
reinforced polymer strengthened reinforced concrete bridge girders. Journal of
Composites for Construction, 8 (6): 501509.
Gheorghiu, C., Labossiere, P., and Proulx, J. 2006. Fatigue and monotonic strength of
RC beams strengthened with CFRP. Composites: Part A, 37: 1111-1118.
Gheorghiu, C., Labossiere, P., and Proulx, J. 2007. Response of CFRP-strengthened
beams under fatigue with different load amplitudes. Constr. Build. Mat., 21: 756-763.
Sutton, M.A., Wolters, W.J., Peters, W.H., Ranson, W.F. and McNeill, S.R. 1983.
Determination of Displacements using an Improved Digital Correlation Method. Image
and Vision Computing, 1(3): 133-139.
Taljsten, B. 1997. Defining Anchor Lengths of Steel and CFRP Plates Bonded to
Concrete. International Journal of Adhesion and Adhesives, 17(4): 319-327.
Yun, Y., Wu, Y.F. and Tang, W.C. 2008. Performance of FRP systems under fatigue
loading. Engineering Structures, 30: 3129-3140.
Bizindavyi, L., Neale, K.W., and Erki, M.A. 2003. Experimental investigation of
bonded fiber reinforced polymer-concrete joints under cyclic loading. Journal of
Composites for Construction, 7 (2): 127134.

202

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

PHENOMENOLOGICAL AND EXPERIMENTAL STUDY OF


PULTRUDED UD FRP COMPOSITES STATIC AND DEFERRED
RUPTURE UNDER COMBINED FLEXURAL, COMPRESSIVE AND
TORSION LOADING
N. Kotelnikova-Weiler1 and J.-F. Caron2
1
2

PhD student, Laboratoire Navier, Ecole des Ponts ParisTech, France


Professor, Laboratoire Navier, Ecole des Ponts ParisTech, France

ABSTRACT
In structural applications the use of composite materials is subject to strict safety
restrictions. It is of particular importance to know exactly the modes and causes of deferred
rupture. Glass fibre reinforced vinylester pultruded composites were studied under
sustained loads. Thin cylindrical samples were tested with flexural, combined torsionflexion and combined torsion-compression loads. Tests were conducted at room and
elevated temperatures (60C). The experimental study shows two different types of rupture
modes for different loadings: progressive and quasi-instantaneous. The use of torsion
loading allows a better understanding of the matrixs role in deferred rupture of composite
materials. A model including resin viscoelastic behaviour is being developed to predict the
lifetime of a composite under constant load.
1.

INTRODUCTION

The interest in composite materials for various applications in aeronautical, naval and civil
engineering fields is constantly increasing. These materials offer designers the possibility to
associate form and function within innovative systems and structures. New design and
manufacturing processes allow the extension of technical possibilities and better
satisfaction of sometimes conflicting demands (weight, function, mechanical properties)
that traditional homogeneous materials fail to meet. Light weight, anisotropy, high
mechanical properties, multifunctionality are the main advantages of fibre reinforced
polymers. Development of pultrusion technique made these materials economically
adequate for civil engineering applications. However in many countries design standards
and regulations for composite structures are not completely established. Although structural
calculations using FRPs are mastered, lifespan predictions and safety factors still need to be
adjusted. Currently composite materials are mainly used in non-bearing components or as
reinforcement of existing steel or concrete structures and when an entirely composite
structure is designed, its initial loading is limited to a maximum of 30% of its ultimate
203

stress as prescribed in Eurocomp design guidelines. In order to achieve accurate lifespan


predictions, it is essential to understand the role of different composites constituents in
creep rupture and to build both efficient and parameters-sensitive models. These models
shall link environmental conditions and material properties of the composite to its
mechanical response and long-term behaviour.
Some authors have worked on the modelling of strength reduction in composites subject to
sustained traction loading. Though the model currently developed is not presented in this
paper, this study particularly concentrated on shear-lag models as they are efficient and
easily implemented. Ohno et al. in 1999 [1] have developed a model allowing the
prediction of a composites lifespan considering its load in traction. Koyanagi et al. in 2007
[2] have also developed a similar model. Zhou et al. in 2002 [3] and 2004 [4] lead an
experimental study of a model composite consisting of graphite fibres embedded in epoxy
matrix. They used Raman spectroscopy and tested their specimens in traction at
temperatures ranging from 20C to 80C. Their results provided a micromechanical view of
the creep and deferred rupture processes. Most of the models cited consider elastic fibres
and matrix and periodic fibre arrangements. Yet as we will see later it is the viscoelasticity
of the matrix that leads to deferred rupture. Beyerlein et al. 1998 [5] develops a 2D model
including viscoelasticity of the matrix, stochastic strengths distribution for fibres and allow
modelling of successive fibre breaks in time. In 2009 Blassiau et al. [6] developed a 3D
finite elements micromechanical model including matrix viscoelastic behaviour, stochastic
distribution of fibres strength and interface rupture between a fibre and surrounding
matrix. This model considers a periodic damage state of the composite.
Our work aims to investigate different rupture modes and understand the role of
composites constituents on the rupture type and the lifespan of the structure subject to a
given loading. First we will present experimental study carried out to explore different
rupture mechanisms. This study enabled us to identify two distinct rupture modes. In the
second part we will show the results of a larger experimental campaign including creep
tests under combined torsion-flexion loadings at different temperatures.
2.

STATIC RUPTURE MODES IDENTIFICATION

2.1 Test Samples


Samples are unidirectional pultruded thin rods composed of E-glass fibres (roving) and
vinylester resin. We first identified their mechanical characteristics such as strength, strain
at rupture, Youngs modulus, fibre volume content and glass transition temperature. In
order to determine the strength of the samples, a four-point bending test was performed on
the Instron testing machine (fig. 3d). In order to prevent the sample from being crushed,
silicon pads were placed under the inner loading noses. Strain was either measured using
strain gages or calculated from the crosspiece displacement data. Bending tests provided us
with the information on the flexural strength, maximum strain and Youngs modulus.
Specific tests (loss on ignition and dynamic mechanical spectroscopy) were carried out in a
partner laboratory to determine fibre volume fraction and glass transition temperature.
Data for the dynamic mechanical spectroscopy is given in the figures below (Fig. 1).

204

E' jonc 5 mm

Tan jonc 5mm


0,16

2,50E+10

E' jonc 5 mm

Tan jonc 5mm

0,14

Module E' (Pa)

Tan

0,12
0,1
0,08
0,06
0,04

2,00E+10
1,50E+10
1,00E+10
5,00E+09

0,02
0

0,00E+00
0

50

100

150

200

250

300

50

100

150

200

250

300

Temprature (C)

Temprature (C)

Fig. 1. Dynamic Mechanical Spectroscopy results.


Additional tests were carried out to determine the variation of flexural strength upon
temperature. The strength tends to decrease but this phenomenon is difficult to quantify
because of the data scattering. Figure 2 shows the results of these tests.

flexural strength (MPa)

strength vs temperature
1800
1600
1400
1200
1000
800
600
400
200
0
0

10

20

30

40

50

60

70

temperature (C)
5mm

Linaire (5mm)

Fig. 2. Evolution of bending strength with temperature.


These tests show that between 20C and 60C, mechanical properties of the composite
material studied can be considered as constant. This range of temperatures will be
considered in this study as it also corresponds to the application temperatures of this type of
composite materials.
Material data is summarised in the following table (Table 1).
Table 1. Mechanical properties of the material studied.
Rod diameter max
max E
Vf
Tg
5mm
1499MPa 2.6% 56GPa 56.08% 137C
2.2 Experimental Devices
The industrial partner of the project noticed a new rupture mode in their applications. In
these applications the material is mainly subject to bending. The ruin mode observed was
deferred and abrupt, it lead to an instantaneous separation of the two pieces of the sample.
In order to identify the loadings that lead to this particular rupture mode, a series of tests
under various static loadings were carried out. In this section different experimental devices
used will be described.
205

Bending tests were performed on the same Instron testing machine as the one used for
strength identification tests. Samples were tested in a four-points bending device (Fig. 3a).
Compression-torsion tests were performed on a hydraulic MTS testing machine (Fig. 3b).
For this test the samples section was locally reduced around the middle of the sample as
shown on the figure 4a. To apply torque to the specimens, we inserted them in aluminium
holders with a screw on the side. A plane surface machined on the sample's side allowed
gripping it inside the aluminium holder, preventing it from sliding. A schematic shows the
assembly figure 4b.Using the same testing machine but with a different gripping system
(shown in figure 4c) a combined torsion-traction could be applied to the samples.
Pure torsion was applied using a specific device shown in figure 3c. Using this device one
can apply a given torque rotation (measured in degrees) to a sample of a given length.

a)

b)

c)

Fig. 3. a) Instron testing machine; b) Hydraulic MTS testing machine; c) Torsion testing
device.

a)

b)

c)

Fig. 4. a) Aluminium holder assembly; b) Specimen in the MTS machine; c) Specific


holders used for torsion-traction

206

2.3 Identification of Various Rupture Types


Rupture modes identified in these tests differ tremendously. In the bending test materials
ruin was progressive and occurred through continuous detachment of thin fibre-matrix
bundles (Fig. 5a). In pure torque test longitudinal cracks appeared (Fig. 5b). When
combined with traction, torsion leads to the ruin of the material through the separation of
large bundles of material by continuous longitudinal cracks (Fig. 5c). In the pure
compressive case the rupture is a well-known kink-band type and the two parts of the
sample hold together (Fig. 5d). Whereas in torsion-compression test specimen abruptly
broke in two parts when a certain torque was applied given a compressive load (Fig. 5e).
In neither cases could we observe the instantaneous crushing of the specimen as the one we
can see on Figs. 5e and f. It is the only case where rupture was both abrupt and leading to
the instantaneous separation of the two parts of the specimen (the residual strength is null).
It should be noticed that torsion-compression and torsion-traction lead to two very different
modes of rupture. The next step was to investigate how torque influenced rupture process in
combined flexion torsion loading where both torsion-traction and torsion-compression
occur.

a)
b)
c)

d)

e)

f)

Fig. 5. Rupture modes in a) Bending; b) Pure torsion; c) Combined traction-torsion; d) Pure


compression; e) Combined torsion-compression; f) Detail of rupture under combined
torsion-compression

207

3.

DEFERRED RUPTURE TESTS UNDER COMBINED TORSION-BENDING


LOADING.

3.1 Experimental Device


Torque plays clearly an important role in the modification of the ruin type of the materials
as shown in the preceding section. As in the industrial applications studied the material is
mainly subject to bending, a combined torsion-bending test was carried out. This time the
deferred, not static, rupture was investigated.
For the combined torsion-bending test a specific device was used. In what follows we will
refer to it as elastica as it has the form of the Eulers post-buckling elastica. To create the
elastica presented in Fig. 6a, we join together the two ends of the cylindrical test sample.
Then we attach them using aluminium holders in the same manner as described in the
previous section (inserting a screw preventing the rod from sliding). By rotating one end
360 in relation to the other end, torque is applied to the sample. This means that torque
level depends directly on the length of the cylindrical sample.
Some samples break in less then a minute. In order to record precisely the lifespan of these
samples, we place electrical wires as shown in Fig. 6a on the specimens. Variation of the
voltage of these wires is acquired. When rupture occurs one of the wires is cut. This
modifies the tension observed.
Electrical wires

a)

b)

c)

Fig. 6. a) "Elastica" sample; b) Oven with controlled temperature; d) "Elastica" samples in


the oven
Samples are placed in an oven where the temperature is stabilised at 60C. When a sample
breaks we record its lifespan under the given flexural and torque loading. Lifespan can then
be expressed in terms of probability of rupture after a certain period for a given load.
Elastica has a complex loading that varies depending on the curvilinear abscissa. The
length of the rod sample used determines the level of maximum bending stress applied at
the middle of the sample as well as the torque level. The following graphics present the
variation of the applied forces and moments depending on the curvilinear abscissa of the
sample (Fig. 7). The first graphic presents the normalised efforts in the general case (Fig.
208

7a) whereas the second (Fig. 7b) and third (Fig. 7c) give an idea of the efforts applied on
the sample for a specific length. Here we chose to present the case where the maximum
bending moment at the middle of the sample equals 60% of the static ultimate bending
moment. This condition gave us the length of the cylindrical sample to be used and
therefore the torque obtained. The level of torque corresponds approximately to 50% of the
static strength in torsion. We also calculated shear stress due to shear force. But it only
represents approximately 2% of the shear stress due to torsion, it is therefore not
represented on the graphics.
Normalised loading

normalised loadings

150,00%
100,00%
50,00%
0,00%
0,00%
-50,00%

20,00%

40,00%

60,00%

80,00%

100,00%

-100,00%
-150,00%

normalised curvilinear abcissa s/smax

Normalised axial force


Normalised bending moment

a)

Normalised shear force


Normalised torque

Bending moment and torque

Axial and shear forces


40

30

10
0
-10 0

0,2

0,4

0,6

-20

0,8

moment (N.m)

force (N)

20

-30

0,2

0,4

0,6

0,8

-4
-6
-8
-10
-12

-40

-14

curvilinear abscissa s (m)

b)

-2

Axial force

curvilinear abscissa s (m)

Shear force

Bending moment

Fig. 7. a) Normalised efforts in the elastica test; b) Axial and shear forces, bending
moment and torque in the sample with the maximum bending moment of 60% of the static
strength
4.

RESULTS AND DISCUSSION

4.1 Experimental Observations


On samples subject to torsion-flexion we observed an abrupt but deferred rupture. A part of
the broken section presents a clear cut. This probably corresponds to the compressiontorsion part. In the rest of the section we do not simply observe longitudinal cracks as we
did in torsion-traction tests but we rather see a concentrated rupture with fuzzy edges
recalling the rupture in pure bending (Fig. 8a). We also made elastica tests without
torsion in order to compare the rupture modes. In this pure bending, the elasticas break as

209

expected, progressively and in the same manner as samples tested in a four-points bending
test (Fig. 8b).
We can therefore conclude that application of torsion leads to the modification of rupture
mode in bending. The level of torsion to be applied in order to do so is still investigated.

a)

b)
Fig. 8. a) Rupture of an elastica with torsion; b) Rupture of an elastica without torsion
The data acquisition device enabled us to record the lifespan of different samples for
various loadings. The following graphic (Fig. 9) presents the results obtained for samples
loaded with the maximum bending moment (at the middle of the sample) of 70% and 60%
with 360 rotation applied between the two ends of the rod specimen. We can see that
greater torsion and flexion levels lead to a dramatic reduction of lifespan.
Probability of rupture for different loading
1,2

probability (%)

70% maximum
bending, 360
rotation
60% maximum
bending, 360
rotation

0,8
0,6
0,4
0,2
0
1

10

100

1000

10000 100000 1E+06

log(s)

Fig. 9. Lifespan of samples subject to different loadings


4.2 Micromechanical Process of Deferred Rupture
The microscopic mechanism generally admitted to explain the deferred rupture is as
follows: when a unidirectional composite is loaded, a certain amount of fibre breaks occur
as the resistance of fibres is non-uniformly distributed. A fibre will have weak points due to
flaws and defects in its structure. When loaded at a sufficiently high level, this section will
yield. Around fibre ends (that exist whatever the load level) and fibre breaks, the matrix is
210

loaded in shear thus re-distributing the load to intact fibres close enough to be affected. The
viscoelasticity of the matrix causes its shear stress to relax leading to the expansion of the
over-load profiles on the intact neighbouring fibres (it is also causing a global viscoelastic
behaviour). Considering such an intact fibre, suppose it is subject to two different non overlapping over-load profiles due to two different fibre breaks (Fig. 10a). Although the
magnitude of each over-load profile is decreasing in time, their expansion finally leads to
their superposition (Fig. 10b). In the range of the superposition area this means that locally
a section sees its stress level increase in time. This might lead to the break of this fibre.
Each new fibre break weakens the whole composite decreasing its strength. This could
finally lead to a deferred rupture. A modelling similar to [5] of these successive ruptures is
in process.

a)

b)

Fig. 10. Schematics on the left represent the two broken fibres and the over-load profiles
caused by their rupture on neighbouring intact fibres. On the right - the stress profiles in the
intact fibre between the two ruptured fibres. a) When the two ruptures just occurred; b)
After a long period of time.
When additional shear is added through torsion, this process accelerates in time and leads
to localisation of the rupture. The matrixs relaxation characteristic time depends directly
on the shear stress level applied, therefore when additional shear stress due to torsion is
applied, the matrixs relaxation is stimulated leading to quicker fibre breaks. Thus this
combined test reveals the qualities or flaws of the matrix by applying greater stresses on it.
Different viscoelastic properties of the matrix will lead to different lifespan in this test.
5.

CONCLUSION AND PERSPECTIVES

We investigated different rupture modes under varied simple and combined loads. Two
different rupture modes were identified: a progressive occurring in flexion and an abrupt
ruin occurring when torsion is added. We showed that torque loading could modify rupture
mode and severely decrease lifespan of a sample subject to bending. Combined torsionflexion testing enables to reveal qualities or defects of the matrix as it is directly subject to
important stress levels. Torsion introduces shear stress that combined with local shear stress
due to fibre breaks leads to an acceleration of the deferred rupture.
In order to predict lifespan of unidirectional composites a model including viscoelastic
behaviour of the matrix is needed. This model should enable calculations in combined
loadings such as torsion-flexion. In combination with the results from previous tests a
precise identification of the matrix in situ behaviour could be accessible. Then calculations
on other loadings would be more accurate. Such a model is being developed based on the
works of Beyerlein and her colleagues [5].
211

REFERENCES
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.

13.

Ohno, N. and Miyake, T. 1999. Stress relaxation in broken fibers in unidirectional


composites: modelling and application to creep rupture analysis. International Journal
of Plasticity, 15: 167-189.
Koyanagi, J., Hatta, H., Ogawa, F., and Kawada, H. 2007. Time-dependent Reduction
of tensile strength caused by interfacial degradation under constant strain duration in
UD-CFRP. Journal of Composite Materials, 41(25): 3007-3026.
Zhou, C.H., Schadler, L.S., and Beyerlein, I.J. 2002. Time-dependent micromechanical
behaviour in graphite/epoxy composites under constant load: a combined experimental
and theoretical study. Acta Materialia, 50: 365-377.
Zhou, C.H., Schadler, L.S., and Beyerlein, I.J. 2004. Stress concentrations in
Graphite/epoxy model composites during creep at room temperature and elevated
temperatures. Journal of Composites Materials, 38(5).
Beyerlein, I., Phoenix, L., and Raj, R. 1998. Time evolution of stress redistribution
around multiple fiber breaks in a composite with viscous and viscoelastic matrices.
International Journal of Solids and Structures, 35(24): 3177-3211.
Blassiau, S., Thionnet, A., and Bunsell, A.R. 2009. Three-dimensional analysis of load
transfer micro-mechanisms in fibre/matrix composites. Composites science and
technology, 69: 33-39.
Foret, G. 1995. Effets d'chelle dans la rupture des composites unidirectionnels ,
PhD thesis of the Ecole Nationale des Ponts et Chausses.
Landis, C.M., McGlockton, M.A., and McMeeking, R.M. 1999. An improved shear-lag
model for broken fibers in composite materials. Journal of Composite Materials, 33(7):
667-680.
Landis, C.M. and McMeeking, R.M. 1999. Stress concentrations in composites with
interface sliding, matrix stiffness and uneven fiber spacing using shear lag theory.
International Journal of Solids and Structures, 36(28): 4333-4361.
Landis, C.M., Beyerlein, I.J., and McMeeking, R.M. 2000. Micromechanical
simulation of the failure of fiber reinforced composites. Journal of the Mechanics and
Physics of Solids, 48(3): 621-648.
Nairn, J.A. 1997. On the use of shear-lag methods for analysis of stress transfer in
unidirectional composites. Mechanics of Materials, 26: 63-80.
Okabe, T., Takeda, N., Kamoshida, Y., Shimizu, M., and Curtin, W.A. 2001. A 3D
shear-lag model considering micro-damage and statistical strength prediction of
unidirectional fiber-reinforced composites. Composites science and technology, 61:
1773-1787.
Richard, F., and Perreux, D. 2001. The safety-factor calibration of laminates for longterm applications: behaviour model and reliability method. Composites Science and
Technology, 61: 2087-2094.

212

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

FATIGUE PERFORMANCE OF CFRP POST-TENSIONED TENDONS


A. El Refai1, J. West2 and K. Soudki3
1

Assistant Professor, Civil and Water Engrg. Dept., Universit Laval, QC, Canada
Associate Professor, Civil & Env. Engrg. Dept., University of Waterloo, ON, Canada
3
Professor & Canada Research Chair, Civil & Env. Engrg. Dept., Univ. of Waterloo, ON, Canada
2

ABSTRACT
The cyclic performance of carbon fibre reinforced polymer (CFRP) tendons was
investigated as part of a broad research program on the long-term performance of CFRP
externally post-tensioned concrete beams. The tests were carried out in two phases. The
first phase consisted of direct tensile fatigue tests on the CFRP tendon-anchor assemblies.
Stress ranges varying between 7% and 17% of the ultimate tensile strength of the tendon
were applied. Two minimum stress levels representing 40% and 47% of the ultimate
strength were considered. The results showed that the number of cycles survived by the
specimens was most affected by the applied stress range with a little influence of the
minimum stress. All specimens withstood an infinite fatigue life at a stress range below
10% for a maximum stress of 50% of its ultimate capacity. The second phase of the study
observed the fatigue performance of the CFRP post-tensioned tendons used to externally
strengthen overloaded concrete beams. During these tests, two tendons were installed
externally on reinforced concrete beams and post-tensioned to 40% of their ultimate
capacity prior to applying the loads on the beams. The tendons were harped at two points
along the sides of the beams and externally anchored at both ends by wedge-type steel
anchors. The CFRP tendons exhibited excellent fatigue performance during the cyclic
loading of the beams, and no tendon failures occurred. No signs of distress were observed
at the harping or anchor locations. This research demonstrated the excellent fatigue
performance of the CFRP tendon and wedge-type anchor system for strengthening concrete
beams subjected to fatigue loading.
1. INTRODUCTION AND BACKGROUND
External post-tensioning implies the placement of the tendons outside of the concrete beam
cross-section. Fibre reinforced polymer (FRP) tendons, with their non-corroding
characteristics, have been proposed as an alternative to steel tendons in external posttensioning applications such as bridges, parking garages, and long-span structures. These
types of structures are typically expected to resist cyclic loading during their lifetime,
which may raise concerns about the fatigue performance of the tendons. Despite the low
variation in the stresses exhibited in the external tendons during service, the locations
213

where the tendons are anchored and harped are critical points where premature fatigue
failure may occur.
FRP tendons are weak in the transverse direction despite their high tensile strength in the
direction of the fibres. They are vulnerable to premature failure at the anchor zone if any
notching in the tendon occurs during the post-tensioning process or during service. It is thus
crucial to examine the fatigue performance of the tendon-anchor assembly prior their use
in practical applications.
Sayed-Ahmed and Shrive (1998) reported on fatigue tests of stainless steel wedge-type
CFRP tendon-anchor systems. Four rates of loading ranging from 1 to 10 Hz were
considered. It was reported that the stress range rather than the mean stress had the most
significant effect on the life of the specimens. Taha and Shrive (2003) conducted fatigue
tests on six wedge-type anchors made of ultra-high-performance concrete (UHPC) concrete
wrapped with CFRP sheets. The tests were carried out in conformance with the PostTensioning Institute standards (1997) developed for steel tendons. Two tendons were
reported to have failed at the anchor due to fatigue.
Another location of potential concern for CFRP external tensions is the deviation or
harping locations. When the FRP tendons are harped, localized curvature stresses are
developed in the tendons leading to a high stress concentration at these locations. Due to
the linear-elastic behaviour of FRP up to failure, the increase in stress at the curved points
may lead to premature failure at the harped points. The reduction in the tensile strength of
the FRP tendons depends mainly on the tendon diameter, the deviator radius, and the
harping angle (Grace and Abdel-Sayed 1998, Dolan et al. 2001). Quayle (2005) conducted
tensile tests on CFRP specimens with 24 combinations of these three parameters. Three
modes of failure were observed during these tests: bending-tension, bending-compression,
and bending-shear failure, with the last two modes occurring at relatively low load levels.
The reduction in the tensile strength of the specimens ranged between 13% and 50% of the
tendon ultimate capacity. It was reported that increasing the harping angle or decreasing the
deviator radius decreased the tensile capacity of the harped tendons.
In the current study, the performance of CFRP tendons subjected to fatigue loading was
investigated on two levels: (a) the tendon-anchor assembly, where fatigue tests with various
load ranges were conducted on CFRP tendon-anchor specimens, and (b) externally posttensioned beams, where harped CFRP tendons were subjected to fatigue stresses while
being used to strengthen reinforced concrete beams. The tendons used in all the tests were
Aslan200TM CFRP tendons of 9.4 mm diameter. The nominal ultimate strength of the
tendon material is 2162 MPa with a modulus of elasticity of 144 GPa and an ultimate strain
of 0.016%.
2. TENDON-ANCHOR FATIGUE TESTS
2.1 Test Specimens and Test Setup
The test specimens shown in Figure 1 consisted of CFRP tendons of 1000 mm length
gripped at both ends with a wedge-type steel anchor, known as the Waterloo anchor.
214

Details of the anchor components and the anchoring procedure are found in (Al-Mayah et
al. 2007 and El Refai et al. 2006). A total of 14 specimens were tested under direct tensile
fatigue tests. The test matrix and the stresses applied on the specimens are shown in Table
1. Two minimum stress levels representing 40% and 47% of the ultimate strength of the
tendon were considered. Stress ranges varied between 7% and 17% of the ultimate strength
of the tendon. Proof tests that represented extreme loading conditions were also considered.

Fig. 1. CFRP tendon-anchor specimen

Test No.
Proof tests
1
2
3
Group (1)
4
5
6
7
8
Group (2)
9
10
11
12
13
14

Table 1. Fatigue test matrix of the tendon-anchor assemblies


Min. stress, MPa
Max stress, MPa Stress range, MPa Fatigue life,
(% of ultimate)
(% of ultimate)
(% of ultimate) No. of cycles
1297 (60)
865 (40)
720 (33)

1427 (66)
1729 (80)
1873 (87)

130 (6)
865 (40)
1153 (53)

>500,000
181
20

865 (40)
865 (40)
865 (40)
865 (40)
865 (40)

1081 (50)
1124 (52)
1153 (53)
1189 (55)
1225 (57)

216 (10)
259 (12)
288 (13)
324 (15)
360 (17)

>2,000,000
378,000
250,000
48,000
10,000

1009 (47)
1009 (47)
1009 (47)
1009 (47)
1009 (47)
1009 (47)

1153 (53)
1225 (57)
1268 (59)
1297 (60)
1333 (62)
1369 (63)

144 (7)
216 (10)
259 (12)
288 (13)
324 (15)
360 (17)

>2,000,000
>2,000,000
363,300
572,700
18,400
9,100

The specimens were instrumented with 5 mm strain gauges placed longitudinally at the mid
point between the two anchored ends. The specimens were then tested vertically in a MTS
loading frame having a capacity of 250 kN. The applied load was measured by a load cell
attached to the hydraulic actuator of the system. Load and strain readings were recorded at
intervals of approximately 10-20 min. over the duration of the test depending on the life of
the specimen. The tested specimens were first loaded to the required mean load and cycled
under constant load amplitude at a frequency of 3-4 Hz until failure. Specimens exceeding
2 million cycles were considered as run-out specimens and were intentionally halted.

215

2.2 Test Observations and Fatigue Life


All specimens showed a perfectly linear-elastic stress-strain relationship until failure, as
shown in Figure 2. Fatigue loading had no discernable effect on the modulus of elasticity of
the CFRP tendons, as indicated by the constant slopes of the stress-strain curves obtained
for all the specimens. A maximum decrease of 4% in the modulus of elasticity of the
tendon was observed. Perfect bond at the anchor ensured negligible slippage of the tendons
at their anchored ends.

Fig. 2. Typical stress versus strain response for specimen No. 12 (stress range 288 MPa,
fatigue life = 572,700 cycles)
All specimens failed due to the fracture of the CFRP tendon within the tendon free-length.
Catastrophic failure took place in a broom-shape manner. Table 1 shows the number of
cycles survived in all tests. The minimum load of the fatigue load history applied had little
effect on the life of the specimens. On the contrary, the fatigue life dropped significantly as
the applied stress range increased. The tendon-anchor assembly exceeded the set limiting
life (2 million cycles) when the applied stress range was less than or equal to 216 MPa
(10% of the ultimate capacity of the tendon), regardless of the minimum stress level applied
(tests No. 4, 9, and 10). Hence, this stress range could be safely considered as the fatigue
endurance limit of the assembly. The lowest stress range that caused fatigue failure of the
assembly was 259 MPa or 12% of the ultimate capacity of the tendon (tests 5 and 11).
The proof test specimens that represented extreme loading conditions also failed due to the
breakage of the tendon rather than failure of the anchor. Test specimen No. 1 withstood
more than 500,000 cycles at a stress range of 130 MPa (6% of the tendon tensile strength)
and the test was stopped intentionally before the specimen failed. In the second test, the
specimen survived 181 cycles at a stress range of 865 MPa (40% of the tendon strength)
exceeding the Post Tensioning Institute (PTI) requirements for steel tendons (50 cycles).
The third specimen survived 20 cycles when subjected to a high stress range of 1153 MPa
(53% of the tendon strength). These results indicated the excellent fatigue performance of
216

both the CFRP tendon material and the tendon-anchor assembly under severe loading
conditions.
A best-fit relationship between the fatigue life of the tendon-anchor assembly, N, and the
applied stress range, Sr, for different minimum stress levels is shown in Figure 3. A loglinear best-fit trend line representing the obtained data can be expressed in the following
mathematical form:
Sr = 541 20.5 log (N)

216 < Sr < 360 (MPa)

(1)

The proposed relationship could be used to predict the fatigue life of the assembly within
the studied stress range.

Fig. 3. Log-linear relationship between applied stress range, Sr, and fatigue life, N.
3. BEAM FATIGUE TESTS
3.1 Test Specimens and Test Setup
The same CFRP tendons were used to strengthen overloaded reinforced concrete beams
through external post-tensioning. Table 2 shows the test matrix of variables for the
strengthened beams. Note that these tests represent a portion of a large experimental
program detailed elsewhere (El Refai et al., 2008). All beams were 152 x 254 x 3500 mm
long, and reinforced with two 15M Grade 400 deformed steel reinforcing bars. The beams
were initially overloaded to twice their yield capacity prior to being strengthened using the
external CFRP tendons. The tendons were post-tensioned at 40% of their ultimate tensile
capacity. The tendons were harped at two points along the beam length by means of large
steel deviators of 500 mm radius of curvature placed at the same locations as the applied
point loads (third points of the span). The large radius of the deviators was selected to
minimize the stress concentration at the harped points in order to prevent significant
reduction in the tensile capacity of the tendons. Thin Teflon sheets were used as cushioning
material to minimize the friction between the tendons and the deviators. The inclined
217

portions of the draped tendons in the shear spans made an angle of approximately 6 degrees
with the horizontal axis of the beam. The tendons were anchored at both ends using the
Waterloo anchors. Four load cells were attached to the ends of the tendons in order to
monitor the variation in the tendon forces during the tests. Figure 4 shows the beam testing
setup (El Refai et al., 2008).
Table 2. Test matrix for strengthened beams
Type of
Max load Range No. of cycles
(kN)
(kN)
to failure
loading
Monotonic
Fatigue
57
42
>1,000,000
Fatigue
67
52
662,911
Fatigue
72
57
438,870
Fatigue
76
61
138,853
Fatigue
82
67
94,931
Fatigue
88
73
87,027

Specimen
OL15-M
OL15-F57
OL15-F67
OL15-F72
OL15-F76
OL15-F82
OL15-F88

2P

Spreader beam

Mode of failure
Steel yielding
Run-out
Steel rupture
Steel rupture
Steel rupture
Steel rupture
Steel rupture

Load cell

100

100

end
anchor

end
anchor

1100

deviators

CFRP tendon
LVDTs
Steel pedestal

`
3300
3500

Elevation
(All dimensions are in mms)

Fig. 4. Beam testing setup


3.2 Test Observations
The stresses in the CFRP tendons exhibited only a small increase as a result of the fatigue
loading applied on the beams. The increase in the maximum stress varied between 0.38%
and 1.37% of the ultimate strength of the tendon material. Most of this increase occurred
during the early part of the fatigue life, and is attributed to the stress redistribution
associated with fatigue-related growth of flexural cracks in the beam. The tendon stress
range was almost constant during the fatigue test until failure. This indicated that no
softening occurred in the tendon material due to the repeated loading. It also implied
perfect anchoring of the CFRP tendon as no slippage was encountered.

218

Table 2 also shows the mode of failure and the fatigue life of the beam specimens. All of
the beams failed due to the fatigue rupture of the reinforcing steel bars. More details about
the fatigue performance of the strengthened beams are explained elsewhere (El Refai et al.
2008). None of the CFRP tendons showed signs of degradation at the deviated points
except for a minor wearing at the surface of the tendon due to friction occurring between
the tendon and the Teflon sheets used. The same tendons were reused multiple times during
the beam fatigue tests, which means that the tendons eventually survived well over a
million fatigue cycles. The design of the post-tensioned beams was based on the results
obtained from the tendon-anchor tests in order to prevent the premature failure at the
anchored and the deviated zones. Despite the large deformation of the severely overloaded
beams prior to strengthening, the CFRP post-tensioned tendons were able to restore the
deflection and maintain the capacity of the beams.
4. CONCLUSIONS
The results of the fatigue tests conducted on the CFRP tendon-anchor assemblies and the
overloaded beams strengthened with external post-tensioned CFRP tendons were presented.
The following conclusions were drawn from this study:

Cyclic loading had a minimal effect on the mechanical properties of the CFRP tendon
material with a maximum decrease of about 4% in the modulus of elasticity.
The fatigue life of the tendon-anchor assembly decreased as the applied stress range
increased regardless of the minimum stress applied.
The fatigue limit of the CFRP tendon-anchor assembly may be taken as a stress range
of 10% of its ultimate capacity (216 MPa). Below this stress range, the assembly is
expected to withstand an infinite fatigue life.
The CFRP tendons had an excellent capacity to restore the rigidity of the overloaded
beams showing no evidence of wear or stress concentration at the deviated points.
CFRP tendons could be an excellent alternative to the conventional steel tendons in
external post-tensioning applications.

5. REFERENCES
1.
2.
3.
4.

Sayed-Ahmed, E. and Shrive, N. 1998. A New Steel Anchorage System for PostTensioned Applications Using Carbon Fibre Reinforced Plastic Tendons. Canadian
Journal of Civil Engineering, 25(1): 113-127.
Taha, M.R. and Shrive, N.G. 2003. New Concrete Anchors for Carbon Fibre
Reinforced Polymer Post-Tensioning Tendons - Part 2: Development/Experimental
Investigation. ACI Structural Journal, 100(1): 96-104.
Post-Tensioning Institute. 1997. Post-Tensioning Manual: 5th Edition. Post-Tensioning
Institute, Phoenix, AZ, USA.
Grace, N.F. and Abdel-Sayed, G. 1998. Behavior of Externally Draped CFRP Tendons
in Prestressed Concrete Bridges. Journal of Prestressed Concrete International, 43(5):
88-101.

219

5.
6.
7.
8.
9.

Dolan, C., Hamilton, R., Bakis, C., and Nanni, A. 2001. Design Recommendations for
Concrete Structures Prestressed with FRP Tendons. Draft Final Report, University of
Wyoming, Report DTFH61-96-C-00019, Laramie Wyoming, USA.
Quayle, T. 2005. Tensile-Flexural Behavior of Carbon Fibre Reinforced Polymer
(CFRP) Prestressing Tendons Subjected to Harped Profiles. M.Sc. Thesis, University
of Waterloo, ON, Canada.
Al-Mayah, A., Soudki, K., Plumtree, A. 2007. Novel Anchor System for CFRP Rod:
Finite-Element and Mathematical Models. Journal of Composites for Construction,
11(5): 469-476.
El Refai, A., West, J., and Soudki, K. 2006. Performance of CFRP Tendon-Anchor
Assembly under Fatigue Loading. Journal of Composite Structures, 80(3): 352-360.
El Refai, A., West, J., and Soudki, K. 2008. Effect of Overloading on Fatigue
Performance of Reinforced Concrete Beams Strengthened with Externally PostTensioned CFRP Tendons. Canadian Journal of Civil Engineering, 35(11): 1294-1307.

220

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

TRANSVERSE SHEAR OF GFRP RODS: TEST METHOD


DEVELOPMENT AND POTENTIAL FOR DURABILITY
ASSESSMENT
T.R. Gentry
Associate Professor, School of Architecture, Georgia Institute of Technology

ABSTRACT
Recent efforts have led to the standardization of test methods for FRP materials for
reinforcing and repairing concrete. These efforts are transitioning standards, originally
published by the American Concrete Institute (Committee 440K), as ASTM standards
under ASTM Committee D30, Composite Materials. One of the methods under
development involves the transverse shearing of FRP rods. FRP rods come under shear
loading when crossing transverse cracks, such as when they are used as dowels in concrete
pavements or as shear reinforcements in cracked wood timbers.
The paper reviews the draft ASTM test method, which has been developed using a modular
fixture to allow for shearing of smooth and deformed bars with a diameter of 10 mm (#3) to
32 mm (#10). The paper presents test results on smooth pultruded rods and commercial
GFRP rebar produced in North America. The paper details test parameters and variables
that impact the outcomes of the tests. The paper also explores the potential for use of the
test for durability screening. Another D30/440 test method, also under development,
involves the alkaline conditioning and subsequent tension testing of FRP rods used for
concrete reinforcement. In this paper, we explore the use of a combination of the alkaline
test method conditioning, with the transverse shear method for tracking mechanical
property changes.
1.

INTRODUCTION

Fiber reinforced polymer (FRP) composites for concrete reinforcement have been under
development since the late 1980s. Early research evaluated the material properties of these
first FRP reinforcements, identified the most promising applications, and demonstrated that
FRP concrete reinforcements were both structurally viable and sufficiently cost-effective to
justify commercial production. Later research and standardization activities led to the
publication of internationally-recognized codes and standards for FRP materials and
products, as well as design guidelines and codes for application of these products. From
the authors perspective, these 25 years of research and standards development has

221

followed a timeline highlighted by the following activities and milestones in rough


chronological order:
1. Materials and Applications Research
2. Demonstration Projects
3. State of the Art Reports
4. Design Guides
5. Building Code Requirements
6. Material and Product Specifications
7. Construction Specifications
8. Test Methods
Each of these eight activities are still in various states of completeness and work is ongoing,
as new materials and products emerge from industry, and new information regarding FRP
concrete reinforcements emerges from the research community. As of 2010, the market for
FRP reinforcements for concrete has matured, and is split between internal reinforcements,
used primarily for new construction in corrosive-prone environments or in other specialty
applications requiring non-metallic reinforcements, and externally bonded reinforcements,
used primarily for repair or strengthening of concrete or other structural materials.
This paper focuses on the last activity in the timeline: the development of test methods.
Standardized test methods are developed late in a product development and commercialization process as the materials, geometries, and application methods for the given
product must be established before the test methods can be finalized. In addition, test
method development is equally the business of the engineers and manufacturers, and thus
requires agreement from both parties during the development of consensus standards.
1.1 Test Method Development in the United States
In the United States, most authoritative work regarding FRP concrete reinforcements has
come from American Concrete Institute (ACI) Committee 440: Fiber-Reinforced Polymer
Reinforcement. The development of test methods for FRP concrete reinforcements began
within Committee 440 in the late 1990s. Early work on test methods for FRP
reinforcements can be found in a papers by Nanni et al [1] and Benmokrane et al [2]. The
first set of guide test methods for FRP concrete reinforcements was published by ACI
Committee 440 in 2004 [3]. Some of these methods were adapted from schematic test
method proposals developed earlier by the Japanese Society of Civil Engineers [4].
In the United States, test method development has transitioned from ACI to ASTM, per an
agreement between the two organizations. This activity has resulted in the publication of
four ASTM standards encompassing the information from seven sections in the original
440R.3R (Table 1). The publication of two additional ASTM standards is anticipated in
2011, with one standard still in the balloting in review process. The original ACI guide
methods have been strengthened through the ASTM balloting and revision process, with
additional laboratory testing and test fixture development taking place for many of the
standards. An example of this process is detailed in a recent paper by Eveslage et al.
detailing the research undertaken to assess and improve the test method for laminate pulloff from a concrete substrate, now published as ASTM D7522 [5].
222

Table 1. ASTM D30 and ACI 440 Test Methods


ASTM
No.
D7205

ACI 440.3R
Methods
B.1, B.2 and
Appendix A
B.8

Title

Standard Test Method for Tensile Properties of Fiber


Reinforced Polymer Matrix Composite Bars
D7337
Standard Test Method for Tensile Creep Rupture of
Fiber Reinforced Polymer Matrix Composite Bars
D7522
Standard Test Method for Pull-Off Strength for FRP
L.1
Bonded to Concrete Substrate
D7565
Standard Test Method for Determining Tensile
L.2 and
Properties of Fiber Reinforced Polymer Matrix
Appendix B
Composites Used for Strengthening of Civil
Structures
L.3
WK22346
Method For Determining Apparent Overlap Splice
(1)
Shear Strength Properties Of Wet Lay-Up FiberReinforced Polymer Matrix Composites Used For
Strengthening Civil Structures
WK22348
Standard Test Method for Transverse Shear Strength B.4
of Fiber Reinforced Polymer Matrix Composite Bars
WK27200
Standard Test Method for Alkali Resistance of Fiber B.6
Reinforced Polymer (FRP) Matrix Composite Bars
used in Concrete Construction
(1) WK identifies a work item a draft test method in the process of being balloted.
(2) Test method has passed ASTM balloting and will be published in 2011.
(3) Test method is undergoing ASTM ballot process, publication date unknown.

Year
2006
2007
2007
2009

2011 (2)

2011 (2)
(3)

2. TRANSVERSE SHEAR OF FRP BARS


The remainder of this paper focuses primarily on the development of the transverse shear
test method, a brief description of typical test results on smooth rods and production GFRP
rebar, and its applicability for durability testing of GFRP bars. FRP bars are loaded in
transverse shear as a consequence of dowel action in reinforced concrete beams. They may
also be inserted as dowels at joints in concrete pavements, or be used to reinforced concrete
or other materials in direct shear. Papers by Amy and Svecova (shear strengthening of
timber) [6], Dulude et al. (punching shear strength of GFRP-reinforced concrete) [7], Eddie
et al (dowels in concrete pavements) [8] and Gentry (shear strengthening of gluedlaminated timber) [9] discuss applications in which the FRP bar are loaded in transverse
shear.
This paper introduces the transverse shear test as developed for the forthcoming ASTM test
method and summarizes details of the test fixture. The paper discusses details of running
and controlling the test, and test conditions that affect the outcomes of the test. Finally the
paper discusses the potential for the transverse shear test as a substitution for the more
complex tension test currently used when assessing the alkaline resistance of GFRP bars.

223

2.1 Summary of the Transverse Shear Test


The transverse shear test fixture cradles a 225 mm length of bar, fully supported along its
length except for a 25 mm gap centered along the length of the bar. The bar is pressed
down onto the supporting seat with set screws to hold it in place. A steel blade, 25 mm
thick and machined to fit snugly around the perimeter of the bar, is pressed on the bar with
a universal testing machine, causing the middle 25 mm section to be sheared from the bar.
Two lower blades, with the same diameter slots as the upper blade, support the bar adjacent
to the shearing planes. Two shear planes are formed during the test so that the bar fails in
double shear.
The transverse shear test fixture is shown in Fig. 1(A), along with a section of 9.5 mm
diameter smooth rod that has been sheared in the fixture. The upper blade is shown sitting
loose at the upper left of the fixture. In Fig. 1(B), the lower assembly is shown
disassembled. The fixture comes apart to allow for swapping of the lower blades, which are
specific to the diameter of the bar being sheared.
The standard calls for the test to be completed by inserting the assembled fixture into a
universal compression machine, and compressing the upper blade with the machine in
displacement control. According to the draft standard, the displacement rate should be
selected so that the test article fails at a time between 1 and 10 minutes.

Upper
blade

Lower
blades
(A)

(B)

Barseat

Fig. 1. Transverse shear fixture: (A) fixture assembled with 9.5 mm rod, sheared; (B) low
portion of fixture disassembled showing two bar seats, two lower blades, two guides, and
tie rods. The upper blade and straps are not shown in (B).
2.2 Typical Test Results with GFRP Bars
The primary test result is the transverse shear stress, calculated as the one-half of the peak
failure load (to account for double shear) divided by the cross-sectional area of the
specimen. Two typical stress-displacement diagrams are show in Fig. 2. In Fig. 2(A), the
diagram exhibits the behavior typically observed when two shearing faces form
simultaneously. In Fig. 2(B), the observed behavior demonstrated when one face forms and
begins to shear, and the second face forms somewhat later. Note that the slope of the
diagram is not consistent, and that no useful stiffness or modulus value is provided by the
test (see notation (1) in Fig. 2).

224

160

(twodatasetsshown)
(3)

(2)

140

(3)

(2)

Stress(MPa)

120
100

(1)

80

(A)

60

(B)

40
20
0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

Displacement(mm)

Fig. 2. Example stress-displacement diagram from transverse shear tests; two tests shown.
In test (A) both failure surfaces form simultaneously, in test (B) the second failure surface
forms after the first failure surface.
A series of tests on smooth 9.5 mm diameter pultruded rods, constructed of unidirectional
E-glass fiber (60% volume fraction) and an unsaturated polyester resin, were completed to
explore the range of test parameters. Two key parameters were identified in the testing: (1)
the speed of testing and (2) the tightness or fitment of the cutting blades relative to the
perimeter of the test article.
Table 2 shows test results for 4 sets of 5 specimens tested at different displacement rates,
spanning the range of times allowed by the draft standard. The baseline set was tested at a
displacement rate of 0.91 mm/min which led to an average failure time of 2.6 minutes. The
coefficients of variation in the test are quite low, especially for a test on FRP composite
materials. The tests at slower displacement rates failed at significantly lower force levels. It
is likely that the transverse shear strength is a matrix dominated property, especially in
uniaxial FRP rods, and that tests at low displacement rates are allowing the matrix to creep,
and thus bringing an element of creep rupture into the test. It is therefore recommended that
the tests take place over a shorter time period, with failure occurring between 1 and 3
minutes.
Table 2. Transverse Shear Strength of 9.5 mm GFRP Rods versus Displacement Rate
Test Speed
Very Slow
Slow
Baseline
Fast

rate
(mm /min.)
0.15
0.31
0.91
1.84

Time
to
Failure (min.)
10.3
5.0
2.6
1.2

Shear Strength
(mPa)
137.0
141.9
162.0
167.9

COV
1.82%
1.22%
1.99%
0.90%

%
Change
from Baseline
-15.43%
-12.43%
0.00%
3.64%

Table 3 shows the result of the same rods tested with 3 different sets of blades, representing
a tight, close and loose fit of the rod relative to the machined width of the slot (all tests run

225

at a displacement rate of 0.91 mm / min.). The results show that the apparent shear
strengths are highly sensitive to the width of the slot. The highest transverse shear strength
is recorded for rods with tight fitting blades. This is a significant finding as the method is
also used with textured FRP bars used as concrete reinforcements, and the ability to fit
these bars into a tight set of blades will be made problematic by the texture on the outside
of the bars (see text that follows).
Table 3. Transverse Shear Strength of 9.5 mm GFRP Rods versus Slot Width
Fitment
Tight
Close
Loose

Rod
Diameter
(mm)
9.53
9.53
9.53

Slot
Diameter
(mm)
9.91
11.40
14.76

Slot to
Rod Ratio

Transverse Shear
Strength (MPa)

COV

1.04
1.20
1.55

162.0
120.3
97.1

1.99%
1.44%
4.14%

% Change
from
Baseline
100.0%
-25.8%
-40.1%

3. TRANSVERSE SHEAR TESTING OF GFRP REBAR


The major suppliers of GFRP reinforcing bars in North America provided bars for testing in
the prototype fixture. Bar A has a helical wrap with a moderate amount of abrasive
surface material to promote bond. Bar B is highly textured, with a significant amount of
larger-size abrasive material. The transverse shear strength of the three smallest bar sizes
are given in Table 4 below. The bars were tested at a displacement rate leading to shear
failures at approximately three minutes into the loading regime (see Table 2). The cutting
blades were the same for the both the A and B bars, and were machined to fit as tightly
as possible on the bars, without necessitating the removal of any of the surface texture on
the bars.
The transverse shear strength is calculated using the so-called standard cross-sectional
area for U.S. reinforcing bars. The apparent shear strengths as calculated are significantly
higher for the GFRP rebar than the 9.5 mm smooth rod, even though the internal construction of the smooth rods and GFRP bar is similar. This is largely because the GFRP
rebar have larger actual cross-sectional areas than the standard area, and have higher
transverse strength values when calculated on a standard area basis.
Table 4. Transverse Shear Strength of Commercially Produced GFRP Bars
Bar Size

Standard CrossSectional Area


(mm2)

#3

Bar Type A

71.0

Transverse Shear
Strength (MPa)
212.5

#4

129.0

#5

200.0

Bar Type B

2.62%

Transverse Shear
Strength (MPa)
235.3

4.74%

209.7

1.36%

192.5

1.99%

199.2

2.07%

188.3

4.44%

COV

4. DURABILITY TESTING USING TRANSVERSE SHEAR TEST

226

COV

A common screening test for durability of GFRP reinforcing bars uses an aqueous alkaline
environment at elevated temperature and ASTM D7205 [10] to measure the change in
tensile strength as a function of environmental conditioning. The aqueous environment is
intended to match the inherent porewater chemistry of Portland cement concrete [11].
Much of the work on alkaline durability of GFRP bars has been reported by Benmokrane
and colleagues [12-14].
A short experiment was completed to determine whether the transverse shear test could be a
substitute for the more complex tensile test, when considering the alkaline conditioning
environment. The tension test is somewhat problematic, due to the requirement for affixing
anchors to the ends of the bars, and the need to isolate the anchors from the conditioning
environment. In this demonstration, short conditioning times were used as the sample rods
were pultruded using an unsaturated polyester resin, which provides significantly lower
protection to the glass fibers, than the vinylester resins typically called for in GFRP rebar
[15]. As a result, the polyester-glass rods degrade quickly. The rods were conditioned at
60C in an an alkaline solution containing calcium, sodium, and potassium oxides, held at a
pH of 12.8 0.2 per the draft ASTM standard on Alkali Resistance.
The results in Table 5 show that the degradation trends match those observed by
Benmokrane and others. Whether the degradation rates for the transverse shear strength
match those for the tensile strength will be the subject of a forthcoming paper by the author.
Table 5. Transverse Shear Strength of 9.5 mm GFRP Rods under Alkaline Conditions
Time
baseline
3 days
7 days
14 days

Transverse Shear
Strength (mPa)
162.0
163.3
130.3
120.6

COV

% Change from Baseline

1.99%
1.05%
5.89%
3.81%

0.0%
0.8%
-19.6%
-25.6%

5. SUMMARY AND CONCLUSIONS


The transverse shear test is a significant addition to the body of test methods and
specifications for FRP reinforcing bars. The transverse shear strength of smooth rods and
textured rebar are easy to establish using the proposed test fixture, and the test results show
little variability when the test parameters associated with testing speed and cutting blade to
bar diameter are held constant. The blades should be manufactured so that they fit as tightly
as possible onto the rod or bar otherwise, a downward bias in measured transverse shear
strength can be expected. Initial test results show that the method could be an inexpensive
means for testing degradation of bars in alkaline environments, eliminating the need for
complex fixturing of tension specimens and the isolation of tension anchors from
environmental conditions. Future research will be needed to correlate the strength loss in
transverse shear to the strength loss in tension in such environmental screening tests.
6. REFERENCES

227

1.
2.

3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.

Nanni, A., Bakis, C.E., and Boothby, T.E. 1995. Test Methods for FRP-Concrete
Systems Subjected to Mechanical Loads: State of the Art Review. Journal of
Reinforced Plastics and Composites, 14(6): 524-588.
Benmokrane, B., Wang, P., Gentry, T.R., and Faza, S. 2001. Tests methods to
determine properties of FRP rods for concrete structures. Proceedings of the
International Workshop on Composites in Construction, American Society of Civil
Engineers, Capri, Italy, 20-21 July.
ACI Committee. 2004. Guide Test Methods for Fiber-Reinforced Polymers (FRPs) for
Reinforcing or Strengthening Concrete Structures. ACI 440.3R-04, American Concrete
Institute: Detroit, Michigan, 40 p.
JSCE. 1997. Recommendation for Design and Construction of Concrete Structures
Using Continuous Fiber Reinforcing Materials. Japanese Society of Civil Engineers,
Concrete Engineering Series 23, ed. Machida, A., Tokyo, Japan, 325 p.
Eveslage, T., Aidoo, J., Harries, K.A., and Bro, W. 2010. Effect of variations in
practice of ASTM D7522 standard pull-off test for FRP-Concrete interfaces. Journal of
Testing and Evaluation, 38(4), 7 p.
Amy, K. and Svecova, D. 2004. Strengthening of dapped timber beams using glass
fibre reinforced polymer bars. Canadian Journal of Civil Engineering, 31(6): 943-955.
Dulude, C., Ahmed, E., and Benmokrane, B. 2010. Punching shear strength of concrete
flat slabs reinforced with GFRP bars. Proceedings of the 2nd International Structures
Specialty Conference, CSCE, Winnipeg, Manitoba.
Eddie, D., Shalaby, A., and Rizkalla, S. 2001. Glass fiber-reinforced polymer dowels
for concrete pavements. ACI Structural Journal, 98(2): 201-206.
Gentry, T.R. 2011. Performance of Glued-Laminated Timbers with FRP Shear and
Flexural Reinforcement. Journal of Composites for Construction. (to appear).
ASTM 2006. Standard Test Method for Tensile Properties of Fiber Reinforced
Polymer Matrix Composite Bars. ASTM D7205, American Society of Testing and
Materials, West Conshohocken, PA, USA.
Gentry, T.R. 2001. Life Assessment of Glass-Fiber Reinforced Composites in Portland
Cement Concrete. Proceedings of the 16th Annual Meeting of the American Society of
Composites, Blacksburg, Virginia, 10-12 September.
Benmokrane, B., Wang, P., Ton-That, T.M., Rahman, H., and Robert, J.F. 2002.
Durability of glass fiber-reinforced polymer reinforcing bars in concrete environment.
Journal of Composites for Construction, 6(3): 143-153.
Debaiky, A.S., Nkurunziza, G., Benmokrane, B., and Cousin, P. 2006. Residual tensile
properties of GFRP reinforcing bars after loading in severe environments. Journal of
Composites for Construction, 10(5): 370-380.
Robert, M., Cousin, P., and Benmokrane, B. 2009. Durability of gfrp reinforcing bars
embedded in moist concrete. Journal of Composites for Construction, 13(2): 66-73.
Gentry, T.R., Bank, L.C., Barkatt, A., and Prian, L. 1998. Accelerated test methods to
determine the long-term behavior of composite highway structures subject to
environmental loading. Journal of Composites Technology and Research, 20(1): 38-50.

228

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

FATIGUE LIFE BEHAVIOR OF ADHESIVELY BONDED


PULTRUDED FRP JOINTS UNDER TENSILE AND COMPRESSIVE
LOADING
R. Sarfaraz1, A.P. Vassilopoulos2 and T. Keller3
1

PhD student, Composite Construction Laboratory (CCLab), EPFL, Switzerland


Research and Teaching Associate, Composite Construction Laboratory (CCLab), EPFL,
Switzerland
3
Professor and Director, Composite Construction Laboratory (CCLab), EPFL, Switzerland
2

ABSTRACT
The fatigue response of adhesively-bonded pultruded GFRP double-lap joints has been
investigated under different loading patterns and the results are presented in this paper.
Experimental results were obtained for three different applied load ratios, corresponding to
tensile, R=0.1, compressive, R=10, and reversed (combined tensile-compressive, R=-1)
fatigue loading. The dominant failure mode for the tensile and the reversed loading was a
fiber-tear failure that occurred in the mat layers of the pultruded laminates, while the failure
was driven by a crack in the roving layer in the case of R=10. The fatigue life of the
examined joint configuration was found to be affected by the applied loading type, being in
general shorter for tensile loads and longer when compressive loads were applied.
1. INTRODUCTION
Adhesively-bonded FRP joints represent a critical element in numerous engineering
structures nowadays, and they must therefore be able to transfer the developed stresses (of a
complex nature) from one part of the structure to another. One of the key objectives of the
scientific research community is thus the development of reliable methodologies for the
fatigue life prediction of adhesively-bonded FRP joints under realistic loading patterns, and
for this purpose the fatigue behavior of the joints must be extensively examined.
The adhesively-bonded FRP joints have been used in civil engineering applications and
their quasi-static and constant amplitude fatigue behavior have been investigated, e.g. [1-5].
Full-scale structural adhesively-bonded joints composed of pultruded GFRP laminates and
epoxy adhesive were investigated under tensile constant amplitude fatigue loading, e.g. [2,
5-7].

229

Despite the vast number of scientific publications concerning the fatigue of FRP composite
joints under constant amplitude loading, numerous aspects related to their behavior under
realistic loading patterns require further examination. One of these aspects is the effect of
the mean stress and the application of compressive loading components on the life of the
examined joints. All experimental studies performed to date are based on tensile fatigue
loads mainly because they have focused on joints with metal substrates, e.g. [8], in which a
cohesive or an adhesive failure is exhibited. A significant mean stress effect was reported in
[8] and a linear Goodman diagram has been shown to be appropriate for representing this
effect. However, there is no evidence that the fatigue behavior of the examined joints is the
same under compressive loads, a loading pattern that is very common during the operation
of a structure. This phenomenon can be more pronounced for pultruded FRP joints in which
cracks in the adherends which are generally not unidirectional lead the failure process,
see [2-7].
The fatigue behavior of adhesively-bonded pultruded FRP Double-Lap Joints (DLJs) under
different constant amplitude loading patterns, including tensile, compressive and reversed
(combination of tensile and compressive) loading has been experimentally investigated in
this work. The failure process of the examined joints was examined and analyzed. The
fatigue life was simulated using load-life curves (similar to the S-N curves used for
composite laminates) and a constant life diagram formulation was employed in order to
model the effect of the load ratio on the fatigue life.
2. EXPERIMENTAL PROGRAM
2.1. Material
Symmetric adhesively-bonded double-lap joints composed of pultruded GFRP laminates
bonded using an epoxy adhesive system were examined. The pultruded GFRP laminates,
supplied by Fiberline A/S, Denmark, consisted of E-glass fibers and isophthalic polyester
resin. The fiber architecture of the laminates is shown in Fig. 1. The laminate comprises
two mat layers on each side and a roving layer in the middle, with a thin layer of polyester
veil on the outer surfaces of the laminate. Each mat layer comprises a woven fabric stitched
to a chopped strand mat (CSM).

Fig. 1. Microscopic view of laminates (cross section perpendicular to pultrusion direction)

230

The longitudinal strength and Youngs modulus of the GFRP laminate were obtained from
tensile experiments, according to ASTM D3039-08, as being 307.54.7 MPa and 25.10.5
GPa respectively. A two-component epoxy adhesive system was used (Sikadur 330, Sika
AG Switzerland) as the bonding material. The tensile strength of the adhesive was 38.12.1
MPa and the stiffness 4.60.1 GPa [9].
2.2. Specimen Fabrication and Experimental Set-Up
Two different joint configurations were prepared; one with a total length of 410 mm, see
Fig. 2, and used only for tensile loading, and another with a reduced total length of 350
mm, which was used when compressive loads were applied to avoid any buckling of the
joints. To achieve the latter configuration, the free length of the inner laminate was
reduced from 100 mm to 40 mm without changing the bonding and gripping length. These
dimensions were selected after preliminary testing and modal analysis using the finite
element software ANSYS, v.10, which indicated that this length reduction sufficed to
prevent the buckling of the laminates. Moreover, the finite element stress analysis showed
that there is no change in the stress field close to the bonded area due to the decreased
laminate length. The bond line was kept constant at 50 mm for both joint configurations.
The gripping areas were also 50-mm long to allow load transfer through shear.
An aluminum frame was employed to assist the alignment of the laminates. All surfaces
subjected to bonding were mechanically abraded in advance (to a depth of approximately
0.3 mm) in order to increase roughness, and then chemically degreased using acetone. The
thickness of the adhesive was controlled by using 2-mm thick spacers embedded in the
bonding area. The specimens were kept for at least 10 hours at 30 C to ensure full curing
of the adhesive.
Aluminum tabs were used to avoid crushing of the laminate by the wedges of the testing
frame. The gripping part (shown on the right side of the specimen in Fig. 2), which is
supported by a bolted connection, was designed to adapt the thickness of the specimen to
the opening of the jaw faces of the machine. No failure or crack initiation was observed in
the gripping part of all specimens during the entire experimental program.
Aluminum tabs

Inner GFRP laminate

Epoxy adhesive

Outer GFRP laminates

Gripping part
10mm

6mm
2mm
6mm
50mm

100mm

100mm

50mm

50mm

50mm

410mm

Fig. 2. DLJ geometry


An INSTRON 8800 servohydraulic machine operating under laboratory conditions
(235C and 5010% RH) was used for all experiments. Quasi-static tensile and
compressive experiments were performed under displacement-control mode with a ramp

231

rate of 1 mm/min. Five specimens under tensile loading and three under compression were
examined in order to define the stiffness and the strength of the joints under investigation.
Fatigue experiments were performed under load control, using a constant amplitude
sinusoidal waveform, at a frequency of 10 Hz. It has already been shown [6] that the fatigue
performance of similar specimens is not affected by the frequency when it lies in the range
between 2 and 10 Hz. Three load ratios were selected to simulate tension-tension (R=0.1),
reversed (R=-1) and compression-compression (R=10) loading. The fatigue experimental
program was designed to derive experimental data that cover the entire lifetime between
one cycle and five million cycles. Seven load levels for R=0.1 and five load levels for R=-1
and R=10 were selected. At least three specimens were tested at each load level in order to
also obtain information regarding the scatter of the fatigue life.
3. EXPERIMENTAL RESULTS AND DISCUSSION
3.1. Quasi-Static Investigation
The examined DLJs showed an almost linear load-elongation behavior up to a brittle failure
under both tension and compression. Similar behavior was reported in [9-10] for DLJs of
the same material under tension loading. The ultimate tensile load (UTL), ultimate
compressive load (UCL), and the stiffness for the examined cases are given in Table 1.
The joint stiffness was defined as the slope of the load-displacement curve in the range of 0
to 10 kN where no crack formed in the bond line. The higher joint stiffness under
compression is due to the reduced length of the inner laminate and the higher strength due
to the different mode exhibited under compressive failure.
Table 1. Quasi-static data
Tension
Compression

UTL or UCL [kN]

Stiffness [kN/mm]

25.50.97
-29.01.07

23.10.20
30.50.39

The observed failure mode was a fiber-tear failure as presented in Fig. 3a for a specimen
tested under tensile loads. A dominant crack initiated from the joint corner of one of the
bond lines the upper in this figure between the adhesive and the inner laminate and then
shifted deeper, between the first and the second mat layers of the inner laminate, and
propagated along this path up to failure. The cracks observed along the lower bond line and
at the right side of the inner laminate of the specimen shown in Fig. 3a are secondary cracks
that occurred after the failure of the specimen. The same failure mode was observed for
double-lap joints composed of similar materials as documented in [10].
A different failure mode was observed for the specimens examined under compression
loading as shown in Fig. 3b. The dominant crack initiated and propagated from the right
side of the inner laminate (as shown in Fig. 3b) inside the roving layer.

232

(a)
(b)
Fig. 3. DLJ failure mode under (a) tension loading (b) compression loading
3.2. Fatigue Investigation
3.2.1 Failure modes
Similar failure modes were observed under the respected fatigue loading conditions
irrespective of the load level. However, different failure modes were observed under
reversed (T-C) fatigue loading (R=-1). In most of the examined cases, the failure process
was similar to T-T mode, although for some of the examined specimens, in addition to the
dominant crack along one of the bond lines, a smaller crack of approximately 1 mm was
observed in the middle of the inner laminate at a similar location as for C-C loading.
However, during the fatigue life the dominant crack was propagating and leading the failure
process, while the crack created by the compressive component of the applied cyclic load
remained short up to the failure of the joint.
3.2.2 Fatigue life
The fatigue life of joints under the applied load ratios is plotted against the maximum
applied cyclic load in Fig. 4 for comparison. The lines that simulate the fatigue behavior
are based on a power law relationship of the form:
Fmax AN B

(1)

where Fmax is the maximum load, N, the number of cycles and A, B are model parameters
that can be obtained after fitting Eq. (1) to the experimental data. The values of parameters
A and B as estimated by a linear regression analysis are given in Table 2. The F-N curve for
R=-1 shows the highest slope (-0.1038) of the three curves, exhibiting the sensitivity of the
examined joints to reversed loading.

233

35
30

Max. Load [kN]

25
20
15
10
5
0
1
10

R=0.1
R=10
R=-1
Exp. data (R=0.1)
Exp. data (R=10)
Exp. data (R=-1)

10

10

10

10

10

10

Nf

Fig. 4. Comparison of load-life data at different R-ratios


Table 2: Estimated material constants of Eq. 1 for all load ratios
R = -1
R =0.1
R =10
39.77
38.49
32.45
A
-0.1038
-0.0828
-0.0426
B
25
N=1e3
N=1e4
N=1e5
N=1e6
N=1e7
R=-1
R=10
R=0.1

Load amplitude [kN]

20

15

10

0
-30

-20

-10

10

20

30

Mean load [kN]

Fig. 5. Linear Goodman diagram


The effect of the load ratio on the fatigue life of the examined joints can also be
demonstrated if the fatigue data is plotted on the mean-amplitude (Fm-Fa) plane. As under
static loading where the examined joints exhibited higher strength and stiffness under
compression, (see Table 1), their compressive fatigue strength is greater as shown in Fig. 4.
The typical linear constant life diagram derived by using the static strengths and the
experimental fatigue data under R=-1 is presented in Fig. 5. In this diagram, each set of
points represents the projection of the power law model at given fatigue lives onto the
mean-amplitude load plane. The lines represent the constant life contours, i.e. each one
corresponds to the same number of cycles. The diagram shows that the linear model is
unable to accurately model the fatigue life of the examined joint configurations at R=0.1
and 10, especially in the high cycle regime, where it seems that the selected constant life
model overestimates the fatigue life under R=0.1, while it underestimates the F-N curve for
R=10
234

4. CONCLUSIONS
The fatigue behavior of adhesively-bonded pultruded GFRP double-lap joints was
experimentally examined under the load ratios of 0.1, -1, and 10 in order to investigate the
effect of the mean load on fatigue life. The results showed that the change in load ratio
significantly affected the fatigue behavior of the examined adhesively-bonded joints. The
following conclusions were drawn:

The examined joints exhibited different behavior under quasi-static tension and
compression loading. A fiber-tear failure was observed under tensile loading, while
the failure of the specimens was dominated by cracks in the mat layers of the inner
laminate. Under compression, failure occurred in the roving layer in the middle of the
inner laminate.

Specimens loaded under quasi-static compression exhibited higher strength and


stiffness compared to those examined under tension loads.

The fatigue failure mode of DLJs under load ratios of 0.1 and -1 were found to be
similar to the failure mode observed under quasi-static tension loads, although under
reversed loading a more complicated failure process was recorded. The failure of
specimens under R=10 was similar to the compressive quasi-static failure mode.

The examined DLJs were found to be more sensitive to tensile than to compressive
fatigue loads. When fatigue data are represented on the Fmax-N plane, the curve
corresponding to R=-1 is the steepest, while the Fmax-N curve corresponding to R=10
exhibits the lowest slope.

The simple linear constant life diagram cannot accurately describe the mean load
effect on fatigue life and more sophisticated phenomenological models must be used
in order to provide better results.
5. ACKNOWLEDGEMENTS
This work has been supported by the Swiss National Science Foundation (Grant No
200020-121756), Fiberline Composites A/S, Denmark (supplier of the pultruded
laminates), and Sika AG, Zurich (adhesive supplier).
6. REFERENCES
1.
2.
3.
4.

Burgueno, R., Karbhari, V.M., Seible, F., and Kolozs, R. 2001. Experimental dynamic
characterization of an FRP composite bridge superstructure assembly. Composite
Structures, 54(4): 427-444.
Keller, T. and Tirelli, T. 2004. Fatigue behavior of adhesively connected pultruded
GFRP profiles. Composite Structures, 65(1): 5564.
Keller, T. and Grtler, H. 2005Quasi-static and fatigue performance of a cellular FRP
bridge deck adhesively bonded to steel girders. Composite Structures, 70(4): 484-496.
Keller, T. and Zhou, A. 2006. Fatigue behavior of adhesively bonded joints composed
of pultruded GFRP adherends for civil infrastructure applications. Composites Part A,
Applied Science and manufacturing, 37(8): 1119-1130.

235

5.

Zhang, Y., Vassilopoulos, A.P., and Keller, T. 2008. Stiffness degradation and life
prediction of adhesively-bonded joints for fiber-reinforced polymer composites.
International Journal of Fatigue, 30(1011): 18131820.
6. Zhang, Y., Vassilopoulos, A.P., and Keller, T. 2009. Environmental effects on fatigue
behavior of adhesively-bonded pultruded structure joints. Composite Science and
Technology, 69(78): 10221028.
7. Zhang, Y., Vassilopoulos, A.P., and Keller, T. 2010. Fracture of adhesively-bonded
pultruded GFRP joints under constant amplitude fatigue loading. International Journal
of Fatigue, 32(7): 979987.
8. Crocombe, A.D. and Richardson, G. 1999. Assessing stress and mean load effects on
the fatigue response of adhesively bonded joints. International Journal of Adhesions
Adhesives, 19(1): 19-27.
9. De Castro, J., Keller, T. 2008. Ductile double-lap joints from brittle GFRP laminates
and ductile adhesives. Part I: experimental investigation. Composites Part B
Engineering, 39(2): 271281.
10. Zhang, Y. and Keller, T. 2008. Progressive failure process of adhesively bonded joints
composed of pultruded GFRP. Composites Science and Technology, 68(2): 461470.

236

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

FATIGUE PERFORMNACE OF COMPOSITE SANDWICH PANELS


WITH AND WITHOUT INTERNAL RIBS
H. Mathieson1 and A. Fam2
1

Doctoral Student, Department of Civil Engineering, Queen's University, Kingston, Canada


Professor and Canada Research Chair in Innovative and Retrofitted Structures, Department of
Civil Engineering, Queen's University, Kingston, Canada

ABSTRACT
Structural applications of sandwich panels composed of lightweight polyurethane core and
GFRP skins have a strong potential in civil engineering. One promising application is the
use of panels as wall cladding of buildings. For this application, the effects of wind
pressure and suction need to be investigated and understood. Another potential application
is for light-weight decking of pedestrian bridges. Due to the cyclic loading nature of wind
on buildings and traffic on bridges, it is imperative to account for repetition of loading and
gain insight to the fatigue behavior of the sandwich panels. The sandwich panel sections
may or may not include ribs connecting the two skins to increase shear rigidity. In this
study, large sandwich panels that had comparable ribbed and non-ribbed cross-sections
were tested under cyclic loading to establish and compare their fatigue lives. It was shown
that to achieve about two million cycles, load should be limited to 20 and 30% of the
ultimate monotonic strength, for panels with and without ribs, respectively. While the ribs
added about 84% to the monotonic ultimate strength, their contribution dropped to less than
5% after 10,000 cycles.
1. INTRODUCTION
Various structural systems involving Glass Fibre Reinforced Polymer (GFRP)
reinforcements have been developed. Benefits include increased environmental resistance
and increased strength-to-weight ratio (1). Sandwich panels are one form of these
composite structures designed with stiff exterior sheets (skins) that overlay a soft core
material. The premise is that spacing of the skins provides sufficient flexural resistance
while the function of the core is to maintain transverse shear transfer between these two
skin (2) (Figure 1a). Additional benefits include allowing sufficient thermal insulation,
while allowing non-structural components and ducts to be passed through (i.e. electrical
wiring if used as part of a building envelope). All flexural resistance of the rather soft foam
core is considered insignificant and typically ignored. In some applications, if additional

237

rigidity is required, orthogonal FRP plates are provided as ribs connecting the two skins
together, to increase shear capacity (Figure 1b).
b)

a)

Fig. 1. a) Control Sandwich Panel Cross-Section, and b) Cross-Section where an Internal


Rib is Used
Fatigue is a specific response all materials experience, where microscopic failure initiates
and propagates, leading to reduction in ultimate strength (3). The fatigue property, defined
as the loading ratio (R), represents the minimum applied load divided by the maximum
applied load. In nearly all cases, a 100% reverse loading case (i.e. an R = -1) results in the
lowest number of cycles for a given amplitude. This is when the minimum applied load is
equal to the maximum load causing opposite strain. This scenario is very important in
fatigue design as it results in the worst case scenario for fatigue loading. Generally, data is
displayed in a Wohler curve, which displays the applied load compared to the number of
cycles to cause failure at that load. This display is more commonly known as the S-N curve
when testing material coupons that report the stress rather than load versus number of
cycles, and will be referred to as such in this research, even though large sandwich panels
and not coupons are being tested in bending fatigue.
Experiments reported in literature have been performed to investigate the effect of loading
ratio on fatigue performance (4). It was shown that material characteristics of the foam core
are affected by the loading rate (5 and 6). Changing the loading rate caused a reduction in
stiffness of the foam by causing the internal temperature of the foam to rise. It was
determined that changing loading frequency from 1 Hz to 5 Hz resulted in a difference of
16oC. The resulting decrease in shear properties was 10% at 5 Hz. However, from a range
of 0.33 Hz to 0.91 Hz, it was found that no measurable change in stiffness occurs (6).
2. TEST PROGRAM
This paper provides a comparison between flexural fatigue behavior of 78 mm thick
sandwich panels with and without GFRP ribs. The panels are made of polyurethane foam
core and GFRP skins. The panels were tested under the worst case scenario of a loading
ratio R = -1. Two groups of sandwich panels were tested and compared. The first group
had no ribs while the second group had a GFRP rib running longitudinal at the mid-width
of the panel (Figure 2). The GFRP ribs were of identical material composition as the skins.
2.1 Material Properties
The foam core was Polyurethane, fabricated from Corafoam U20. This is a closed-cell
polyurethane foam with a density of 32.04 kg/m3 (2 pcf). The skins and ribs were
238

fabricated from 54 oz. 3 weave E-glass 2022 silane sized and CoPoxy 4281A resin with
Copoxy 4284 hardener with an average thickness of 1.6 mm.

b)

a)

Fig. 2. Design Sketches (Top View) of Both test Samples, a) Non-Ribbed (A Sections) and
b) One Internal Rib (B Sections)
2.2 Test Setup and Instrumentation
A specially designed steel frame was fabricated with the intent of mimicking a four-point
bending scheme of a simple one-way slab with completely reversed loading. Loads were
applied at one and two thirds of the 1145 mm long span of the panel. The nominal width of
all panels was about 635 mm. The testing machine used was an Instron 8800, that has
opposing hydraulic grips. Both of these grips can be manually raised or lowered, but during
testing only the top grip can oscillate, as the bottom grip remains stationary. The loading
part of the frame was gripped by the upper hydraulic grips of the machine while the
supporting component of the frame was gripped by the lower gripping system. Figure 3
shows the schematic and picture of the test setup. To facilitate the push-pull action of
loading, the specimen was clamped from top and bottom at both loading and support points
(black boxes in Figure 3a). The clamping system also incorporated a hinge-roller system to
allow for free rotation and sliding under loading. This allows for consistency of the loading
application while also allowing instantaneous reverse bending.

a)
b)

Fig. 3. a) Display Basics of Frame, and b) Image of Frame with Sample


3. TESTING SCHEDULE AND RESULTS
3.1 Testing
As mentioned earlier, two groups were tested, one without any rib and one with a central
longitudinal Z-shape rib. For each group, the first test was a monotonic unidirectional
bending test to establish the ultimate load capacity of the panel, as a control test.

239

Following, were four specified high-cycle fatigue tests. For the non-ribbed sections (A
samples), maximum loads of 70%, 50%, 40% and 30% of ultimate control strength were
applied. For the ribbed samples (B samples) 60%, 40%, 30% and 20% loads were applied.
All cyclic loading tests aimed for establishing the number of cycles to failure under fully
reversed loading (i.e. worst case scenario for fatigue). Therefore, a loading ratio of -1 was
used. The loading rate was maintained between 0.6 and 0.7 Hz. Video and computer
monitoring were continuous to detect failure and the corresponding number of cycles.
3.2 Failure Modes
Shear failure of the soft foam core occurred in the high shear region of all five A samples,
by means of diagonal tensile stresses (Figure 4a). Upon investigation after failure, the
propagation could be observed to advance within the foam, fairly close to the bond line
between the skin and the foam. The advancement of these cracks is fairly small per cycle,
and not visibly noticeable until more than 90% of the test has occurred (i.e. less then 10%
of cycles to failure remaining).
The B samples also failed in the shear region. However, the initial failure was found to be
the de-bonding of the FRP resin at the contact where the 4 to 9 mm wide flange of the Zshape longitudinal rib and skin meet. After this, shear failure would propagate through the
foam core (Figure 4b), as a secondary failure, in the same fashion as the A samples.
a)

b)

Fig. 4. a) Sample A1 (100% load) Failure, and b) Sample B3 (40% load) Failure
3.3

The A-Sample Results

The number of cycles to failure at each loading level of the A-samples is presented in Table
1. These data points have been plotted (Figure 5a) in terms of the loading percent and
number of cycles to failure. It can be noted that the scale of the plot is best viewed by
displaying the number of cycles (x-axis) in a base 10 logarithmic scale.
Table 5: Test Results for A-Samples
ID
A1
A6
A2
A5
A3

Loading Amplitude
Metric Value
% of
(kN)
Ultimate
14.551
10.186
7.28
5.82
4.365

100
70
50
40
30

Cycles to
Failure
1
5,089
46,208
134,500
2,358,795

240

Total
Thickness
(mm)
79
77
79
76
77

Sample
Width
(mm)
637
634
635
635
628

(a)
(b)
Fig. 5. a) Test Results for A Samples Compared to A1 Ultimate Test, and b) Test Results
for B Samples Compared to B1 Ultimate Test
To construct a useful model, a general equation was formulated for this particular data.
Three known facts were used to construct this mathematical model. First, all relevant cycles
will be between 1 and infinite number of cycles. Secondly, the load value cannot exceed
100% of ultimate, and will have a minimum greater than or equal to zero that will never fail
at infinite cycles.
Previous FRP fatigue experiments suggest an s-shape to the S-N Curve for FRP soft foam
core sandwich panels (5) and (7). Therefore, for initial assumptions, a sigmoid function is
used to establish the s-shape of the graph, as given by Equation 1.
1
(1)
f ( x)
1 ex
Through transformation of this base equation, simplification and rearrangement of
constants, Equation 2 is derived:
S(N )

A
1 B* Nm

(2)

These determined values for the A-samples are as follows; m = logarithmic conversion
constant, 0.423822; C = fatigue limit as a fraction of ultimate, 0.22875; N = number of
cycles to failure, S(N) = Stress amplitude as fraction of ultimate; A and B are undefined
material constants, with A = 0.786675 and B = 0.02. The best fit relationship is displayed
in Equation 3.
S(N )

0.786675
0.22875
1 (0.02) N 0.423822

(3)

From this, it is estimated that the fatigue limit is about 23% of ultimate (Figure 5a). This
means any amplitude of less than 23% will never cause the sample to fail (i.e. infinite
number of cycles).
3.4

The B-Sample Results

Similar data was compiled for the five ribbed B-samples, as shown in Table 2.

241

Table 6: Test Results for B-Samples


ID
B1
B2
B3
B5
B4

Loading Amplitude
Metric % of Ultimate
Value
(B-Samples)
(kN)
26.79
100
16.074 60
10.716 40
8.037
30
5.358
20

% of Ultimate Cycles
Failure
(A-Samples)

to Overlap
Range
(mm)
1
6-8
250
4-8
5,116
6
37,227
6-9
2,302,019
7

184
110
74
55
37

Total
Thicknes
s (mm)
77
79
79
77
77

Sample
Width
(mm)
632
638
636
633
633

Similar to the A-samples, Equation 2 can be manipulated to represent the fatigue response
of B-samples, as follows:
S(N )

1.110158
0.124245
1 (0.267659) N 0.283983

(4)

The equation fit of the experimental results is displayed in Figure 5b.


3.5 Contribution of Ribs to Fatigue Performance
The difference of contact area between skins and the foam core for both types A and B
samples is very minimal. Therefore, the primary difference in fatigue resistance is
dominantly due to the addition of the ribs. By plotting the absolute value of loads versus
number of cycles, the comparison of both S-N curves for both types of panels can be
directly made (Figure 6a).

(a)
(b)
Fig. 6. a) Combination of S-N Curves, and b) Approximate Contribution of Internal Ribs to
Fatigue Resistance

242

The contribution of the rib to fatigue resistance is illustrated again in Figure 6b. It is
determined that the additional load resistance by adding ribs is 12,239 N for the monotonic
loading case (i.e. about 84% increase from the non-ribbed sample capacity) but this
contribution reduces to less than 600 N (i.e. only 4-5% increase from the non-ribbed sample
capacity) after only 10,000 cycles.
4. CONCLUSION
The following preliminary conclusions are drawn:
4
Under fully reversed bending fatigue, in order for the sandwich panel to achieve one
to two million cycles, the load level should be limited to 30% of ultimate in panels
without ribs and limited to 20% of ultimate for panels with internal rib tested in this
study.
5
Shear failure by diagonal tension occurred in the foam core of the panels without ribs.
In panels with internal rib, delamination between the flange of the Z-shape rib and
skin occurred, followed immediately by shear failure of the core.
6
The contribution of the rib to the monotonic ultimate strength was about 84%.
7
The contribution of the rib to the ultimate strength reduces to about 4% at 10,000
cycles. It can be concluded that the contribution of ribs to fatigue resistance is
negligible beyond 10,000 cycles, for panels tested in this study.
5. FUTURE WORK
Investigations will need to be conducted in order to determine if increasing the bonding
area of the rib to skin will effectively improve the panels fatigue resistance. If not, it may
be sufficient to ignore any rib effects for all high cycle fatigue of GFRP-polyurethane
sandwich panels. Future work is also required to follow ASTM C273/C273M-07a and
ASTM C394-00 to determine the scientific precision of the results found. For this, it is
recommended to use smaller, more manageable samples. Future results intend to relate
physical material properties of the GFRP panels and relate to the determined mathematical
S-N curves.
6. REFERENCES
1.
2.
3.
4.
5.

Teng, J.G., Chen, J.F., Smith, S.T., and Lam, L. 2002. FRP-Strengthened RC
Structures. John Wiley & Sons Ltd., Chichester, 266 p.
Barbero, E.J., Lopez-Anido, R. and Davalos, J.F. 1993. On The Mechanics of ThinWallled Laminated Composite Beams. SAGE Publications, Journal of Composite
Materials, 27(8): 806-829.
Osgood, C.C. 1982. Fatigue Design. 2nd Edition. Oxford, UK : Pergamon Press, 1982.
Burman, M. 1998. Fatigue Crack Initiation and Propagation in Sandwich Structures.
Doctoral Thesis. Stolkholm, Sweden: Kungliga Tekniska Hgskolan (Royal Institute of
Technology), ISSN 0280-4646.
Shenoi, R.A., Clark, S.D. and Allen, H.G. 1995, Fatigue Behaviour of polymer
composite sandwich beams. Journal of Composite Materials, 29(18): 2423-2445.

243

6.
7.
8.
9.

Challis, K.E., Hall, D.J. and Paul, D.B. 1986. A Novel Method for Determining the
Temperature Dependence of Shear Properties of Structural Foams. Cellular Polymers,
5: 91-101.
Halfpenny, A. 2005. A Practical Introduction to Fatigue. Sheffield, UK: nCode
International Ltd.
Abelkis, P.R. 1980. Design of Fatigue and Fracture Resistant Structures. Bal Harbour,
Florida : ASTM, 81-68806.
Allen, H.G. 1969. Analysis and Design of Structural Sandwich Panels. Toronto:
Pergamon of Canada Ltd, ISBN: B0006 BVJ44.

244

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

HEMP FIBER COMPOSITES FOR STRUCTURAL RETROFIT


D. Asprone 1, M. Durante2 and A. Prota1
1

Department of Structural Engineering - University of Naples Federico II, Naples, Italy


Department of Materials and Production Engineering - University of Naples Federico II,
Naples, Italy

ABSTRACT
Currently, sustainability represents a primary issue for construction industry. New material
and technological solutions are widely proposed and investigated to meet sustainability
requirements and natural fibers represent one of the most studied materials. The work
presented here investigated the mechanical behavior of a sustainable composite system
made by pozzolanic mortar reinforced with hemp fiber grids. To improve the durability of
the system and in particular of the fibers in the pozzolanic mortar environment a latex
coating was used. The objective of the study was to investigate the mechanical behavior of
the proposed composite system and assess the feasibility of using the system for structural
retrofit applications on existing structures. A mechanical characterization of the fibers was
conducted and the effectiveness of the latex coating in improving the durability of the fibers
was investigated. The mechanical behavior of the composite system was studied, through a
three-point bending test program.
1. INTRODUCTION
Currently sustainability of civil structures represents a primary requirement in
construction industry. Low environmental impact of operations on built environment is
often a primary condition to be respected by construction designers and operators. Thus,
recently, new technological solutions are proposed and new materials are investigated
and used. For this reason, in recent years, natural fibers have been widely investigated,
to be used as an alternative to carbon, glass or plastic fibers, in several composite
applications for construction industry. In fact, given their low environmental impact both
in production and in dismission phase, natural fibers represent a highly sustainable
material. Furthermore, natural fibers can be locally supplied, ensuring a sustainable
production chain.
The recent increasing scientific interest in natural fibers as a component of construction
applications is also due to the good mechanical properties exhibited by natural fibers.
Available literature provides the mechanical characterization of natural fibers, in terms

245

of elastic properties and tensile strength [1,2]; in particular, attention has been focused on
different fibers, e.g. flax [3], jute [4, 5], hemp [6, 7], sisal [8]. The review presented in
Bledzki and Gassan [1] reports the main mechanical properties of various natural fibers.
It can be observed that the tensile strength can reach more than 1000 MPa, in case of
flax fiber and vary from about 400MPa to 800MPa in case of jute fiber, whereas hemp
fiber exhibits a tensile strength of 690MPa. Furthermore, the ultimate tensile strain
varies from 1.5% for the jute fiber to 3.2% for the flax fiber, whereas hemp fiber
presents an ultimate tensile strain of 1.6%. Young modulus is equal to 26.5GPa and 27.6
GPa for jute and flax fiber, respectively, whereas, according to Dhakal et al. [7], Young
modulus for hemp fiber varies from 30GPa to 60GPa. Thus, the values of the main
mechanical properties of natural fibers are not so far from those exhibited by the most
used synthetic fiber, i.e. glass or carbon.
The objective of the present paper is to propose and investigate an inorganic composite
system, made by a pozzolanic mortar reinforced by a hemp fiber grid, to be potentially
used in seismic retrofitting operations of existing masonry structures, in Italy. In fact,
Italian guidelines from National Research Council [9] permit to use, for seismic
retrofitting of masonry structures, externally bonded fiber reinforced composite systems,
using inorganic matrices. From the structural point of view, the use of natural fibers and
in particular hemp fibers guarantees a good mechanical compatibility of the composite
system with the masonry elements, given the low Young modulus of the fibers. Here, the
results of the preliminary experimental characterization of the composite system are
presented. In particular, the single hemp fibers have been characterized in terms of
tensile strength; furthermore, the durability of the fibers in the matrix environment was
also addressed and a latex coating was investigated. Finally, the flexural behavior of the
composite was investigated using different configurations of the fiber reinforcement.
It is emphasized that the choice of the specific components for the composite system is
led by sustainability criteria. In fact, both hemp fibers and pozzolanic mortar are very
common in Italy and the production of the composite system can be provided by a local
supply chain.
2. MACHANICAL CHARACTERIZATION AND DURABILITY ASSESSMENT
OF THE HEMP FIBERS
The hemp fibers employed in the investigated composite systems were produced in Italy,
from a local growing. In order to characterize the mechanical properties of the fibers 10
tensile failure tests were carried out on fiber specimens, using a displacement control
testing machine. The tests were conducted and elaborated according to ASTM C1557
[10]. The free length of the fiber specimens was 15mm and the tests were conducted at
0.1mm/sec of elongation velocity. The specimens used for the tests came from 5 fibers;
each of them were divided into 2 specimens. During the tests, the load and the specimen
elongation were acquired and elaborated into stress-strain relationships. To do this, the
cross area A of the fibers was evaluated as

246

P
L

(1)

being P the weight of the fiber, L the length of the fiber and g the specific weight of the
fiber, which can be considered equal to 1.4104 N/m3 for cellulose based materials [11]. The
specimens exhibited an elastic behavior up to failure. As it was expected, the values are
highly variable; the obtained average tensile failure stress and average Youngs modulus
were equal to 898MPa and 21.3GPa, respectively, whereas the average ultimate strain was
equal to 6.1%.
The durability of natural fibers represents a critical issue for the use of such fiber in
composite systems. Physical and mechanical properties of natural fibers can be affected by
significant modifications due to environmental factors [12, 13]. In particular, inorganic
matrix environment can induce a significant degradation of fiber properties [13, 14]. In case
of cementitious environment, durability of natural fibers is affected by different factors:
alkali attacks, chemical reactions with products of cement hydration and fibers volume
variation due to water absorption [14-17].
Actually, in the present study, the use of a pozzolane-based mortar, characterized by a
lower alkalinity than those presented by purely cementitious mortars, can reduce the effects
of the alkali attacks on the hemp fibers. However, in order to improve the durability of the
hemp fibers in the proposed composite system, a latex coating was investigated, as
described in the following sections.
3. MECHANICAL CHARACTERIZATION OF THE COMPOSITE SYSTEM
The investigated composite system consists of a thin pozzolanic mortar slab reinforced with
different layers of hemp fiber grids. To assess the mechanical properties of the hemp fiber
reinforced composite system a bending test program was conducted on different composite
specimens, with different grid reinforcement configurations. Figure 1 depicts a composite
specimen during preparation, where a hemp fiber grid has just been placed on the fresh
mortar. In particular, three-point bending tests were carried out. The samples were almost
300mm long, 15mm thick and 100mm wide. The span length L used for the test was equal
to 240mm. Up to 6 reinforcement layers were placed in the composite samples. The
position through the thickness of each layer is such as indicated in Figure 2; in particular,
two different configurations were tested, using 4 and 6 layers, adding the inner layers in the
latter case.

247

Fig. 1. A composite specimen during preparation

top

reinforcement layer 2

reinforcement layer 4

reinforcement layer 6

reinforcement layer 5

reinforcement layer 3

reinforcement layer 1

bottom

Fig. 2. Reinforcement layers position


Two different values of grid spacing through the width of the sample, Sw, were used, equal
to 10mm and 20mm, whereas the grid spacing through the length of the sample, Sl, was
equal to 40mm in each specimen. Two different values for the diameter of the single fiber
bundle, Db, were used, equal to 1mm and 0.3mm, corresponding to a weight per length of
the bundles of 1.1 g/m and 0.1 g/m, respectively. Both latex and resin coatings were used.
Table 1 reports the main characteristics of the tested samples.

248

Table 1. Three-point bending tests

sample

coating
type

P
B
A
D
H
L
C
F
E
G
I

--

number of
bundle
reinforcing diameter, Db
layers [#]
[mm]
0
4

latex
6
4
resin
6

-0.3
1.0
0.3
0.3
1.0
0.3
1.0
0.3
1.0
1.0

Bundle
Average
spacing
reinforcement
maximum
through the weight ratio
flexural
width, Sw,
[%]
stress [MPa]
[mm]
--7.03
20
1.2
10.20
20
4.0
9.35
20
1.8
9.27
10
3.0
13.70
10
12.0
13.20
20
1.2
9.90
20
4.0
10.27
20
1.8
11.10
20
7.0
15.02
10
12.0
21.64

The tests were conducted after 28 days of curing the mortar, using a displacement control
esting machine, with a maximum load capacity of 50kN. During the tests, the applied load
P, and the midspan deflection D, were acquired. According to ASTM C1018 approach [18],
the load and deflection data were elaborated to obtain flexural stress-strain curves, through
the following relationships:

3PL
2bd 2

(2)

6Dd
L2

(3)

where:
s the flexural stress;
e the flexural strain;
P, the applied load;
D, the midspan deflection;
L, the span length;
b, the width of the specimen;
d, the thickness of the specimen.
The average maximum flexural stress is reported in Table 1. It can be observed that in the
unreinforced specimen a maximum stress of 7.03MPa is reached. On the contrary, for the
reinforced specimen, it was experienced that after a peak stress occurring almost at the
same flexural strain value, the stress increases again and higher values of strain are
obtained. In particular, higher strain values are reached for the specimens with latex coated
249

reinforcing fibers, whereas higher values of stress are obtained in case of the specimens
with resin coated reinforcing fibers. With regard to the diameter of the fiber bundles, it can
be observed that, in case of the latex coating, unlike the resin coating, the increase of the
diameter does not correspond to an increase of the maximum flexural stress. This reveals
that latex coating is less effective than resin coating in transferring bond stress inside the
bundles. Figure 3 depicts a specimen with latex coated fibers after failure.

Fig. 3. Specimen reinforced with latex coated fibers after failure


Finally, it can be observed that, both for latex and resin coating, an increase of the
reinforcement ratio produces a growth of the maximum flexural stress, both in case of
reduction of the bundle spacing and in case of increase of the reinforcing layers. However,
such variation is much less significant in the latter case, as it could be expected, since the
added reinforcing layers are less effective than the others, since they are closer to the
neutral axis in the bended cross section of the specimen.
4. CONCLUSIONS
The activities here presented addressed a preliminary mechanical characterization of an
inorganic composite system that could become a potential solution for the retrofit of
existing structures. The system is made by pozzolanic mortar reinforced with hemp fiber
grids, with a latex coating. Based on the results of the conducted experimental activities it
can be concluded that hemp fiber grid reinforcement can provide a significant improvement
of the flexural behavior of the pozzolanic mortar, increasing the flexural strength and
providing a considerable durability enhancement.
At this step, further tests will be conducted to assess the durability of the composite system
within different environmental conditions and to investigate the feasibility of applications
of the reinforced mortar on structural elements. However, the preliminary outcomes here
illustrated reveal that the composite system presents considerable mechanical properties
and can be further developed to be used as an effective solution for structural retrofit of
existing structures. Furthermore, once the mechanical behavior of the composite per se has
been optimized, it will be necessary to assess issues related to its bond to existing structural
members.

250

As a final word, it is underlined that, as it was expected for natural fibers, results from both
fiber and composite tests presented a high dispersion. This represents a critical issue for
structural applications, where a design value for each mechanical property of the used
materials is needed.
5. REFERENCES
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.

Bledzki, A.K. and Gassan, J. 1999. Composites reinforced with cellulose based fibres.
Progress in Polymer Science, 24(2): 221-274.
Gassan J. 1997. Natural fibre-reinforced plastics Correlation between structure and
properties of the fibres and the resultant composites. Dissertation at the Institute of
Materials Engineering, University of Kassel, Kassel.
Stamboulis, A., Baillie, C.A., and Peijs, T. 2001. Effects of environmental conditions
on mechanical and physical properties of flax fibers. Composites Part A: Applied
Science and Manufacturing, 32(8): 1105-1115.
Mannan, K.M., and Talukder, M.A.I. 1997. Characterization of raw, delignified and
bleached jute fibres by study of absorption of moisture and some mechanical
properties. Polymer, 38(10): 2493-2500.
Gassan, J. and Bledzki, A.K. 1999. Possibilities for improving the mechanical
properties of jute/epoxy composites by alkali treatment of fibres. Composites Science
and Technology, 59(9): 1303-1309.
Kostic, M., Pejic, B., and Skundric, P. 2008. Quality of chemically modified hemp
fibers. Bioresource Technology, 99(1): 94-99.
Dhakal, H.N., Zhang, Z.Y., and Richardson, M.O.W. 2007. Effect of water absorption
on the mechanical properties of hemp fibre reinforced unsaturated polyester
composites. Composites Science and Technology, 67(7-8): 1674-1683.
Li, Y., Mai, Y.W., and Ye, L. 2000. Sisal fibre and its composites: a review of recent
developments. Composites Science and Technology, 60(11): 2037-2055.
CNR, 2004. Guide for the Design and Construction of Externally Bonded FRP Systems
for Strengthening Existing Structures - Materials, RC and PC structures, masonry
structures. CNR-DT 200/2004, Rome, Italy, 144 p.
ASTM. 2008. Standard Test Method for Tensile Strength and Young's Modulus of
Fibers. ASTM C1557, 10 p.
Fakirov S. and Bhattacharyya, D. 2007. Handbook of Engineering Biopolymers:
Homopolymers, Blends, and Composites. Hanser Gardner, 932 p.
Gram, H.E. 1983. Durability of natural fibres in concrete. CBI Research No. 1-83,
Swedish Cement and Concrete Research Institute, Stockholm, 255 p.
Jayamol, G., Sreekala, M.S. and Sabu, T. 2004. A review on interface modification and
characterization of natural fiber reinforced plastic composites. Polymer Engineering
and Science, 41(9): 1471 1485.
Bledzki, A.K., Reihmane, S. and Gassan, J. 1998. Properties and modification methods
for vegetable fibers for natural fiber composites. Journal of Applied Polymer Science,
59(8): 1329 1336.
Filho, T.R.D., Scrivener, K., England, G.L. and Ghavami, K. 2000. Durability of alkali
sensitive sisal and coconut fibres in cement based composites. Cement and Concrete
Composites, 6(22): 127143.

251

16. Canovas, M.F., Selva, N.H. and Kawiche, G.M. 1992. New economical solutions for
improvement of durability of Portland cement mortars reinforced with sisal fibres.
Materials and Structures, 25: 417422.
17. Filho, T.R.D. 1997. Natural fibre reinforced mortar composites: experimental
characterisation. PhD thesis, DEC-PUC-Rio, Brazil, 472 p.
18. ASTM. 1997. Standard Test Method for Flexural Toughness and First-Crack Strength
of Fiber-Reinforced Concrete (Using Beam with Third-Point Loading). ASTM C101897, Pennsylvania. USA, 7 p.

252

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

EFFICIENCY OF DIFFERENT FRP-BASED FLEXURAL


STRENGTHENING TECHNIQUES IN BEAMS SUBMITTED TO
FATIGUE LOADING
J. Sena-Cruz1, J. Barros2, M. Coelho3 and L. Silva4
1

Assistant Professor, ISISE, Dept. of Civil Engrg., Univ. of Minho, Portugal.


Associate Professor, ISISE, Dept. of Civil Engrg., Univ. of Minho, Portugal.
3
Researcher, ISISE, Dept. of Civil Engrg., Univ. of Minho, Portugal.
4
MSc Student, Dept. of Civil Engrg., Univ. of Minho, Portugal.
2

ABSTRACT
Flexural strengthening of reinforced concrete (RC) structures with fiber reinforced
polymers (FRP) has been used mostly by two main techniques: Externally Bonded
Reinforcement (EBR) and Near-Surface Mounted (NSM). In both strengthening techniques
the FRP systems are applied on the cover concrete, which is normally the weakest region of
the element to be strengthened. Consequently, the most common problem is the premature
failure of the system, which occurs more frequently when using the EBR technique. In an
attempt of overcoming this weakness, another flexural strengthening technique, named MFEBR Mechanically Fastened and Externally Bonded Reinforcement, is analyzed in the
present paper. This technique uses multidirectional carbon fiber laminates that are
simultaneously glued and anchored to concrete. To compare the efficiency of NSM, EBR
and MF-EBR techniques, four-point bending tests with RC beams were carried out under
monotonic and fatigue loading. In this work the tests are described and the obtained results
are presented and discussed.
1. INTRODUCTION
Fiber reinforced polymers (FRP) are gaining acceptance in the rehabilitation and/or
strengthening of existent structures since they are good alternatives to traditional
strengthening systems, due to their high stiffness and tensile strength, low weight, good
fatigue behavior, immunity to corrosion, and geometric versatility. The FRPs are used,
mainly, by two strengthening techniques (ACI 2008): the Externally Bonded
Reinforcement (EBR) and Near-Surface Mounted (NSM). The EBR technique has been
used to increase the flexural and the shear resistance of reinforced concrete (RC) elements,
as well as the concrete confinement, and to control the cracking process. Despite the
advantages of this technique, the bond between FRP and concrete surface can be
susceptible of degradation, particularly due to environmental conditions such as fire, high

253

temperatures, UV radiation, humidity and even vandalism acts. The NSM technique was
proposed as an alternative strengthening technique and, when compared to the EBR, several
advantages can be pointed out, mainly: the amount of site installation work and the surface
preparation may be reduced; the NSM technique is less prone to the debonding from the
concrete substrate; the FRP reinforcements can be more easily anchored into adjacent
members (preventing debonding failures); the FRP reinforcements are protected by the
concrete cover and, consequently, less exposed to mechanical damage and impact loading,
and less exposed to the fire and vandalism acts. Furthermore, the aesthetic of the
strengthened structure is virtually unchanged (De Lorenzis and Teng 2007). Several
studies, however, have shown the frequent occurrence of FRP debonding when the EBR
(CNR 2004) is used, and concrete cover rip-off failure mode when NSM technique is
applied (Barros and Fortes 2005). Besides being both fragile failure modes, they do not
allow the adequate exploitation of the tensile potentialities of FRP systems (the maximum
stress installed in the FRP at failure of the strengthened element is much lower than its
tensile strength).
In alternative to EBR and NSM, a quite new technique, called Mechanically Fastened FRP
(MF-FRP), has been proposed based on the use of steel fasteners applied along the
laminates length. The application of the MF-FRP technique in the flexural strengthening of
RC elements improves the flexural capacity with little or no loss in ductility (Martin and
Lamanna 2008). In the last decade the MF-FRP technique has been investigated by several
researchers and some benefits have been pointed out, namely, quick installation with simple
hand tools, no special labor skills are needed, no surface preparation is required, and the
strengthened structure can be used immediately used after installation of the FRP (Elsayed
et al. 2009, Lee et al. 2009, Bank and Arora 2007, Quattlebaum et al. 2005, Lamanna et al.
2004). Nevertheless, some notable disadvantages of this system have been observed,
including scale effects, cracking induced by the impact of fasteners in high-strength
concrete, and less-effective stress transfer between the FRP and concrete due to the discrete
attachment points (Ray et al. 2000).
Based on the MF-FRP technique, the Mechanically Fastened and Externally Bonded
Reinforcement (MF-EBR) technique has been proposed. The MF-EBR combines the
fasteners from the MF-FRP technique with the externally glued properties from the EBR. In
addition, the fasteners are pre-stressed and multidirectional laminates exclusively made
with carbon fiber reinforced polymers (MDL-CFRP) with high longitudinal tensile
strength, modulus of elasticity and bearing strength are used in order to mobilize high
levels of efficiency (Sena-Cruz et al. 2010a). To assess the efficiency of EBR, NSM and
MF-EBR techniques, four-point bending tests with RC beams were carried out under
monotonic and fatigue loading. The tests are described and the results are presented and
discussed in detail.
2. EXPERIMENTAL PROGRAM
2.1 Specimens and Test Configuration
The experimental program was composed of two series of four beams each, being the
distinction between the series associated to the loading configuration: one subjected to
254

20
31

69

69

CFRP
2x(30x1.4)
31

200

6@100

NSM
4x(15x1.4)

MDL-CFRP
2x(30x2.1)

45 30 50 30 45
200

6@100

510

300

510

100

6@100

300

510

300

6@100

510

300

20

monotonic loading and the other to fatigue loading. Each series was composed of a
reference beam (REF), and a beam for each investigated strengthening technique. The RC
beams have a cross section of 200 mm wide and 300 mm height, being 2000 mm the
distance between supports. All the beams have three longitudinal steel bars of 10 mm
diameter (310) at the bottom, and 210 at the top (see Fig. 1). The transverse
reinforcement is composed of steel stirrups of 6 mm diameter (6) with a constant spacing
of 100 mm in order to avoid shear failure. Fig. 1 includes the cross section of the beams.

45 30 50 30 45
200

40 40 40 40 40
200

(a)
(b)
(c)
(d)
Fig. 1. Cross section of the all beams: (a) REF; (b) EBR; (c) MF-EBR; (d) NSM. Note: all
dimensions are in [mm].
Table 1 presents the main properties of the beams. In this table tf, Lf and wf are the
thickness, the length and the width of the laminates, respectively, and s,eq is the equivalent
longitudinal steel reinforcement ratio defined by Eq. 1:
s,eq

E f Af
As

bd s Es bd f

(1)

In this equation b is the width of the beam, As and Af are the cross sectional area of the
tensile longitudinal steel bars and FRP systems, respectively, Es and Ef are the modulus of
elasticity of steel and FRP, respectively, and ds and df are the distance from the top concrete
compression fiber to the centroid of the steel bars and FRP systems, respectively. In all the
strengthened beams similar s,eq was applied.
Beam
REF
EBR
MF-EBR
NSM

Table 1. Properties of the beams.


Laminate type
Quantity
tf [mm]
Lf [mm]
Unidirectional
2
1.41
1400
Multidirectional
2
2.07
1400
Unidirectional
4
1.41
1400

wf [mm]
30
30
15

s,eq [%]
0.439
0.550
0.553
0.561

In this experimental study, a four-point loading test configuration was adopted for the
monotonic and fatigue tests (see Fig. 2a). A servo-controlled hydraulic system was used to
perform the monotonic tests under displacement control, with a deflection rate of 20 m/s,
using the linear variable differential transducer (LVDT) located at the mid-span of the
beam (LVDT3 in Fig. 2). The load was applied through an actuator equipped with a load
cell of 500 kN of maximum capacity. The fatigue tests were performed between a
minimum fatigue level of Smin=25% and maximum fatigue level of Smax=55%, where the S
is the ratio between the applied load and the monotonic beams load carrying capacity, Fm.
The fatigue tests were composed by three main steps: initially, a monotonic loading was
255

applied, under force control and at a load rate of 100 N/s up to the maximum level (Smax), in
order to register the initial response of the specimen; then, 1 million cycles was imposed at
2 Hz of frequency between SminFm and SmaxFm; finally, a monotonic loading up to the
failure, with the same configuration of the monotonic tests, was applied to the specimens.
In addition to the LVDT3, others four LVDTs were used to record the deflections in the
loaded sections (LVDT2 and LVDT4) and at the sections coinciding with the free ends of
the FRP systems (LVDT1 and LVDT5), see Fig. 2a. Strain gauges were glued on both the
longitudinal steel reinforcement and FRPs to measure the strains during the tests, see
Fig. 2b-e.
F/2

F/2

CL

(a)

300

LVDT5

LVDT4

LVDT3

600

200

300

600

(b)

All beams

300

SGs2

100

500

200

SGs1

310

100

(c)

EBR beam

100

SGf3

SGf4

UD-CFRP
laminates
150

300

SGf2

150

300

100

(d)

MF-EBR beam
45 30 50 30 45

100

SGf1

100

100

300

SGf2

100

100

200

SGf3

100

100

100

300

SGf4

100

100

100

100

100

100

100

MDL-CFRP

100

(e)

NSM beam
40 40 40 40 40

100

200

45 30 50 30 45

SGf1

200

100

100

300
Support Line

SGf5
SGf1

150
FRP end

SGf6

SGf7
SGf3

SGf2

150

300
Load line

SGf8
SGf4

UD-CFRP
laminates

200

100

LVDT2

LVDT1

300

All beams

100
Midspan

Fig. 2. Instrumentation adopted: (a) vertical deflection; (b) strains on the steel bars; (c)
strains on the laminate of the EBR beam; (d) strains on the laminate of the MF-EBR beam;
(e) strains on the laminates of the NSM beam. Note: all dimensions are in [mm].
2.2 Material Characterization
The mechanical characterization of concrete was assessed by means of compression tests.
For this purpose six cylindrical concrete specimens were tested at the age of the tested
beams to evaluate the compressive strength and the modulus of elasticity according to the
recommendations NP EN 12390-3:2009 and LNEC E397:1993, respectively. From the
compression tests, an average compressive strength value of 53.08 MPa, with a coefficient
of variation (CoV) of 4.0%, and an average value of 31.17 GPa (CoV=4.4%) for the
modulus of elasticity, were obtained. The age of the concrete beams at the date of
experimental program was about two years. The steel of the longitudinal bars and stirrups
has a denomination of A400 NR SD according to NP EN 1992-1-1:2010. Additional
256

information related with the experimental characterization of the steel bars can be found
elsewhere (Bonaldo 2008).
In this work two different types of CFRP laminates were used: unidirectional (UD-CFRP)
for the case of EBR and NSM techniques, and multidirectional (MDL-CFRP) for the case
of the MF-EBR technique. Tensile tests were performed according to the ISO 527-4:1997
for both laminates (UD-CFRP and MDL-CFRP) to assess their tensile properties. From
these tests it was obtained a tensile strength, a modulus of elasticity and an ultimate strain
of 1866 MPa (CoV=5.1%), 118 GPa (CoV=2.8%) and 1.58 % (CoV=5.1%) for MDLCFRP, and 2435 MPa (CoV=5.8%), 158 GPa (CoV=3.9%) and 1.50 % (CoV=4.7%) for
UD-CFRP, respectively (Sena-Cruz et al. 2010a). From the bearing tests with MDL-CFRP
specimens performed according to the ASTM D5961/D5961M05 standard, a bearing
strength of 316.4 MPa (CoV=11.8%) and 604.4 MPa (CoV=5.8%) was obtained for the
case of unclamped and clamped with a torque of 20 Nm, respectively (Sena-Cruz et al.
2010a). To bond the laminates to concrete the S&P Resin 220 epoxy adhesive was used.
According to the supplier, this adhesive has a flexural tensile strength, a compressive
strength and a bond concrete/laminate strength of 30 MPa, 90 MPa e 3 MPa, respectively.
A Hilti chemical anchors system was adopted to fix mechanically the laminate to
concrete for the case of the MF-EBR beam. This system is composed by the resin HITHY 150 max and the HIT-V M8 8.8 threaded anchors. The preparation of the strengthened
beams required several steps that are described elsewhere (Sena-Cruz et al. 2010b).
3. RESULTS
Table 2 resumes the main results obtained in the performed tests, while Fig. 3 depicts the
relationship between force and displacement at mid-span during the tests. In this table the
meaning of the symbols are: cr, deflection at concrete crack initiation; Fcr, load at concrete
crack initiation; y, deflection at the yield initiation of the longitudinal steel bars; Fy, load at
the yield initiation of the longitudinal steel bars; max, deflection at the maximum load;
Fmax, maximum load; fu, ultimate strain in the FRP according to the results obtained in
tensile tests; fy, maximum strain in the FRP at Fy; fmax, maximum strain in the FRP at
Fmax.
In terms of monotonic testes, it can be concluded that the most effective strengthening
technique was the MF-EBR, not only due to the maximum load reached (Fmax=148.2 kN),
but also in terms of deflection at failure and maximum/failure strain ratio in the FRP. When
compared with the EBR, the MF-EBR system had an increase of the load carrying capacity
of about 37%. This superior behavior cannot be explained by the higher axial stiffness, EfAf,
of the laminate, since the ratio between the EfAf of the MDL-CFRP and EfAf of the UDCFRP (used in the EBR beam) is only 1.08. The pre-stressed anchors have contributed for
this higher strengthening effectiveness of MF-EBR technique. In fact, while EBR FRP
systems failed by FRP peeling (Fig. 4a), and NSM FRP systems by concrete cover rip-off
(detachment of the concrete cover that includes the CFRP strips, Fig. 4b), the MF-EBR
FRP laminates failed by bearing (Fig. 4c-d). The presence of the anchors avoided the
premature debonding (peeling) of the laminates, as well as the detachment of the concrete
cover (rip-off). Defining the level of ductility as the ratio between the deflection at the
257

maximum load and the deflection at the yielding of the longitudinal steel bars (max/y), in
the MF-EBR beam the max/y was equal to 4.35, which was much higher than the values
registered in the EBR (1.80) and NSM (2.98) beams. Apparently, in the MF-EBR beam the
force corresponding to the crack initiation, Fcr, is higher than the Fcr of the other beams.
This behavior can be explained by the existence of pre-stress. In fact, the pre-stress
provided by the anchors may have induced a compressive stress state on the concrete cover
which has delayed the concrete crack initiation. This phenomenon could also explain the
higher stiffness in the phase between the concrete crack initiation and the steel yield
initiation of the MF-EBR beam. After the longitudinal steel bars have yielded, a slight
higher stiffness can be observed in the NSM beam, when compared with the MF-EBR
beam. This behavior can be justified by the confinement that surrounding concrete provides
to the NSM CFRP laminates.
Table 2. Main results obtained in the tests.
Crack
initiation
Beam
Fcr
cr
[mm] [kN]

Yielding

Ultimate

max/y fy/fu fmax/fu Failure


mode
[%]
[%]
Fmax
max
[kN]
[mm]
[mm]
MONOTONIC
REF 0.36
29
3.8
70
79.3
22.6
5.95
108.4
EBR 0.27
25
4.1
90
7.4
1.80
24.0
36.6 Peeling
(37%)*
MF148.2
18.3
4.35
15.8
69.3 Bearing
0.38
32
4.2
96
EBR
(87%)*
147.3
NSM 0.40
29
4.9
104
14.6
2.98
23.4
63.3 Rip-off
(86%)*
FATIGUE
REF 0.26
20
2.5
66
79.9
23.3
9.32
114.2
EBR 0.32
27
3.0
94
7.1
2.37
14.6
29.6 Peeling
(43%)*
MF147.2
0.35
31
3.7
101
12.9
3.49
15.0
63.4 Bearing
EBR
(84%)*
160.7
NSM N/A
N/A
3.3
105
22.2
6.73
15.4
55.7 Rip-off
(101%)*
*
(Fmax-Fmax,REF)/Fmax,REF where Fmax,REF is the maximum load of the reference beam in the
series.
y

Fy
[kN]

In terms of fatigue tests, the behavior at crack initiation, was similar to the one observed in
the monotonic tests, i.e., the best one was registered in the MF-EBR beam. After the fatigue
loading, the NSM was the most effective strengthening technique for both the maximum
load (Fmax=160.7 kN) and ultimate deflection capacity (max/y=6.73). When compared with
the corresponding monotonic tests, marginal variation in terms of maximum load was
obtained for the case of the REF, EBR and MF-EBR beams, whereas an increment of 9%
was attained in the NSM beam. No rational explanation can be pointed out for this
behavior. The inferior performance of the MF-EBR beam, when compared with the
monotonic one, can be attributed to a possible loss of efficiency of the pre-stressed
258

anchorages along the fatigue cycles, and due to a bearing strength degradation of the MDLCFRP during the cycles. The EBR and NSM beams exhibited the same failure modes
occurred in the monotonic tests. Despite the performance in the monotonic tests, the MFEBR beam presented a more fragile failure mode with bearing and inter-laminar failure of
the FRP. Stiffness degradation was also evaluated for the fatigue cycles. Marginal
variations were observed. In fact, a decrease of 8.3%, 3.0%, 0.3% and 12.1% in terms of
stiffness was observed for the REF, EBR, MF-EBR and NSM beams, respectively.
175

175
150

150

MF-EBR

NSM
125

125

MF-EBR

Force [kN]

Force [kN]

NSM
100
REF
75
EBR

50

100

EBR

75

REF

50
25

25

10

20

30

40

10

20

30

Displacement at mid-span [mm]

Displacement at mid-span [mm]

(a)
(b)
Fig. 3. Force vs. displacement relationship of the tested beams: (a) monotonic loading; (b)
fatigue loading.

(a)

(b)

(c)
(d)
Fig. 4. Failure modes: (a) EBR beam; (b) NSM beam; (c) and (d) the MF-EBR beam.

259

4. CONCLUSIONS
In this paper a flexural strengthening technique, called MF-EBR Mechanically fastened
and externally bonded reinforcement, is studied. This technique combines the fasteners
from the MF-FRP technique and the epoxy bond-based performance from the EBR
technique. In addition, all the fasteners are pre-stressed. Multidirectional laminates
exclusively made with carbon fibers reinforced polymers (CFRP) were used. To compare
the efficiency of the MF-EBR, EBR and NSM strengthening techniques, an experimental
program was carried out, composed of two series of four beams, one submitted to a
monotonic loading, and the other to a fatigue loading. For both series the beams were
subjected to a four-point bending loading configuration. In the fatigue tests 1 million cycles
at a frequency of 2 Hz was applied to each beam. Each series is composed of a reference
beam (REF) and one beam for each of the strengthening techniques analyzed: EBR, MFEBR and NSM. When compared to the reference beam, an increase on the loading carrying
capacity of 37%, 87% and 86% was obtained for the EBR, MF-EBR and NSM
strengthened beams, respectively. When compared to the EBR beam, an increase of about
37% on the load carrying capacity was obtained for MF-EBR technique. The most
favorable aspect of the MF-EBR technique was, however, the normalized deflection
capacity at maximum load (4.35), which was much higher than that registered in the EBR
(1.80) and NSM (2.98) beams. In addition, more ductile failure mode was observed for MFEBR technique. In the fatigue tests, after having been subjected to 1 million cycles, the
NSM beam has supported the highest ultimate load, corresponding to an increase of 101%,
while the MF-EBR and EBR beams presented an increase of 84% and 43% in the load
capacity, respectively, when compared with the maximum load of the control beam. In the
fatigue tests the NSM beam presented the highest normalized deflection capacity at
maximum load (6.7), while a value of 3.5 and 2.4 was registered in the MF-EBR and EBR
beams, respectively.
5. ACKNOWLEDGEMENTS
The
present
work
was
supported
by
Program
PIDDAC,
Project
No. PTDC/ECM/74337/2006 by FCT. The authors acknowledge the following companies:
Hilti, S&P, SECIL, and TSwaterjet, Lda.
6. REFERENCES
1.
2.
3.
4.

ACI Committee 440. 2008. Guide for the design and construction of externally bonded
FRP systems for strengthening concrete structures. ACI 440.2R-08, American
Concrete Institute, Farmington Hills, MI, USA, 76 p.
De Lorenzis, L. and Teng, J.G. 2007. Near-surface mounted FRP reinforcement: An
emerging technique for strengthening structures. Composites: Part B, 38: 119-143.
CNR-DT 200. 2004. Guide for the Design and Construction of Externally Bonded FRP
Systems for Strengthening Existing Structures. National Research Council, Rome,
157 p.
Barros, J.A.O. and Fortes, A. 2005. Flexural strengthening of concrete beams with
CFRP laminates bonded into slits. Cement and Concrete Composites, 27(4): 471-480.

260

5.
6.
7.
8.
9.
10.
11.

12.
13.
14.

Martin, J.A. and Lamanna, A.J. 2008. Performance of mechanically fastened FRP
strengthened concrete beams in flexure. Journal of Composites for Construction,
12(3): 257-265.
Elsayed, W., Neale, K., and Ebead, U. 2009. Studies on mechanically fastened fiberreinforced polymer strengthening systems. ACI Structural Journal, 106 (1): 49-59.
Lee, J.H., Lopez, M.M., and Bakis, C.E. 2009. Slip effects in reinforced concrete
beams with mechanically fastened FRP strip. Journal of Cement & Concrete
Composites, 31: 496504.
Bank, L. and Arora, D. 2007. Analysis of RC beams strengthened with mechanically
fastened FRP (MF-FRP) strips. Composite Structures, 79: 180191.
Quattlebaum, J.B., Harries, K.A., and Petrou, M.F. 2005. Comparison of three flexural
retrofit systems under monotonic and fatigue loads. Journal of Bridge Engineering,
10(6): 731-740.
Lamanna, A.J., Bank, L.C., and Scott, D.W. 2004. Flexural strengthening of reinforced
concrete beams by mechanically attaching fiber-reinforced polymer strips. Journal of
Composites for Construction, 8(3): 204-209.
Ray, J.C., Scott, D.W., Lamanna, A.J., and Bank, L.C. 2000. Flexural behavior of
reinforced concrete members strengthened using mechanically fastened fiber
reinforced polymer plates. Proc. 22nd Army Science Conf., The United States Army,
Washington, D.C., pp. 556560.
Sena-Cruz, J.M., Barros, J.A.O., and Coelho, M. 2010a. Bond between concrete and
multi-directional CFRP laminates. Advanced Materials Research, Vols. 133-134: 917922.
Bonaldo, E. 2008. Materiais compsitos e fibras discretas de ao no reforo de
estruturas laminares de beto. PhD thesis, Dept. of Civil Engineering, University of
Minho, Portugal, 404 p.
Sena-Cruz, J.M., Barros, J.A.O., Coelho, M., and Silva, L. 2010b. Efficiency of
different techniques in flexural strengthening of RC beams under monotonic and
fatigue loading. Submitted to Journal of Composites for Construction.

261

262

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

A REACTIVE MULTIPHASE MULTICOMPONENT APPROACH


FOR PREDICTING THE PERFORMANCE OF FRP COMPOSITES IN
CONCRETE
M. Boulfiza
Department of Civil and Geological Engineering, University of Saskatchewan, Canada

ABSTRACT
Classical accelerated aging procedures have failed to provide reliable means for predicting
the service performance/life of Glass Fiber Reinforced Polymers (GFRP) reinforced
concrete structures due to the inadequacy of variety of major assumptions at the hart of
those techniques. Among the most important questionable assumptions one can mention:
equivalence between time and temperature, automatic assimilation of chemical species
penetration with degradation, equivalence between a GFRP rod immersed in high pH
solution and a rod imbedded in concrete.
To overcome the above limitations, it is necessary to build more robust models that can
describe more accurately the penetration of potentially harmful substances into GFRP
composites and their subsequent interaction with the resin and fiber. Based on the current
state of knowledge, as represented by first principles of chemistry and physics, a
mechanistic model is proposed in this study for allowing a rational and accurate prediction
of the service performance/life of GFRP reinforced concrete structures. It is shown that
improved modeling capabilities are achieved by using an analytical model for the physical
chemical ageing of GFRP rods imbedded in concrete. The proposed model is based on the
potential chemical attack of the inner glass fiber surface - by diffusing Sodium Hydroxide and subsequent decrease of interfacial matrix - fiber strength. The mechanical retention of
the GFRP rod is described through a model that accounts for the eventual degradation of
the fiber-VE resin interface as a function of the extent of chemical attacks on the glass fiber
surface. It is worth noticing that the proposed approach is able to explain the discrepancies
between the results of traditional accelerated tests (high temperatures and high pH values)
and the measured or observed field conditions.
1. INTRODUCTION
Over the last few years, durability has been identified as one of the major parameters
controlling the long-term behavior of concrete structures [1]. Concrete structures are
typically designed to perform for 50 to 100 years with minimal maintenance. However, the
263

premature degradation of numerous structures exposed to chemically aggressive


environments has raised many questions with respect to the long-term durability of concrete
structures under such conditions. A major component of infrastructure decay is the
degradation of bridges and highways due to corrosion of the embedded reinforcing steel.
This problem has prompted the development of a variety of alternative strategies, such as
use of Fiber Reinforced Polymer composites (FRPs), for increasing the service life of
reinforced concrete structures exposed to harsh environments. However, adoption of these
materials by industry has fallen far short of initial expectations, due in large extent to a lack
of tools to evaluate their long term in-service performance. At present, no technically sound
procedure is available for evaluating the field performance of advanced reinforcement
materials available in the market over the service life of the RC structure [2]. Long-term
performance assessment of concrete structures by accelerated laboratory tests [6] or in situ
measurements is very difficult because of the extrapolations associated with the process.
This has led to a great deal of effort towards developing sound models that can reliably
predict the long-term performance of concrete structures. Despite the abundant scientific
and technical literature published on the durability of concrete over the past few decades
and FRP more recently, very little modeling work has been reported on comprehensive
models regarding the degradation of concrete. These concerns have motivated many
researchers to take advantage of recent improvements in computer engineering to develop
numerical models specifically devoted to the prediction of the long-term performance of
concrete structures subjected to harsh environments and loading. Based on first principles, a
comprehensive model is proposed in this study for predicting moisture flow and transport
of dissolved species in concrete and FRP reinforcement under stressed or unstressed
conditions. The ingress of the various aqueous species available in concrete along the depth
of concrete cover are calculated using a Pitzer-based reactive transport model which
considers all dominant chemical interactions with the concrete solid phases, pore water
solution, and FRP composite.
2. FRAMEWORK OF PROPOSED DURABILITY MECHANICS APPROACH
Depending on their service environment, concrete structures are exposed to various types of
aggressions (water, carbon dioxide, acids, chlorides, etc). Hence, their durability is
generally controlled by the transport properties of the material. One can distinguish the
following transport mechanisms in cement-based materials: transport of fluid by
permeation or diffusion (especially for the vapor phase); and transport of molecules or ions
by advection and/or diffusion. Accurate prediction of the durability of concrete structures
requires good knowledge of transport properties of concrete and their evolution over time.
The transport properties of concrete vary with age and deterioration caused by structural
and environmental loads. As the hydration reaction proceeds, more hydration products are
formed and concrete becomes tighter to water and chemicals penetration. Mechanical
loading as well as other environmental factors (temperature, freezing and thawing cycles
etc.) often lead to creation of new cracks and/or extension and widening of existing ones.
Cracks provide pathways for water movement and ions transport, hence, increasing the
penetration of harmful substances into concrete. Very little information exists on the
transport of substances in cracked concrete. The average fluid velocity across the crack
width, has been shown to obey a Darcy-type law. Whereas, mass transport in variably
saturated concrete or through the crack has been assumed to be controlled by an advective264

diffusive equation that can be considered as a generalization of Ficks second law of


diffusion since it includes transport by molecular diffusion as well as by advection (with
moving water).
To perform a dependable prediction of service performance of reinforced concrete
structures in their operating environment, one needs reliable models in the following areas:
Solid/structural mechanics models for predicting the structural response (deformation,
stresses, cracking, ); Models for predicting water flow, transport of aqueous chemical
species and their interactions with both the solid phase of concrete and reinforcement; and
Procedures or schemes to account for the interaction of all the above models. Fig 1
illustrates the framework of the proposed approach.
3. MATHEMATICAL MODELS
3.1. Reactive Transport in Concrete and GFRP
Concrete, considered here to be composed of a porous matrix or skeleton whose interstitial
volume, is occupied by two phases: liquid and gas. The two fluid components of interest in
the case of concrete are water and air; each of which can partition into each phase. The gas
phase is treated as a mixture of air and water vapour, where each constituent may exist in
any fraction between zero and one. On the other hand, the liquid phase is mostly water,
although small amounts of dissolved air can exist in solution.
Component balances are developed under the assumption of thermodynamic phase
equilibrium. Component balance equations for water and air take the form:
d w
Fw Q w
t
d a
Fa Qa
t

(1)

where d w ( da ) is the bulk density of water (air), Fw ( Fa ) denotes the net mass flux vector
of water (air), and Q w ( Qa ) denotes mass source for water (air). These fluxes encompass
both advective and diffusive transport processes. The reactions and sources take on positive
values for processes that increase concentrations (sources) and negative values for
processes that decrease them (sinks). In the above equation the mass of species are
increased or destroyed by a number of mechanisms depending on the problem being
considered. These mechanisms may include both homogeneous and heterogeneous
reactions. Homogeneous reactions, refer to processes that occur within the aqueous phase
itself while other fluid phases present in the void space and the solid matrix are not
involved. Heterogeneous reactions, on the other hand, refer to processes that involve
interactions between components in the aqueous phase and constituents of the solid phase,
or transfer across interphase boundaries within the void space. Examples include reactions
between aqueous species and solid phases in concrete and reinforcement. Reactions of
either type that contribute to the overall production or consumption of species may be
265

described by using equilibrium or kinetic models [3,4]. In equilibrium models the


concentrations of the multiple species are simultaneously adjusted within the aqueous phase
in accordance with equilibrium relationships and mass balance constraints. Such models are
appropriate when the rates of reactions are relatively fast (in comparison with other
transport and transformation mechanisms). The bulk densities are given by:
d w Ywl l Sl Ywg g S g

(2)

d a Yal l Sl Yag g S g

where Ywl ( Yal ) denotes the mass fraction of the water (air) component in the liquid phase
l ( g ) is the liquid (gas) phase density, and Sl ( S g ) is the liquid (gas) phase saturation,
the fraction of the porosity occupied by the liquid (gas) phase. The pore space is assumed to
be fully occupied by fluid, and hence, Sl S g 1 .
The advective fluxes are assumed to be properly described by the extended Darcy law, in
which relative permeabilities are introduced to account for the multiphase motion of fluids
[6]. Assuming that each phase has its own phase pressure, the mass fluxes of liquid and gas
phases are given by:

l krl
k Pl l g
l
g krg
k Pg g g
fg
g
fl

(3)

where P is pressure, g is the gravitational acceleration vector, and is the dynamic


viscosity. The intrinsic permeability tensor of the medium is k and the relative
permeabilities are denoted krl and krg . The intrinsic permeability tensor is assumed to be a
property of the material under consideration, and as such is a spatially heterogeneous
quantity. The water and air net mass fluxes appearing in the balance equations (1) and (2),
are assumed to be a superposition of component fluxes in each phase,

Fw Fwl Fwg

(4)

Fa Fal Fag

and each component phase-flux can be written as a sum of an advective (pressure-driven)


flux and a diffusive flux(concentration-driven),

266

Fwl Ywl fl J wl

Fwg Ywg f g J wg
Fal Yal fl J al

Fwg Ywg f g J wg

(5)

Because the gas is a mixture, each component will undergo interdiffusion whenever a
gradient in concentration exists [6]. The diffusive fluxes in the gas are approximated by

J wg g Dwg Ywg

(6)

J ag g Dag Yag

On average, the gas mixture as a whole moves with the average mass flux given by
f g Fwg Fag

(7)

Substituting eqn. (5) into (7), and using the fact that the mass fractions in a phase must sum
to unity, the diffusive fluxes in the gas phase must satisfy,
J wg J ag 0

(8)

In general, the diffusive fluxes of components dissolved in the liquid phase could be
described in the form
J wl Dwl wl
(9)
J al Dal al
where wl ( al ) is the concentration of the water (air) component in the liquid phase.
Diffusion of dissolved air in the liquid phase has been neglected in this implementation.
Transport of dissolved chemical species is assumed to take place in the liquid phase both by
advection and diffusion. Hence, the mass flux for a chemical species Cl is given by

FlCl ul ClCl ( Sl DlCl ) ClCl ;

ul k

krl

Pl l g

(10)

where u l is Darcy velocity of the liquid phase, ClCl is the chemical species concentration,

DlCl is its effective diffusion coefficient.


Since most chemical and physical properties are affected, to various degrees, by
temperature, it is important to account properly for the phenomena of temperature and heat
evolution over time in both the solid as well as the fluid phases of concrete as well as the
GFRP rod. The equation that describes heat transfer by convection and conduction reads
Ceq

T
. K eq T CL uT QH
t

(3)

267

The dependent variable is temperature, T. In the equation, Ceq denotes the effective
volumetric heat capacity; Keq defines the effective thermal conductivity; and CL is the
volumetric heat capacity of the moving fluid. The total fluid velocity, u, is a vector of
directional velocities u, v, and w for 3D systems. The terms on the right-hand side of the
equation, QH and denote a heat source. The equation coefficients and source terms can be
spatially and temporally varying, remain constant, depend on the temperature T, and link to
any other physics in the model. Thermodynamics suggests that systems that are in a nonequilibrium state will try to achieve thermodynamic equilibrium. However, typical
degradation reactions involved in high performance polymers, such as vinylester resins,
tend to proceed very slowly toward equilibrium, and concepts of chemical kinetics must be
invoked to characterize how fast the reaction is moving towards equilibrium conditions. Fig
3 shows the models ability to predict the synergetic effects of carbonation and chlorides
ingress in concrete. Knowing the drop of pH as a function of time is a particularly desirable
feature for GFRP given the susceptibility of e-glass fibers to highly alkaline environments.

Fig 1. Simulation of free and bound chloride and pH profiles after 10 years in the presence
and absence of carbonation.

3.2. Coupling of Chemical Attacks and Mechanical Properties


This section describes a thermodynamics based approach for coupling of mechanical
loading and chemical attacks in both concrete and GFRP. The formulation is based upon
the thermodynamics of open porous media composed of a skeleton and several fluid phases
occupying the porous space. A sketch of a general procedure for achieving this goal is
described in this section. The generalized Claussius-Duhem inequality, which expresses
locally the second law of thermodynamics is given by
1 2 0
(5)

268

with 1 being the intrinsic dissipation (irreversible behavior of the skeleton) and 2 is the
dissipation associated with chemical reactions. 1 can be expressed in the following form:

d
dT o

S
g mj
:

dt
T dt
m j

dm j

t
dt

(6)

where is the stress tensor, the strain tensor, S the entropy, T is the temperature, g mj is
the chemical potential (free enthalpy per unit mass of fluid phase j), m j denotes the mass of
fluid phase per unit of macroscopic representative volume element (rve), and is the free
energy of open elementary system. In the case of a reversible behavior, 0 and the
constitutive laws reduce to state equations

,S
and g mj
, j 1,..., N

T
m j

(7)

In the case of an irreversible behavior of the skeleton (matrix), these state equations are still
valid and the dissipation law becomes


0
t

(8)

Constitutive equations encompass the state equations and the complementary evolution
laws which describe the irreversibility of the matrix. This latter law is usually expressed in
terms of a relationship linking to . The case of dissipation due to chemical reactions
can be illustrated with the case of the following simple reaction
X Y

(9)

2 Am o 0 , where Am k ( g mX g mY ) is the affinity of the chemical reaction, o the


chemical rate, and k is a constant linking the reaction rate to the mass creation associated
with eq. (9). This latter expression of 2 holds irrespective of transport phenomena of the
reactant and product phase through the structure. By choosing a suitable free energy under
the form ( , T , mi , ) , i 1,..., N c it would be possible to account for the effect of
chemical attacks on the mechanical properties through the chemical reaction evolution as
represented by mass creation mi and degradation of the material under external loading
through the evolution of internal variables such as damage and/or plastic strains [5].
Similarly, it is perfectly possible to make the evolution of the transport properties depend
on the extent of both the chemical attacks and material degradation under mechanical
loading.

269

4. A SIMPLE EXAMPLE ILLUSTRATING THE USE OF THE PROPOSED


APPROACH
The main objective to this section is to illustrated the application of the proposed modeling
approach for the physicalchemical ageing of a GFRP rod. The model presented is based
on chemical attack of the inner glass fibre surface -by diffusing Sodium Hydroxide- and
subsequent decrease of interfacial matrix-fibre strength. Here, we will specifically focus on
the diffusion and subsequent glass fibre corrosion by Sodium Hydroxide.

4.1 Effect of Water


Both VE resins and glass fibers are known to be very resistant to chemical attack by water.
The interphase between the two, however, might be susceptible to some degradation due to
microblister formation at interface defects which always exist due to suboptimal wetting of
fibers by sizing. This is usually caused by the formation of a macroscopic solution as a
result of reaction or by intrinsic poor adhesion, giving rise to an osmotic pressure with
subsequent microblister formation. A growing blister could alter the dimensions of the
composite, giving rise to microcracks and leaching of water and composition material
through the blister surface or along the fibre. Whether a blister will grow or not depends on
the following factors:
- the generated osmotic pressure at some point in time
- the interfacial VE resin - sizing strength and
- the elastic properties of the composite.
The condition for growth can be determined by use of a Gibbs energy balance that contains
these three factors. If the blister can grow, the growth rate is usually determined by the
diffusion rate into the blister. Use of a mass balance and a diffusion equation, provide the
necessary tools to simulate the growth rate [7]. The term for the change in elastic energy for
the hemispherical blister is defined by:
2
p
dGelastic . p.z. .dA
3
E

(10)

The resulting total energy expression has a minimum for a certain blister size. This
minimum defines an equilibrium condition, which we express in terms of critical pressure:

pcrit

2. .E
d

(11)

Figure 2 shows that, even if the reaction equilibrium lies far to the right, the energy that is
required for blister expansion in this sort of composite is much too high to allow blister
growth initiated by water in such a cavity. Figure 7 clearly shows that the osmotic pressure
that could develop at the fibre-matrix interphase is much smaller than the critical pressure
needed to drive a blister growth. Hence, it is expected that no growth of any macroscopic
blister could take place. As a consequence, the volumetric change of the composite, due to
blister growth, is approximately zero.
270

Fig. 2. Pressure profiles for blister growth.

4.2. Effect of Sodium Hydroxide Ions and Mechanical Retention


In this section we will determine the mechanical retention of the GFRP rod by the
degradation of the fibre-VE resin interface. Assuming that the interfacial strength is a
direct function of the extent of chemical degradation reaction of Sodium Hydroxide on the
glass fibre surface, it is shown that at pH 12, the lifetime of the E Glass VE resin
composite exceeds a 100 years at ambient conditions of temperature and pressure. If, on the
other hand, pH is spiked to 14, the lifetime could drop as low as 40 years.

Fig. 3. Mechanical strength retention of GFRP at pH 14 and pH 12.


Deviations of accelerated tests from the findings of this analysis can be explained as
follows: At higher temperatures (typical of accelerated tests, around 80% of glass
271

transition) swelling stress exceeds the restrain stress of the fibrematrix interface and result
in volume change and subsequent appearance of macroscopic holes. Following that,
degradation rates increase to extreme values that bear no resemblance to real-life
circumstances.

5. CONCLUSIONS AND RECOMMENDATIONS


The following conclusions can be drawn:

Classical accelerated aging procedures have failed to provide reliable means for
predicting the service performance/life of GFRP reinforced concrete structures as has
been demonstrated by the ISIS GFRP durability project (no detectable degradation
after 10 years exposure to the concrete environment in service conditions).

To overcome the above limitations, it is necessary to build more robust models that
can describe more accurately the penetration of potentially harmful substances into
GFRP composites and their subsequent interaction with the resin and fiber. A reactive
transport approach where the thermodynamics and kinetics of reactive porous media
is used as a general framework for the integration of durability and structural analysis
in concrete has been proposed in this study.

This approach has the potential to allow a reliable prediction of the service life of
GFRP reinforced concrete structures through its ability to account for the following
major processes which govern the performance of GFRP reinforced concrete
structures: Accurate models for the prediction of water flow and transport of aqueous
chemical species in concrete and in the GFRP composite; Advanced mechanical
models for predicting the structural response (deformation, cracking, etc.) of both
materials (concrete and GFRP); Reliable degradation/corrosion models for evaluating
the extent of rebar degradation; An integrated approach to account for the dynamic
interaction of the above models.

At pH 12 the lifetime of the E Glass VE resin composite exceeds a 100 years at


ambient conditions of temperature and pressure. If, on the other hand, pH is spiked to
14, the lifetime could drop as low as 40 years.

More research is still needed to fine tune the various models described which would
lead to high fidelity forecasts over the service life of structures, more than 100 years
in advance.

6. ACKNOWLEDGEMENT
The authors gratefully acknowledge the financial support of ISIS Canada and NSERC.

272

7. REFERENCES
1.
2.
3.
4.
5.
6.

7.

Benmokrane, B. and El-Salakawy, E. 2007. Durability & Field Applications of Fiber


Reinforced Polymer (FRP) Composites for Construction. Proceedings of the third
CDCC Conference, 22-24 May, Quebec City, Quebec, Canada.
Nkurunziza, G. et al. Durability of GFRP bars: A critical review of the literature. Prog.
Struct. Engng Mater., 2005.
Boulfiza, M. and Munshi, S. 2006. A reactive transport model for carbonation and
chlorides ingress in concrete. ISTC06 Int. Symp. Theoretical Chemistry, 12-15 June.
Lichtner, P.C., Steefel, C.I., and Oelkers, E.H. 1997. Reactive Transport in Porous
Media. Reviews in Mineralogy,Vol. 34, Mineralogical Society of America,
Washington, DC.
Coussy, O. and Ulm, F.J. 1996. Creep and Plasticity due to chemo-mechanical
couplings. Archive of Applied Mechanics, 66: 523-535.
Katsuki, F. and Uomoto, T. 1995. Prediction of deterioration of FRP rods due to alkali
attack. Non-metallic (FRP) reinforcement for concrete structures, Non-Metallic (FRP)
Reinforcement for Concrete Structures, Proceedings of the 2nd International RILEM
Symposium (FRPRCS-2), Edited by L. Taerwe, F & FN EPSON, London, Great
Britain, pp. 82-89.
Walter, E. and Ashbee, K.H.G. 1982. Osmosis in composite materials, Composites, pp.
365-368.

273

274

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

VALIDATION OF DIGITAL IMAGE CORRELATION TECHNIQUE ON


CFRP AND SFRP WRAPPED CYLINDERS SUBJECTED TO FREEZETHAW ENVIRONMENTAL EXPOSURE
K. Abdelrahman1 and R. El-Hacha2
1
2

MSc Student, Dept. of Civil Engrg., University of Calgary, Canada


Associate Professor, Dept. of Civil Engrg., University of Calgary, Canada

ABSTRACT
This paper examines the behaviour of plain concrete cylinders wrapped with Fiber Reinforced
Polymer (FRP) exposed to freeze-thaw conditions. The environmental effect of freeze-thaw
deteriorates the concrete affecting the strength and ductility. Thus, a need for repair and
rehabilitation is established. One of the repair and rehabilitation methods is to wrap the
concrete with FRP materials. In this study, circular cylinders (150 300 mm) are wrapped by
two types of fibers; Carbon FRP (CFRP), and Steel FRP (SFRP). The cylinders are exposed to
471 cycles of freeze-thaw (+34oC to -34oC). The experimental program includes nine
cylinders divided into three groups. The first group consists of three control cylinders, the
second group consists of three cylinders wrapped with CFRP sheets and the third group
consists of three cylinders wrapped with SFRP sheets. This paper introduces a new technique
based on digital image correlation technique (DICT) that provides hoop strain readings over
the surface of FRP confined concrete. Results indicate that FRP sheets increase the strength
and ductility of wrapped concrete cylinders. The digital image correlation technique has
proven to be reliable and reasonably accurate to predict hoop strain readings from the surface
of FRP confined concrete.

1. INTRODUCTION
Fiber reinforced polymers (FRP) materials have replaced traditional materials for
strengthening concrete structures due to their light weight, high strength to weight ratio,
corrosion resistance. FRP materials are composed of high-strength fibers (carbon, glass and
aramid) embedded in a polymer matrix. Steel Fiber Reinforced Polymer (SFRP) is a new class
of composites recently introduced in the strengthening applications of concrete. SFRP is made
from ultra high strength twisted steel wires which is 11 times stronger than typical steel plate
(Hardwire, 2009). Several researches have shown that confinement of concrete columns using
FRP increases their strength and ductility. Environmental exposure of concrete has gained
significant research interest due to its crucial deterioration effect on concrete. Several
researchers have studied the effect of various severe environmental exposures on the strength
and ductility of FRP wrapped concrete columns (few are listed such as Karbhari and Eckel
275

1994, Hanna and Jones 1997, Soudki and Green 1997, Toutanji 1999, Toutanji and Deng
2002, Green et al., 2006, and El-Hacha et al., 2010).
Several researchers have implemented DICT to measure strains from the surface of materials.
White et al., (2003) measured the movement of soil patches using DICT. The study also
included a series of experimental validation to demonstrate the precision, accuracy and
resolution of DICT system. Bisby et al., (2007) implemented DICT to measure strain variation
from the surface of FRP confined concrete cylinders. The study included validation of DICT
on CFRP wrapped specimens which showed reasonable accuracy with conventional strain
gauge results. Bisby and Take (2009) studied the variation of hoop strains and axial strains
from the surface of FRP confined concrete. Results showed large variation in the axial and
hoop strain along the height of the cylinders. Bisby and Stratford (2010) studied the behaviour
of FRP wrapped cylinders with varying aspect ratio (diameter of 150 mm and aspect ratio of
2, 4 and 6). DICT was implemented to measure the hoop strain variation from the cylinders of
varying aspect ratio to provide an accurate and reliable strain efficiency of CFRP sheets.
Based on 525 hoop strain readings, the strain efficiency suggested for CFRP sheets was 0.5.
The objective of this paper is to study the effect of freeze-thaw environmental exposure on the
stress-strain curve, strength and ductility of concrete. This paper also examines the
implementation and validation of digital image correlation technique to measure hoop strain
readings from the surface of FRP confined concrete.

2. EXPERIMENTAL PROCEDURE
2.1 Materials and Test Matrix
A total of nine cylindrical plain concrete specimens were used each measuring 150 mm in
diameter and 300 mm in height. The concrete mix contained a water/cement ratio of 0.4, 1%
air content and a maximum aggregate size of 20 mm. The average 28-day compressive
strength of the concrete was 43.1 MPa. The concrete cylinders were wrapped with one layer of
unidirectional CFRP and SFRP sheets. The CFRP sheet used in this project was SikaWrap
Hex 230C made of unidirectional carbon fibre fabric with a thickness of 0.381 mm. According
to the manufacturer, the ultimate tensile strength and modulus of elasticity are 894 MPa and
65402 MPa, respectively, and the strain at failure is 1.33% (Sika, 2010). The SFRP sheet used
in this project was type 32-20-12 made of unidirectional brass coated ultra high strength
twisted steel wires with a thickness of 1.23 mm and a net cross-sectional area of 0.38
mm2/mm. According to the manufacturer, the ultimate tensile strength and effective modulus
for the SFRP sheet are 986 MPa and 66100 MPa, respectively, and the ultimate tensile strain
at failure is 1.5% (Hardwire, 2009).
The specimens are divided into three groups; Group I consists of three control unwrapped
cylinders, Group II consists of three cylinders wrapped with CFRP sheets, and Group III
consists of three cylinders wrapped with SFRP sheets. The specimens were exposed to freezethaw for 471 cycles. Each freeze-thaw cycle was 7 hours and 8 minutes between -34C to
+34C and a relative humidity of 75% for temperature above +20C.

276

2.2 Specimen Preparation and Test Set-Up


The concrete cylinders were cleaned and completely dried to ensure proper bonding of the
FRP sheets to the concrete surface. A thin layer of epoxy (Sikadur 330) was applied to the
concrete cylinders. The unidirectional sheets were then saturated with epoxy and applied
directly using wet lay-up method. Another layer of epoxy was added to the exterior of the FRP
sheets to ensure complete saturation of the fibers. All the specimens were wrapped with the
fibers in the hoop direction. To ensure full development of composite strength, an overlap
length of 100 mm was used. All cylinders were wrapped with one layer (300mm wide) of
CFRP sheet or SFRP sheet having almost the same axial stiffness (EACFRP= 7.48kN and
EASFRP= 7.51kN, where E is the modulus of elasticity of the FRP sheet and A is the
area/300mm width). All the specimens were loaded in uniaxial compression at a loading rate
of 10kN/s until failure using 2 MN Amslar hydraulic testing machine. Both ends of the
cylinders were grinded to ensure uniform distribution of loading. Hoop strains were measured
using two conventional foil strain gauges (120) with gauge length of 10 mm installed at
mid-height of the specimen located 180o from each other.

2.3 Digital Image Correlation Technique Instrumentation


The test setup of digital image correlation technique (DICT), instrumentation, and field of
view of the camera are as shown in Figure 1. The pictures were taken every 5 seconds using a
15.1 megapixel resolution Canon Digital Rebel camera. However, before the failure of the
cylinder, the camera was removed to protect it from being damaged by the catastrophic failure
of the cylinders. The images were analyzed using an image processing program called
geoPIV. The program works by defining particular sets of regions that are traced along
subsequent images. As shown in Figure 2, the particular region of interest were identified as
square patches of 64 64 pixels along the sides of the specimen to calculate the hoop strain.
The gauge length between two virtual points is approximately 150 mm. In order for the
technique to work, sufficient variation in the intensity and distribution of pixel colors must be
provided in order to trace the region of interest accurately. Thus, the cylinders were painted
with white paint and then sprayed randomly with black spray to provide the required texture.
More information about DICT can be found in White et al., (2003), Bisby et al., (2007), Bisby
and Take (2009), and Bisby and Stratford (2010).

Fig. 1. Instrumentation and field of view


(Bisby and Stratford, 2010)

277

Fig. 2. Virtual patches for DICT


for DICT

The DICT was implemented on all FRP confined cylinders; however, for CFRP-1 and CFRP2 specimens, the technique did not yield reasonable results since it was the first attempt of the
technique by the authors. Also, for specimen SFRP-2, the pictures were very blurry and the
DICT was not able to capture deformations. The method was successfully implemented on
CFRP-3, SFRP-1 and SFRP-3 specimens. This method highly depends on the users technique
during set-up and analysis of the data.

3. RESULTS AND DISCUSSION


3.1 Stress-Strain Behaviour
The stress-hoop strain behaviour of unconfined and confined cylinders is shown in Figure 3
and has been well established. The behaviour mainly features a bi-linear curve. The graph can
be divided into three regions (Lam and Teng, 2003); the first region of FRP confined concrete
follows the behaviour of unconfined concrete. In this region, the concrete is known to behave
elastically and the transverse (hoop) strain is linearly related to the longitudinal strain by
Poissons ratio (Youssef et al, 2007). Since the dilation of concrete is low in this stage, the
FRP confinement is usually inactive (or provides very little confinement). As the load
increases, the microcracks start to coalescence forming macro cracks which increases the
transverse strain of concrete thus activating the FRP wrap. The stress-strain curve starts to
soften forming a transition zone which is identified as the second region of FRP confined
stress-strain curve. In the third region, the concrete experiences dramatic increase in the
transverse strain which induces tensile forces on the FRP wrap. The FRP sheet resists the
expansion of concrete by providing a confinement pressure forcing the concrete to be in a triaxial state of stress. The FRP confinement is completely activated and increases proportional
to the applied load until failure (Youssef et al, 2007). This region of the stress strain is mainly
dependant on the cross-sectional geometry of confined concrete and the amount of FRP
confinement (Lam and Teng, 2003).

Axial Concrete Stress (MPa)

120
SFRP - 1

SFRP - 3

SFRP - 2

100
80
CFRP - 1

60

CFRP - 2

CFRP - 3
C-2

40

C-1

C-3

20
0
0

2000

4000

6000

8000

10000 12000 14000 16000

Hoop Concrete Strain ()

Fig. 3. Stress-hoop strain behaviour of unconfined and confined concrete.

278

Table 1 provides the unconfined strength (fc), maximum (i.e. peak) confinement strength
(fcc), ultimate confinement strength (fcu), hoop strain corresponding to maximum
confinement strength (cc) and the strain corresponding to the ultimate confinement strength
(cu) for all nine specimens. It can be seen that all confined specimens experienced significant
increase in hoop strain. CFRP wrapped specimens showed percentage increase in hoop strain
ranging from 290% to 450% compared to unwrapped specimens. SFRP wrapped specimens
showed percentage increase in hoop strain ranging from 400% to 650% compared to
unwrapped specimens. Based on the average strains from each group, SFRP wrapped
specimens showed an increase in strength of 34% compared to CFRP wrapped specimens.
Although the cylinders were wrapped with the same axial stiffness of both FRP sheets, the
rupture strain of SFRP is greater than the rupture strain of CFRP sheets which explains the
enhanced performance of hoop strain and maximum strength for SFRP wrapped specimens
over CFRP wrapped specimens.
It can be seen from Table 1 that the freeze-thaw exposure had a positive effect on the
compressive strength of unwrapped concrete cylinders. The average compressive strength
improved from 43.1 MPa to 62 MPa with an increase of 44 %. This could be related to the fact
that the proper curing and exposure of heat and moisture inside the chamber attributed to the
improvement in strength. Similar behaviour was also noticed by Toutanji (1999). CFRP
wrapped specimens had an insignificant increase in strength by an average of 3 % compared
to unwrapped specimens. This insignificant increase in strength is because the unconfined
concrete also experienced an increase in strength. For specimen CFRP-1, it can be noticed that
the maximum strength is different than the ultimate strength. The stress-strain curve features a
post-peak descending curve where the compressive strength is reached before the rupture of
FRP. However, since the ultimate confined stress fcu is still greater than the unconfined
strength fc, the cylinder is still classified as sufficiently confined (Lam and Teng, 2003). All
SFRP wrapped specimens showed a significant increase in strength by an average increase of
70% and 66% compared to unwrapped specimens and CFRP wrapped specimens,
respectively.

FRP
Unwrapped
(control)
CFRP
sheet
SFRP sheet

ID
C-1
C-2
C-3
CFRP-1
CFRP-2
CFRP-3
SFRP-1
SFRP-2
SFRP-3

Table 1. Results from compression tests


fc (MPa)
fcc (MPa)
fcu (MPa)
64.2
59.9
62.1
68.5
65.0
62.3
60.5
105.6
107.0
106.4
-

279

cc ()
1564
2278
453
2382
-

cu ()
8566
11641
7597
10194
14073
11783

3.2 DICT Analysis


Figures 4, 5 and 6 compare DICT analysis with experimental results from conventional strain
gauges until the point where the camera was removed. For this analysis, virtual optical strains
at mid-height of the specimen located at the same height as the conventional strain gauge was
used. It can be clearly seen from the figures that DICT was able to capture accurately the
stress-strain behaviour of confined concrete when compared to experimental results from
conventional strain gauges. For all the three specimens which used DICT, the virtual strain
readings had a percentage error of less than 10 % compared to conventional strain gauges.
These figures also show that for data points below the unconfined concrete strength, the data
are slightly scattered because the strains in this region are very small. However, in the third
region of the curve where FRP wrap is active and the concrete experiences large hoop strains,
DICT seems to capture the trend accurately. This is fairly reasonable since the major concern
is to measure large strains after the FRP wrap is active.
60
50

Axial Concrete Stress (MPa)

Axial Concrete Stress (MPa)

80

40
30
20
10

Strain Gauge
Digital Technique

70
60
50
40
30
20
Strain Gauge
Digital Technique

10

500

1000

1500

2000

2500

3000

500

Hoop Concrete Strain ()

1000

1500

2000

Fig. 4. DICT and strain gauge for CFRP-3


Height of Concrete Cylinder (mm)

Axial Concrete Stress (MPa)

3000

Fig. 5. DICT and strain gauge for SFRP-1


300

90
80
70
60
50
40
30
20
Strain Gauge
Digital Technique

10

2500

Hoop Concrete Strain ()

45 Mpa from DICT


55 MPa from DICT
57 MPa from DICT
58.8 MPa from DICT
45 MPa from Strain Guage
55 MPa from Strain Gauge
57 MPa from Strain Gauge
58.8 MPa from Strain Gauge

250
200
150
100
50
0

0
0

1000

2000

3000

4000

5000

Hoop Concrete Strain ()

6000

Fig. 6. DICT and strain gauge for SFRP-3

7000

1000

2000

3000

Hoop Concrete Strain ()

4000

Fig. 7. Strain variation for CFRP-3

280

5000

300

250
200
150
60 MPa from DICT
73 MPa from DICT
81.7 MPa from DICT
60 MPa from Strain Gauge
73 MPa from Strain Gauge
81.3 MPa from Strain Gauge

100
50
0
0

1000

2000

3000

4000

Height of Concrete Cylinder (mm)

Height of Concrete Cylinder (mm)

300

70 MPa from Strain Gauge


80 MPa from Strain Gauge
85 MPa from Strain Gauge
90 MPa from Strain Gauge
70 MPa from DICT
80 MPa from DICT
85 MPa from DICT
90 MPa from DICT

250
200
150
100

5000

50
0
0

Hoop Concrete Strain ()

1000 2000 3000 4000 5000 6000 7000 8000 9000 10000

Hoop Concrete Strain ()

Fig. 8: Strain variation for SFRP-1

Fig. 9: Strain variation for SFRP-3

Since DICT measures strains from virtual optical points on the cylinder, this method can now
be used to measure hoop strain variations over the height of the cylinder. Figures 7, 8, and 9
provide the hoop strain variation over the height of the cylinder. The top and bottom 50 mm
was excluded from the analysis since they are affected by friction induced by the platens.
These figures show great variation of hoop strains over the height of the cylinder. Although
there is no unique pattern that can be recognized, the patterns for each cylinder seem to
increase consistently. The figures also compare hoop strain results from the conventional
strain gauge at mid-height of the specimens with DICT at several load increments. It can be
clearly seen how close the readings are from the DICT compared to the conventional strain
gauge. Similar behaviour of the strain variations was also found by Bisby et al., (2007), Bisby
and Take (2009), and Bisby and Stratford (2010). Based on the comparison of the above
figures between DICT and conventional strain gauge data, it is reasonable to state that DICT
is a reliable and reasonably accurate method to measure strains from the surface of FRP
confined concrete. It is important to note that DICT provides a great insight into the variation
of strains over the surface of FRP confined concrete which was impossible to achieve using
conventional strain gauges. This variation simply states that the reading of conventional strain
gauge depends greatly on the location of the strain gauge along the height of the cylinder. It
has been also reported by Bisby and Take (2009) that the strain variation occurs
circumferentially along the perimeter of the cylinder.

4. CONCLUSIONS
From the compression tests of the CFRP and SFRP wrapped cylinders, the following
conclusions were drawn:
1. CFRP and SFRP sheets enhanced significantly the hoop strain of concrete and axial
strength. CFRP specimens showed a percentage increase in hoop strain ranging from
290% to 450% compared to unwrapped specimens whereas SFRP wrapped specimens
showed percentage increase in hoop strain ranging from 400% to 650% compared to
unwrapped specimens.
2. CFRP wrapped specimens had an insignificant increase in strength by an average of 3%
compared to unwrapped specimens whereas SFRP wrapped specimens showed a
281

significant increase in strength by an average increase of 70% and 66% compared to


unwrapped specimens and CFRP wrapped specimens, respectively.
3. Digital Image Correlation Technique (DICT) accurately captured the stress-strain curve
specifically in the third region where the FRP wrap is active when compared to the results
from conventional strain gauges.
4. DICT was able to capture the strain variations over the height of the cylinders which
provided very accurate results when compared to the conventional strain gauge at the same
location and specific load.

5. ACKNOWLEDGEMENTS
The authors would like to acknowledge the assistance of Dr. Luke Bisby to implement the
DICT, and Dr. Andy Take from Queens University for providing the geoPIV program to
perform the analysis of the DICT. The authors would like to express their gratitude to
Hardwire for providing the SFRP and Sika Canada for providing the epoxy and CFRP.

6. REFERENCES
1.

Raupach, M. 2006. Concrete Repair according to the new European Standard EN 1504.
(M. Alexander, H.-D. Beushausen, F. Dehn, & P. Moyo, Eds.) Concrete Repair,
Rehabilitation, and Retrofitting, 6-8.
2. Hardwrie. 2009. Hardwire Composite Armor System. (http://www.hardwirellc.com).
3. Karbhari, V. and Eckel II, D. 1994. Effect of Cold Regions Climate on Composite
Jacketed Concrete Column. Journal of Cold Regions Engineering , 8 (3): 73-86.
4. Soudki, K.A. and Green, M.F. 1997. Freeze-Thaw Durability of Compression Members
Strengthened by Carbon Fibre Wrapping. Concrete International, 19(8): 64-67.
5. Hanna, S., and Jones, R., 1997. Composite Wraps For Ageing Infrastructure: Concrete
Columns. Composite Structures. 38 (1-4): 57-64.
6. Toutanji, H. 1999. Durability Characteristics of Concrete Columns Confined with
Advanced Composite Materials . Composites Structures , 44(2-3): 155-161.
7. Toutanji, H., and Deng, Y. 2002. Strength and Durability Performance of Concrete
Axially Loaded Members Confined with AFRP Composite Sheets. Composites: Part B
33: 255-261.
8. Green, M.F., Bisby, L.A., Fam, A.Z., and Kodur, V.K.R. 2006. FRP Confined Concrete
Columns: Behaviour under Extreme Conditions. Cement and Concrete Composites, 28:
928-937.
9. El-Hacha, R., Green, M., and Wight, G.R. 2010. Effect of Severe Environmental
Exposure on CFRP Wrapped Concrete Columns. Journal of Composites for Construction,
14(1): 83-93.
10. Sika.
2010.
Sikawrap
Hex
230C.
Product
Technical
Data
Sheet,
(http://www.sikaconstruction.com/tds-cpd-SikaWrapHex230C-us.pdf), Feb. 21.
11. White, D., Take, W., and Bolton, M. 2003. Soil Deformation Measurement using Particle
Image Velocity (PIV) and Photogrammetry. Gotechnique , 53 (7): 619-631.
12. Bisby, L., Take, W., and Bolton, M. 2007. Quantifying Strain Variation in FRP Confined
Concrete using Digital Image Analysis. Proceedings of the 1st Asia-Pacific Conference
on FRP in Structures, (APFIS 2007), Hong Kong, China, 12-14 December, pp. 599-604.

282

13. Bisby, L. and Take, W. 2009. Strain Localisations in FRP Confined Concrete: New
Insights. Structures and Buildings , 165 (5): 301-309.
14. Bisby, L. and Stratford, T. 2010. The Ultimate Condition of FRP Confined Concrete
Columns: New Experimental Observations & Insight. Proceedings of the 5th International
Conference on FRP Composites in Civil Engineering (CICE 2010). Beijing, China, 27-29
September, pp. 599-602
15. Lam, L. and Teng, J. 2003. Design-oriented Stress-Strain Model for FRP-Confined
Concrete. Construction and Building Materials, 17: 471-489.
16. Youssef, M., Feng, M., and Mosallam, A. 2007. Stress-Strain Model for Concrete
Confined by FRP Composites. Composites: Part B, 38(5-6): 614-628.

283

284

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

EXPERIMENTAL STUDY OF THE INFLUENCE OF MATRIX


SYSTEM ON PULTRUDED GFRP COMPOSITES CREEP
BEHAVIOUR UNDER FLEXURAL LOADING
S. Barboura1, N. Kotelnikova-Weiler2, J.-F. Caron3 and O. Baverel4
1

Post-Doctoral Reseaurcher, Laboratoire Navier, Ecole des Ponts ParisTech, France


PhD student, Laboratoire Navier, Ecole des Ponts ParisTech, France
3
Professor, Laboratoire Navier, Ecole des Ponts ParisTech, France
4
Professor, Laboratoire Navier, Ecole des Ponts ParisTech, France
2

ABSTRACT
Behaviour of glass fiber reinforced pultruded composites with three different resins
(polyester and two vinylesters) was studied under flexural sustained loading. Three
different geometries of samples were used: rectangular bars, thin rods and tubes. Bars and
rods were tested in laboratory conditions whereas tubes were part of an outside gridshell
structure for 3 years before dismantling. For lab samples different stress levels were applied
while creep deformation was recorded. Their strength evolution was also assessed. Residual
deformation was measured for the gridshell tubes. A widely used model integrating
viscoelastic behaviour of the material was built in order to predict the material long-term
physical properties evolution. The final purpose is to integrate this kind of models in
mechanical calculations in order to ease the use of FRP materials in structural applications.

1. INTRODUCTION
Development of industrialised production of composite materials such as pultrusion
technique encourages the use of composite materials in a wide range of applications in
various fields from aeronautical to civil engineering. In the latter, composite materials are
used for reinforcement and rehabilitation of existing concrete, masonry or steel structures.
Works of Ascione et al. 2008 [1] and Tavares et al. 2002 [2] as well as Al-Salloum and AlMussallam 2007 [3] show their application for the rehabilitation of concrete structures, and
Seica and Packer in 2007 [4] discuss their use for steel tubes reinforcement. New
construction typologies are developed to explore mechanical potential of composite
materials. Pultruded composites bring together high mechanical properties and
industrialized production which makes them a competitive solution for entirely composite
structures. In the Institut Navier two structures of that kind have been developed. The first
one is a footbridge made of sandwich composite material (Caron et al. 2009 [5]). The
second one is a gridshell structure made of pultruded composite tubes (Douthe et al. 2006

285

[6], 2010 [7]). In theses two applications the rigidity of the structure is obtained through
bending of the composite material. Thus the material is subject to sustained flexural loading
during the structures entire service life. It is therefore essential to investigate creep and
rupture behaviour of these materials under maintained load. One of the prototypes
(presented in figure 1) of a gridshell structures remained outside for three years has been
dismantled. In this study we investigate the residual deformation of the tubes that composed
this structure.

Fig. 1. Gridshell prototype of the Navier Institut.


FRPs creep behaviour and strength evolution depends on the loading history,
environmental conditions and the constituents of the material: fibers, matrix and interface.
Foreseeing the time evolution of a structural components mechanical characteristics is
essential during the design stage of a project (Richard and Perreux 2001 [8], Choi and Yuan
2003 [9]). To effectively integrate the variation of the materials mechanical properties in
structural calculations, a mathematical model needs to be developed. This model shall link
modifications in materials response to the changes in loading conditions (loading being
mechanical or hygro-thermal). Findelys empirical model for linear viscoelastic composites
is one of the most popular models to be used for pultruded composites. In this study
materials will be characterized using this model to ease the comparison with other materials
and integration into existing predictive models.
Many authors have studied creep of pultruded structures. To cite a few: Bank and
Mossallam in 1992 [10] studied viscoelastic behaviour of an H-shaped frame made of
pultruded glass-fiber reinforced vinylester resin. The frame is subject to the quarter of its
ultimate flexural load. Findey model is then used to predict the softening of the structure.
Other researchers such as Mcclure and Mohammadi 1995 [11], Scott and Zureick 1998
[12], Choi and Yuan, Shao and Shanmugan 2006 [13] have investigated pultruded
structures creep. Most of the experiments lasted approximately a year and Findley model is
used to characterise materials behaviour.
The aim of this work is to complete theses studies and provide a characterisation and
comparison of three different composite systems made of E-glass fibers and three distinct
resins: isophtalic unstaturated polyester, a standard resin in pultrusion industry, and two
vinylester resins (Derakane 441 and Derakane 470). Creep behaviour and reduction of
residual stress are both studied in this work. Viscoelastic behaviour of composite materials
comes essentially from the viscoelasticity of the matrix composing the material. A detailed
description of the micromechanical processes can be found in the paper entitled
Phenomenological and experimental study of pultruded UD FRP composites static and
deferred rupture under combined flexural, compressive and torsion loading by the same
286

authors in this conference. One can identify the three phenomena of interest in the study of
deferred behaviour of a composite: viscoelastic creep deformation, strength reduction and
deferred rupture. In the following we will present our experimental results for the study of
these phenomena. In the first part we will present laboratory experiments: creep of thin rods
and rectangular bars in bending. We will also present a few results obtained on tubes part of
the outside gridshell structure exposed to variable (and uncontrolled) environmental
conditions during three years. In the second part we will present the study of strength
reduction in bending.

2. CREEP IN BENDING
2.1 Test Samples
The samples tested are pultruded composite profiles. They are unidirectional glass-fiber
reinforced thermosetting matrices. Three different matrices are used: polyester and two
vinylesters (Derakane 441 and high-performance Derakane 470). The profiles have three
different geometries: thin rods, tubes and rectangular bars. The products properties are
presented below in terms of composition (resins and fibers properties: Table 1) geometry
of the samples (Table 2), mechanical properties of the rod samples (Table 3), bar samples
(Table 4) and tube samples (Table 5). Resin properties are given by the producer. Tg stands
for temperature of glass transition. Tg for the composites was measured by dynamic
mechanical spectrometry at the LCPC laboratory. Mechanical properties of the composite
samples were determined experimentally at our laboratory using the same devices as for the
long-term testing. These devices are described in the next page.
Table 1. Mechanical properties of the fibers and resins used in the rods and bars samples.
Resin/Properties
rupt
rupt EL
Tg composite
Tg resin
Polyester
D441
D470
Fibres/Properties

120 MPa 3.7% 3.9 GPa 174C


125C
145 MPa 5-6% 3.4 GPa 159C
135C
160 MPa 3-4% 3.8 GPa 197C
145C
Diameter
Youngs modulus Poissons ratio

E-Glass fibers

17-24 m

72Gpa

0.2

Table 2. Geometry of the samples.


Geometry Rod
Tube
Rectangular bar
Dimensions 5mm diameter 42mm external diameter 40mm large
35mm internal diameter 16mm thick

Properties of rod
samples
Glass/polyester
Glass/D441
Glass/D470

Table 3. Mechanical properties of rod samples.


E
rupt
rupt
Standard
(GPa) (%)
(MPa)
deviation for rupt
49.24
2.328
1320
4.44 %
43.37
1.865
894
3.14 %
44.34
2.314
1197
4.65 %

287

Fibre volume
fraction
52%
47%
47%

Rectangular
bar
Glass/D441
Glass/D470

Table 4. Mechanical properties of bar samples.


E
rupt
rupt
Relative deviation
(GPa) (%)
(MPa)
for rupt
37.7
1.616
602.13
3.77 %
36.10
2.311
846.4
3.17 %

Fibre volume
fraction
47%
47%

Table 5. Mechanical properties of tube samples


Tubes
E (GPa) rupt (MPa)
Glass/polyester 26.7
350
The tubes tested have different material constituents and mechanical properties from the
glass/polyester systems used in rods and rectangular bars. They are part of a previous
project involving different materials. To determine their properties, specific flexural testing
was carried out therefore characteristics presented in the table (Table 5) above correspond
to flexural loading.
2.2 Experimental Devices
In this section we will present testing devices used to investigate the creep behaviour of
pultruded composites. Tests were carried out on laboratory samples and on samples
obtained after the dismantling of an outdoor structure, a gridshell shown in the figure 1.

a)
b)
Fig. 2. Testing machines used for laboratory bending experiments. a) For bar samples; b)
For rod samples.
The figures above show the testing machines used for laboratory experiments: a standard
MTS testing machine (Fig. 2a)) with a capacity of 100kN and an Instron with 10kN
capacity (Fig. 2b)). Samples were tested in four-points bending under constant force.
Different load levels were applied. Rectangular bars samples were tested using the MTS
machine and the rods using Instron. Strain is measured using strain gauges or calculated
using crosspiece displacement.
L0
h

L/n
L

Fig. 3. The schematic of a four-points bending test.


288

Table 6. Testing characteristics for the four-points bending test.


Testing characteristics
L
h
n
Rectangular bars
600 mm 16 mm 2
Rods with Derakane resins 100 mm 5 mm 3
Rods with polyester resin 150 mm 5 mm 3
The gridshell structure was built using polyester tubes. It is represented in figure 1. The
structure remained outside covered with a canvas sheet for approximately three years. It
was then dismantled and the residual curvature was assessed using a three point device
described below (Fig. 4).

Displacement
sensor
580 mm
290 mm

Fig. 4. Schematic of a three-point device for measuring of the residual curvature


Only a few points were selected to evaluate the creep deformation of the tubes. These
points correspond to the most loaded sections of the structure. Therefore the initial loading
is very close for all of the points chosen: it is approximately 30% of the ultimate initial
strength.

2.3 Results and Findley Model


A constant force (constant stress) is maintained during creep testing of rectangular bars and
rods. The strain is assessed in time using a strain gauge (for static tests) or calculated from
the crosspiece displacement (for creep tests). The total strain (t) can be expressed as a sum
of the instantaneous strain 0 that accompanies the applied stress immediately at t0 and the
deferred strain d(t). Results presented here were obtained for creep tests in bending at room
temperature of 22C. The duration of the applied load was approximately a week. Four
different stress levels were applied to the samples: 40, 50, 60 and 70% of the initial ultimate
stress. Data is captured at variable frequency: every second at the beginning of the
experiment and once every 15 minutes after a few hours.
Following figures (Fig. 5) show the results for rectangular bars of glass/Derakane 441 and
glass/Derakane 470 composites for different applied stress levels. For glass/Derakane 441
at 70% level the rupture was very sudden and occurred after 25hours. Unfortunately we
could not capture the evolution of the deformation rate preceding the rupture. The
glass/Derkane 470 specimen at the same stress level resisted for over 5 days. Though after
this period leaps in the specimens strain can be observed being the result of fiber bundles
brutal rupture in tension on the lower side of the specimen. For lower stress levels,
289

whatever the material, strain tends to level off characteristic of primary and secondary
creep. Using experimental data, we could calculate creep deformation rate using the
following formula:
p (%)

(t ) 0
100
0

(1)

Figure 6 presents the evolution of the creep deformation rate in comparison to the initial
deformation as a function of time. One notices that after a 24h period more than 50% of the
final deformation, obtained after 5 days of testing, is already achieved. An average raise of
2 to 3% of the deformation in comparison to the initial one is recorded on these two
pultruded materials for loadings below 60%. The glass/Derakane 470 composite presents a
better resistance to creep deformation than the glass/Derakane 441, and the curve almost
levels off more rapidly for glass/Derakane 470 with loadings below 70% (see figure 6) .

a)
b)
Fig. 5. Creep deformation of bar samples of glass/D441 (a) and glass/D470 (b) for different
stress levels.

a)
b)
Fig. 6. Evaluation of the creep deformation rate for bar samples of glass/D441 (a) and
glass/D470 (b) at different stress levels.

290

a)
b)
Fig. 7. Evaluation of the creep deformation rate for cylindrical samples of glass/D441 (a)
and glass/D470 (b) at different stress levels
Figure 7 presents the results obtained with the rod samples of glass/Derakane 441 and
glass/Derakane 470 for various bending moments, which are kept constant in time. The
creep deformation rate evolution, calculated using equation (1), is around 3% after 5
minutes of testing the two composite rods for loadings below 55% of the initial ultimate
stress. On this same figure, one can observe the tertiary creep followed by rupture for
loading of 70%. This tertiary creep is characterized by a high creep speed and an inflexion
point. Rupture occurs far more rapidly on samples of glass/Derakane 470 (after 1.5
minutes) than on samples of glass/Derakane 441 (40 minutes), which corresponds to the
opposite trend from what was observed for the rectangular bar samples.
Creep behaviour is therefore dependent on the structure and scale of the composite sample.
What can be clearly observed on these tests is that tertiary creep which leads to a deferred
rupture is strongly dependent on the geometry parameters. Even though Derakane 470
composite system globally seems to have a better resistance to creep, rod samples of this
material will break before the rod samples of the D441 composite and the trend is inversed
for the bar samples.
Measurements of the residual flexural deformation on the polyester gridshell tubes gave us
the following result (Fig. 8).

Fig. 8. Residual stress (in % of initial strength) versus initial stress (in % of the initial
strength) after three years of continuous bending load.

291

It can be seen that over the range of creep bending load between 15% and 40% of initial
strength, the residual stress after a period of three years tends to increase but with a low
rate: when initial load goes from 15% to 40%, the residual stress increases from 13,7% to
14,5% (additional 0,78%).

2.4 Creep Model of Pultruded Composites


To evaluate the creep deformation of the GFRP composites tested, in terms of flexion, one
must have a simple and suitable model for unidirectional composites. As stated before, the
Findley power law model is the most commonly used for predicting the long term
behaviour of pultruded composites. This model has the advantage of requiring only creep
results to determine the different terms of its equation. The general equation of the Findley
model is given by:
t
(t ) 0 m
t0

and

n m t
r (t )

0 t0 t0

n 1

(2)

where (t) is time depending total creep deformation, 0 the initial deformation, which is a
function of the initial stress applied, m is the creep coefficient depending on the stress, n a
constant independent from the loading level, t the time after the creep loading, t0 the instant
when the creep loading is met and r (t ) the creep deformation relative speed. Generally,

Equation 2 describes the deformation for a material subject to a constant loading. 0 , a time
independent constant, and m are often calculated using a moderate-constraint hyperbolic
function (McClure and al. 195 [12]). These hyperbolic functions depend on the loading and
can be written as:


0 0' sinh
m m ' sinh

and
m
(3)
'
where 0 is the instantaneous deformation associated to a reference loading

, m ' is the

creep parameter at the reference loading m , and the loading level. These different
parameters are all constants, independent from time and temperature, which are determined
from the creep tests results at various loading levels. The parameters m and n can be
determined using test results. By re-organizing equation (2) and applying the Napierian
logarithm, the non-linear portion becomes linear:
t
Log ( (t ) 0 ) Log (m) n Log
t0

(4)

Some results of this model are presented on the figure 9. On these graphics, the
experimental results are superimposed to the one estimated by the model. The models
parameters are determined for each material and each loading (Table 7). The coefficient n
characterizes the viscoelastic effect of the material. Thus, the greater n is, the more
292

pronounced the viscoelastic effect becomes. It is generally acceptable this exponent be a


constant parameter for the material. In anterior creep studies of pultruded composites, two
ranges of values have been distinguished : a first group of 0.22 0.25 (McClure and
Mohammadi 1995 [12], Scott and Zureick 1998 [13]) and a second around 0.33 (Bank and
Mossallam 1992 [11]). Although the values of n are strongly dependent on the material, it
is difficult to conduct a direct comparison. It is however noticeable to know its order of
magnitude. From Table 7, one can observe the materials studied belong to the second
aforementioned group.
Bar
Loading
m
n
0 (%)
Erreur (%)
Rods
Loading
m
n
0 (%)

Table 7. Findley model parameters for bar and rod samples


Glass/D441
Glass/D470
40%
50%
60%
70%
50%
60%
-3
-3
-3
-3
-3
0,526 10
1,83 10
3,88 10
3,3 10
1,66 10
12 10-3
0,25
0,28
0,22
0,264
0,257
0,145
0,67
0,87
0,98
1,17
1,22
1,53
0,3
3
0,3
0,3
0,45
0,62

Glass/D441
39%
6,9055 10-5
0,2289
0,9527

48%
2,9325 10-4
0,1125
1,3167

55%
1,7409 10-4
0,2028
1,3979

Glass/D470
42%
1,8031 10-4
0,1576
1,3586

51%
3,9453 10-4
0,1392
1,7026

70%
3,69 10-3
0,28
1,8
0,12
56%
2,1258 10-4
0,2265
1,6092

a)
b)
Fig. 9. Superposition of experimental results and Findley model for a stress level of 50%
for rod samples: (a) creep in bending of glass/D441, (b) creep in bending of glass/D470

a)
b)
Fig. 10. Findley model of the relative creep deformation rate of rof samples of glass/D441
(a) and glass/D470 (b)
293

Figure 10 presents the relative deformation ratio for the rod samples, calculated using the
Findley model. The overall model behaviour is satisfactory for duplicating the experimental
results (an error lower than 10% between the experimental points and the Findley
calculated ones). An estimation of the relative creep speed for the different composite
materials after 10.000 minutes is presented in Table 8. One can then compare the creep
tendency of the two studied materials, glass/Derakane 441 and glass/Derakane 470.
Glass/Derakane 470 samples globally present a better creep resistance. However, these
results seem to indicate the creep speed increases more rapidly with the loading level for
glass/Derakane 470 rod samples than for glass/Derakane 441. Thus, for a glass/Derakane
470 rod sample, this represents an increase of 0.32% per day of the creep speed between
40% and 60% loading, where it is limited to 0.02% per day for similar loadings with
glass/Derakane 441. This speed evolution explains the fast rupture of the glass/Derakane
470 samples (after 1.5 minutes) in comparison with the glass/Derakane 441 samples
(rupture after 40 minutes) under a 70% loading of its initial ultimate stress.
For the bar samples, one observes an increase of the creep speed for the two materials of
0.02% per day for the same loading conditions. On the other hand, the creep rupture of the
rectangular sample glass/Derakane 441 is much quicker (25 hours) than the one of the
glass/Derakane 470 sample under a loading of 70%, which seems to resists up to 5 days. It
should be noted however that the creep test have being carried out for short and various
time lengths for different loadings, the trends may invert in the long term.
Table 8. Evaluation of long-term relative creep strain of tested composites
Creep rate after 10.000 minutes
Glass/poly Glass/D441
Glass/D470
(in %/day)
Loading
50%
40% 50% 60% 40% 50% 60%
0,31 0,12 0,33 0,17 0,20 0,49
Rods
0,02 0,07 0,09 0,04 0,06
Bar
-

3. RESIDUAL STRENGTH OF PULTRUDED COMPOSITES UNDER CONSTANT


BENDING
3.1 Testing Devices
By using different tests, we aim to underline the phenomenon of deferred rupture, the
characteristic times and the resistance drop associated with it and the influence of the
matrix constituents. The samples will be subjected to different loading levels for
predetermined periods of time. The samples are then tested for resistance. The loading is
done by attaching the two ends of the sample to an anchor system that allow free rotation
(see figure 11a)). One obtains a post-buckling elastica (see figure 11b)). Testing is done
in bending, for a more detailed study of different rupture modes in various loading
condition, see the article entitled Phenomenological and experimental study of pultruded
Ud frp composites static and deferred rupture under combined flexural, compressive and
torsion loading by the same authors in this conference.

294

a)
b)
Fig. 11. a) Anchor system for the "elastica" samples; b) Final geometry of the samples
The loading obtained varies along the sample. The moment is null at the samples anchor
and maximum in its center. One imposes the maximal central loading simply by choosing
the length of the sample.
Each section from the elastica will undergo a different loading history. To measure the
residual resistance time evolution, it is necessary to make as many elasticas with the
appropriate form as there are points required. One simply has to unload a sample after a
given period of time, to cut the sections (which loading history is known) and test their
flexural resistance.
5 elasticas were built for each composite material, among which 4 are unloaded after 3, 7
21 and 35 days. The last one is kept loaded until rupture occurs. The maximum loadings
imposed to the different materials are 60, 65 and 90% of the initial ultimate stress of each
material (glass/Derakane 441, glass/Derakane 470 and glass/polyester). These supplies
samples sections that where loaded respectively between 11 to 60%, 13 to 65% and 25 to
89% for each 3 material.
3.2 Results
The sections cut from the elasticas are tested by four point bending until rupture occurs
with the same experimental device as for the creep tests. Here however, rubber skids are
added under the inner supports in order to protect the sample from being crushed when the
stress becomes important (just before rupture). The force applied to the sample is recorded.
The rupture stress and the residual resistance are deduced from these measurements. For the
samples that are not unloaded and left until break occurs, the rupture is marked as soon as
the first fiber is broken (see figure 12).

Fig. 12. Rupture observed on a sample of glass/polyester


The results obtained for the different samples do not allow, for the time period taken into
account, to clearly identify the differences in terms of resistance drop between the samples
that underwent different loadings. Indeed, the physical properties and characteristics of
each sample (geometry and material) lead to very scattered results, and the resistance drop
under different loadings are in the same order of magnitude.

295

One calculates the mean resistance obtained at a given instant. This mean value is then
normalized using the resistance of a brand new material for the different samples of the
same elastica. This leads to the experimental graphics traced on figure 13 where the 3
materials are compared.
It is difficult to compare the three materials on such short time periods. Although it can be
noted that the initial resistance drop is very close for the three materials (the representative
points for the three materials are very grouped until 400h). Then, the material differentiates
from one another and large gaps between glass/Derakane 441 and glass/polyester is
observed after 1600h: the resistance of the glass/Derakane is down to 60% of its initial
value whereas the glass/polyester only dropped to 90%.

a)
b)
Fig. 13. a) Evolution of flexural strength of different samples; b) Rupture of an "elastica"
sample of glass/D470

4. CONCLUSION
This article presents an experimental test campaign for the study of creep phenomenon in
different pultruded composite materials using various geometries. The tests are conducted
during almost a week. Several loadings are defined. The Findley model was successfully
applied to reproduce the test results. The first conclusions on the composite creep behaviour
can be drafted.
The glass/Derakane 470 composite shows a better strength to creep. The creep speed
estimation allows to confirm this result. Creep tests by 4 points flexion for longer periods
(one year) are foreseen. To achieve this, a climatic chamber has been prepared to receive a
4 point flexion frame capable of testing 18 samples at a time.
A specific experimental protocol easing the deferred tests was also implemented. The tests,
performed for periods lasting up to 3 months, permitted the observation of the deferred
rupture as well as the flexion resistance drop. The mechanism leading to the deferred
rupture is closely linked to the behaviour of the matrices.
As such, several research leads allowing different solicitations of the matrix are foreseen, to
further investigate its role: high temperature tests to fasten the matrix creep and combined
torsion-flexion tests, increasing the mechanical solicitation of the matrix.
296

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

Ascione, F., Berardi, V.P., Feob, L., and Giordano,A. 2008. An Experimental Study on
the Long-Term Behavior of CFRP Pultruded Laminates Suitable to Concrete Structures
Rehabilitation. Composites: Part B, 39: 1147-1150.
Tavares, C.M.L., Ribeiro, M.C.S., Ferreira, A.J.M., and Guedes, R.M. 2002. Creep
Behaviour of FRP-Reinforced Polymer Concrete. Composite Structures, 57: 47-51.
Al-Salloum, Y. A., and Al-Musallam, T. H. 2007. Creep Effect on the Behaviour of
Concrete Beams Reinforced with GFRP Bars Subjected to Different Environments.
Construction and Building Materials, 21: 1510-1519.
Seica, M. V., and Packer, J. A. 2009. FRP Materials for the Rehabilitation of Tubular
Steel Structures for Underwater Applications. Composite Structures, 80: 440-450.
Caron, J.F., Julich, S., and Baverel, O. 2009. Selfstressed Bowstring Footbridge in
FRP. Composite Structures, 89(3): 489496.
Douthe, C., Baverel, O., and Caron, J.F. 2006. Form-Finding of a Grid Shell in
Composite Materials. Journal of the International Association for Shell and Spatial
Structures, 47(1): 53-62.
Douthe, C., Caron, J.F., and Baverel, O. 2010. Gridshell Structures in Glass Fibre
Reinforced Polymers. Construction and Building Materials, 24(9): 1580-1589.
Richard, F., and Perreux, D. 2001. The Safety-Factor Calibration of Laminates for
Long-Term Applications: Behaviour Model and Reliability Method. Composites
Science and Technology, 61: 2087-2094.
Choi, Y., and Yuan, R.L. 2003. Time-Dependent Deformation of Pultruded Fiber
Reinforced Polymer Composite Columns. Journal of Composites for Construction,
7(4): 356-362.
Bank, L.C., and Mosallam, A.S. 1992. Creep and Failure of a Full-Size FiberReinforced Plastic Pultruded Frame. Composites Engineering, 3(2): 213-227.
McClure, G., and Mohammadi, Y. 1995. Compression Creep of Pultruded E-GlassReinforced-Plastic Angles. Journal of materials in Civil engineering, 7(4): 269-276.
Scott, D.W., and Zureick, A.H. 1998. Compression Creep of a Pultruded EGlass/Vinylester Composite. Composites Science and Technology, 58: 1361-1369.
Shao, Y., and Shanmugam, J. 2006. Deflection Creep of Pultruded Composite Sheet
Piling. Journal of Composites for Construction, 8(5): 471-479.
Dillard, D.A., Straight, M.R., and Brinson, H.F. 1989. The Nonlinear Viscoelastic
Characterization of Graphite/Epoxy Composites. Composites Structures, 27(2): 116123.
Papanicolaou, G.C., Zaoutos, S.P., and Cardon, A.H. 1999. Prediction of Non Linear
Viscoelastic Response of Unidirectional Fiber Composites. Composites Science and
Technology, 59: 1311-1319.
Le Moal, P., and Perreux, D. 1994. Evaluation of Creep Compliances of Unidirectional
Fiber-Reinforced Composites. Composites Science and Technology, 51: 469-477.
Barboura, S., Caron, J.F., and Sab, K. 2010. Micromechanical Model of Viscoelastic
Behavior for Glass Fiber Reinforced Polymers. Third Euro-Mediterranean Symposium
on Advances in Geomaterials and Structures, 1: 239-244.
Guedes, R.M., Morais, J.J.L., Marques, A.T., and Cardon, A.H. 2000. Prediction of
Long-Term Behaviour of Composite Materials. Computers and Structures, 76: 183194.
297

298

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

CFRP CONFINEMENT OF NON-SEISMIC COLUMNS UNDER


EXPLOSIVE LOADING
A. Lloyd1, M. Saatcioglu2 and T.K. Tikka3
1

Graduate Student, Department of Civil Engineering, University of Ottawa


Professor, Department of Civil Engineering, University of Ottawa
3
Associate Professor, Department of Civil Engineering, Lakehead University
2

ABSTRACT
The possibility of a high explosive attack on structures has become a significant topic in
research and design within Canada. Explosion induced shock waves are very short duration,
high pressure loading that can cause severe damage to localized parts of the structure. It is
critical that reinforced concrete columns, as load bearing members, withstand blast loading
and retain some degree of post blast axial capacity. This paper explores the capacity of asbuilt reinforced concrete columns that were designed for non-seismic regions in Canada.
The use of carbon fibre reinforced polymer (CFRP) confinement as an external retrofit is
also explored with the goal of increasing strength and ductility of the columns.
As-built and retrofitted columns were tested under simulated blast loading using the
University of Ottawa shock tube. Using high pressure air, the shock tube creates a shock
wave that is representative of high explosive driven shock waves. Strain, displacement,
acceleration, pressure and high speed video were recorded during the shock tube testing.
The experimental data was used to develop a single degree of freedom numerical
integration model of the dynamic system. This numerical model is used to explain the
benefits of confinement retrofit of reinforced concrete columns with CFRP under blast
induced shock wave loading.

1. INTRODUCTION
Explosion induced shock waves are a very short duration, high pressure loading that can
cause severe damage to localized structural elements. As load bearing members, it is
critical that reinforced concrete columns withstand blast loading and retain sufficient postblast axial capacity. This paper explores the response of an as-built reinforced concrete
column and a column retrofitted with external carbon fibre reinforced polymer (CFRP)
confinement wraps subjected to shock tube induced shock wave loading. The columns were
tested at the University of Ottawa using a shock tube to simulate explosive induced shock
wave loading. The benefits of using CFRP confinement to mitigate the effects of shock

299

wave loading are discussed in terms of experimental results and an analytical model used to
predict column response.

2. EXPERIMENTAL METHOD
The two columns, presented in this paper, were designed to represent main storey columns
designed for low-seismicity zones within Canada. The columns were constructed using 410M longitudinal reinforcing steel bars with a yield strength of 483 MPa. Closed hoops
were constructed out of 6.3 mm diameter steel wire with a yield strength of 580 MPa and
spaced at a distance of 75 mm along the entire length of the column. The column crosssection measured 150 mm by 150 mm square with clear cover to the outside of longitudinal
steel bars of 10 mm. The columns had a total length of 2438 mm with a clear span between
supports of 1981 mm. Concrete strength was measured at 69 MPa at the time of testing.
One of the columns was retrofitted with a single layer of confinement wraps of carbon fibre
reinforced polymers which extended the entire length of the column terminating at the
supports. The FRP material was Tyfo SCH-41S-1 508 mm wide CFRP sheets [1] with 100
mm overlap on the column side. Prior to application of the sheets, the column corners were
rounded to obtain a 25 mm radius and any surface defects were filled with epoxy putty. The
CFRP had a specified composite laminate ultimate tensile strength in the primary fibre
direction of 876 MPa, a nominal thickness of 1 mm, and a specified ultimate tensile strain
of 1.2 %.

2.1 Shock Tube


The shock tube used to apply shock wave loading can simulate a wide range of explosions
with varying pressures and impulses [2]. This shock tube, shown in Fig. 1(a) and (b),
consists of a driver section, a double diaphragm spool section and an expansion section
terminating in a rigid test frame with a 2032 mm by 2032 mm square opening. The shock
tube has been shown to produce shock waves that are predictable and repeatable, allowing
for accurate tests of companion sets of specimen under specific impulse and pressure
combinations. Shock waves at the rigid test frame have been found to be planar in pressure
and arrival time [2].

a) Shock Tube Driver Section


b) Shock Tube End Frame
Fig. 1. University of Ottawa Shock Tube Testing Facility

300

2.2 Test Configuration


A composite steel beam and steel membrane system was designed to collect the pressure
generated by the shock wave and act as the lateral load transfer mechanism by applying a
series of point loads along the length of the column. The composite steel beam and steel
membrane system was fabricated from a series of 76.2 mm by 76.2 mm by 4.8 mm thick
hollow square structural steel transfer beams which were attached to a 0.71 mm thick steel
sheet. These transfer beams were placed perpendicular to the column height and were 2083
mm long. The steel sheet covered the entire area of the end frame. Neither the sheet nor the
steel beams were attached to the end frame during testing and were therefore free to move
and distribute the load entirely to the column being tested.
The columns were clamped to the shock tube test frame at both ends with stiffened steel
plates welded to HSS beams. These supports were intended to represent fully fixed end
conditions. To test the rigidity of the supports, a 152 by 152 by 6.3 mm thick hollow
structural steel member was clamped at the bottom support and pulled as a cantilever beam
at a point 2100 mm above the support. Both tip displacement and total applied force were
measured and used to find the support rotational spring constant, which was 903 kN.m/rad.

2.3 Instrumentation
The columns were instrumented with electrical resistance strain gauges on the tension and
compression reinforcing steel at locations of high strain. The change in axial load during
response was monitored with an electrical resistance-based load cell placed beneath the
column. Three linear variable displacement transducers (LVDT) were used to measure
lateral and axial displacements of the column. One LVDT measured in-plane flexural
displacement at the column mid-span and two LVDTs measured the axial displacement at
the column ends. Three piezoelectric shear accelerometers were used to measure
acceleration at the same locations and directions as the LVDTs.
Beyond the instrumentation listed above, several additional pieces of equipment were
required for dynamic testing. Dynamic piezoelectric pressure sensors were mounted within
the shock tube to measure the reflected shock wave. A high-speed digital oscilloscope was
used as the primary data acquisition controller and was configured to record at 100,000
samples per second. The data acquisition system was triggered and began recording data
when the shock front passed over one of the pressure sensors. The data acquisition system
recorded for a total duration of 3 seconds. In addition to the specimen instrumentation, a
high speed camera was used to monitor the tests. This camera, set to record 500 frames per
second, was triggered by the data acquisition system such that video was synchronized with
the recorded data-time history.

2.4 Test Regimen


Each column was loaded with three shock waves. The magnitude of the pressure and the
impulse of these shock waves were selected to activate a different level of response in the
as-built column. The first test was intended to cause the column to remain fully elastic,
while the second test was intended to cause the column to reach its elastic limit and the
301

third test was intended to cause significant damage to the column. The same three shock
waves were used to load the retrofitted column to facilitate direct comparison of the
experimental results.

3. EXPERIMENTAL RESULTS
3.1 As-Built Column
The first test on the as-built column did not cause any signs of damage. The second test
slightly exceeded this range resulting in small tension cracks and residual displacement of
13.1 mm. The third shock wave had a 73.0 kPa reflected pressure, a 839.5 kPa.ms reflected
impulse over the positive phase, and a 23.4 ms positive phase duration. This shock wave
resulted in a maximum displacement of 135.3 mm at mid-height which occurred 51.7 ms
after the start of the shock wave loading. Residual displacement was 136.5 mm after the
third test for a cumulative residual of 149.6 mm. The reflected pressure and mid-height
displacement-time histories are shown in Fig. 2(a) and the damaged column is shown in
Fig. 3(b).

a) As-Built Column
b) External CFRP Column
Fig. 2. Pressure and Displacement-Time History for Columns

a) Undamaged
b) Damaged
c) Undamaged
d) Damaged
As-Built
As-Built
Retrofitted
Retrofitted
Fig. 3. Damaged and Undamaged As-Built and CFRP Retrofitted Columns

302

The specimen was considerably damages after test 3. Plastic hinges formed at mid-height
and at the support regions, as shown in Fig. 3(b). The mid-height cover concrete on the
compression side spalled off. Rebar buckling was observed between the hoops at midheight. Several large cracks were formed on the tension face at the location of the supports
and at mid-height. Some of the tension concrete at mid-height outside of the hoops fell off
where cracks intercepted one another. Tension cracks were distributed over approximately
half the length of the column and were most significant at mid-height.

3.2 Column with External CFRP Retrofit


The first and second tests did not produce any visible signs of damage, but there was 10 mm
residual displacement indicating that the elastic limit of the column had been slightly
exceeded. The third shock wave had 71.0 kPa reflected pressure, a 652.4 kPa.ms reflected
impulse over the positive phase, and a 21.1 ms positive phase duration. This shock wave
resulted in a maximum displacement of 85.5 mm at mid-height that occurred 21.1 ms after
the start of the shock wave loading. Residual displacement was 38.7 mm after test 3. The
reflected pressure and mid-height displacement-time histories are shown in Fig. 2(b).
After the third test, there was minimal visible damage to the specimen other than the
noticeable residual displacement. The column was subjected to a fourth shock wave test,
unfortunately a problem with triggering of the data acquisition system caused a loss of all
data. Figure 3(d) shows the damaged column after the fourth test. Plastic hinges formed at
mid-height and at the support regions. The compression concrete at mid height did not
appear have any signs of crushing. One large crack opened at mid-height and at the
supports in the plastic hinge regions. The tension reinforcing steel ruptured at the locations
of plastic hinges at the supports and the mid-span. From other similar tests, the pressure and
impulse combination for this test can be estimated as 83.1 kPa and 836.1 kPa-ms,
respectively.

3.3 Comparison of As-Built and Retrofitted Column


Figure 4 compares the maximum and residual displacements recorded for the as-built and

retrofitted columns for similar pressure-impulse combinations. This figure shows that the
use of CFRP confinement wraps as a blast-retrofit technique can significantly improve the
performance of as-built columns. For levels of loading that do not cause significant damage
to the as-built column, there is no noticeable difference in the column responses. The
benefit of the CFRP confinement retrofit seems to be in its ability to prevent or delay
concrete crushing, providing higher resistance and ductility capacity at high displacements.

303

Fig. 4. Maximum and Residual Displacements for As-Built and Retrofitted Columns

4. ANALYSIS
4.1 Resistance Curve Development
The resistance-displacement relationship of a column under distributed load is found by
integrating the curvature of the section along the member length. This method involves first
determining the moment-curvature of the column section through strain compatibility.
Once the moment-curvature relationship is known, the displacement and moment resistance
is calculated at various critical levels including cracking, steel yield, and concrete crushing.
In the resistance-displacement calculations, the plastic hinge length of the column is taken
as equal to the section height, in this case 150 mm. The total applied force and displacement
at mid-span of the column form the resistance curve used in SDOF analysis.

4.1.1 Material Models


The reinforcing steel model used in the analysis was obtained through coupon tests which
provided yield, strain hardening and ultimate stresses and strains. The Saatcioglu and
Razvi3 confined normal strength concrete model was used for the core concrete of the asbuilt column. Unconfined concrete was modelled using Hognestads parabola4. The
Saatcioglu and Razvi confinement model was used for the cover and core concrete of the
column retrofitted with CFRP. For the core concrete, the confinement stress of the steel
hoops was added to the confinement stress of the CFRP wrap. Dynamic increase factors
(DIF) for material strength of 1.17 for reinforcing steel yield stress and 1.19 for concrete
strength were included in the material models5.

4.1.2 Moment Curvature


Figure 5(a) and (b) show the moment curvature relationship for the as-built and retrofitted

columns. Various moment-curvature curves have been developed for each section under
different axial loads. The variation of axial load with displacement is discussed later in this
section. From these figures, it is apparent that the maximum moments of the retrofitted
column are higher than the as-built column under all axial load levels. Furthermore, the
curvature capacity of the retrofitted column is consistently greater than the as-built column
indicating higher ductility capacity.

304

4.1.3 Variation in Axial Load


During experimental testing, a hydraulic jack was used to apply an axial compressive force
to each end of the column. As the column deflected horizontally, the projected vertical
length decreased. This change in length caused an equivalent reduction in axial load as the
horizontal displacement increased and the end of the column moved away from the fixed
vertical support (hydraulic jack). A composite resistance curve accounting for the change in
axial load with displacement was developed for this study. The magnitude of the axial load
at any given displacement was determined as a function of the initial strain due to the initial
axial load and the vertical displacement at the top of the column at a specified horizontal
displacement. In Fig. 5(c) and (d), the composite curve accounting for varying axial load is
plotted along with the resistance-displacement curves for constant axial loads of varying
magnitude. In general, a decrease in stiffness and resistance occurs with a drop in axial load
as the strength capacity of the section decreases. Although it is not shown in this paper, the
vertical displacements at the ends of the column were measured during the experimental
tests and the relationship between horizontal displacement and the shortening of the
projected length of the column was verified.

a) Moment Curvature As-Built

b) Moment Curvature Retrofitted

c) Resistance-Displacement As-Built
d) Resistance-Displacement Retrofitted
Fig. 5. Section and Member Properties

4.2 SDOF Method


A common method of member analysis under shock wave loading is the equivalent SDOF
approach. In this method, member properties (e.g. mass and stiffness) and the external
loading are converted into equivalent point loads acting on a single degree of freedom
oscillator by using load and mass transformation factors6. The forcing function was the
experimentally recorded pressure-time history multiplied by the area of the load transfer
305

device. The total mass of the SDOF system was taken as the mass of the load transfer
device (209 kg) plus the mass of the column (107 kg). The solution to the equation of
motion was obtained by numerical integration using average acceleration method. This
method is recommended by US Department of Defence Structures to Resist the Effects of
Accidental Explosions6 as it is unconditionally stable and converges upon the exact solution
with a small enough time step.

4.3 SDOF Results


Figure 6(a) and (b) shows the SDOF displacement-time histories for the as-built and CFRP
retrofitted columns after test 2 and test 3. From these plots it is evident that the SDOF
method predicts the response of the columns up to maximum displacement with a
reasonable degree of accuracy. Both the model and the experimental results show the
benefit of the CFRP retrofit in terms of limiting maximum displacement for high pressure
and impulse combinations (Fig. 6(b)).

(a) As Built Moderate Pressure


(b) CFRP Retrofit Moderate Pressure
Fig. 6. Single-Degree-of-Freedom and Experimental Displacement-Time Histories

5. CONCLUSIONS
The experimental results and analytical model of two reinforced concrete columns
subjected to shock tube induced shock wave loading, have been discussed in this paper.
These tests were undertaken to explore the benefits of using externally bonded CFRP wraps
as a retrofit to mitigate the effects of blast loading on columns by providing increased
confinement. The following conclusions were drawn from this research:

The use of confinement CFRP sheets limits damage and displacement of columns
subjected to shock wave loading.
Consideration of the variation of axial load is crucial in the development of accurate
resistance-displacement curves for analysis.
The complex flexure response of reinforced concrete columns seeing significant
damage can be accurately modeled by using simple SDOF models.

306

6. REFERENCES
1.
2.
3.
4.
5.
6.

Fyfe, CO. LLC. 2009. Tyfo SCH-41S-1 Composite Product Data. Manufacturer
material data sheet.
Lloyd, A. 2010. Performance of reinforced concrete columns under shock tube induced
shock wave loading. Masters thesis, University of Ottawa, Ottawa, Ontario, Canada.
Saatciogul, M. and Razvi, S.R. 1992. Strength and Ductility of Confined Concrete.
Journal of Structural Engineering, 118(6): 1590-1607.
Hognestad, E. 1951. A study of combined bending and axial load in reinforced
concrete members. Bulletin Series No. 399, Univ. Of Illinois Engrg. Experimental
Station, Urbana, Ill.
Unified Facilities Code. 2008. Structures to Resist the Effects of Accidental
Explosions. (UFC) 03-340-02, USA DoD, Washington, D.C.
Biggs, J.M. 1964. Introduction to Structural Dynamics. McGraw-Hill, New York, NY,
341 p.

307

308

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

PERFORMANCE OF CONCRETE BRIDGE BARRIERS


REINFORCED WITH CONDITIONED GFRP BARS IN ALKALINE
SOLUTION
E.A. Ahmed 1 and B. Benmokrane 2
1

Post-Doctoral Fellow, Dept. of Civil Engrg., University of Sherbrooke, Quebec Canada/ Lecturer,
Minoufiya University, Shebin El-Kom, Egypt.
2
Canada & NSERC Research Chairs Professor, Dept. of Civil Engrg., University of Sherbrooke,
Quebec, Canada

ABSTRACT:
The use of glass fibre-reinforced polymer (GFRP) bars has become an effective solution to
substitute the conventional steel bars and prevent the corrosion and related deterioration
problems. Many studies were conducted to evaluate the durability of the GFRP bare bars in
severe exposure conditions such as alkaline solutions by comparing the properties of the
bars before and after exposure to evaluate the residual properties and degradation. This
paper presents the residual capacity of concrete bridge barrier Type MTQ 210 reinforced
with GFRP bars that were subjected to alkaline solution at high temperature. The GFRP
stirrups were conditioned in alkaline solution at 60oC for 3 months in accordance with the
CSA-S807-10. Thereafter the stirrups were employed in fabricating a barrier prototype
Type MTQ 210 and tested using monotonic loading until failure. The behaviour of the
bridge barrier fabricated using the conditioned GFRP bars was compared with that of the
control prototype which was reinforced with unconditioned GFRP stirrups. The comparison
revealed that the behaviour of the barrier reinforced with conditioned GFRP stirrups was
similar to that of the control prototype (unconditioned).

1.

INTRODUCTION

The anticipated service life of many steel-reinforced concrete structures is shortened due to
corrosion of steel reinforcement and related deteriorations problems. Reinforced concrete
bridges are among those structures that suffer from corrosion and related deteriorations.
This problem is exacerbated when large amounts of de-icing salts are used on bridges
during the winter season. Whereas there is a multitude of methods to protect the steel
reinforcement from corrosion, such as epoxy coating or galvanizing, the use of noncorrodible fibre-reinforced polymer (FRP) materials can eliminate it.

309

Many investigations were conducted to evaluate the structure performance of FRPreinforced concrete (RC) members. The long-term durability properties were also addressed
through the accelerated aging exposure of GFRP bare bars and evaluating the residual
properties and the degradation. However, the structure behaviour of concrete members
reinforced with conditioned FRP bars may provide a realistic evaluation procedure to
evaluate the residual capacity of the RC members. Besides, it may be used to quantify the
effects of the changes of the mechanical properties due to the conditioning of the FRP
materials on the structural behaviour of the RC member.
The strength of the FRP stirrups is usually conducted according to the ACI 440.1R (2006)
(B.5 test method) or the CAN/CSA S806 (2002) (Annex E) as shown in Figure 1. In this
test, the bend strength of the FRP stirrup is determined through embedding the FRP stirrup
in concrete block with one side debonded to force the failure to occur in this side. The two
concrete blocks were pushed apart using a hydraulic jack till the failure of the stirrup which
normally happened at the bent (Figure 1). However, due to the geometry and smaller
dimensions of the GFRP stirrups that are being using in the curbs of the MTQ 210 barriers,
representative specimens with larger dimensions and different geometry to fit with the B.5
test method requirements were needed. However, more realistic tests were to test a concrete
barrier type MTQ 210 with unconditioned GFRP stirrup and another one with conditioned
stirrups. Thereafter, the behaviour of both prototypes can be compared and the effect of the
conditioning of GFRP stirrup on the behaviour of the barrier could be evaluated.
Thus, in this paper the main objective was to investigate the structural behaviour of the
concrete bridge barrier reinforced with pre-conditioned GFRP stirrups and compare its
performance with that reinforced with un-conditioned GFRP stirrups. This study included
construction and testing of two concrete barriers type MTQ 210. One of them was
reinforced with GFRP stirrups No. 15 which were conditioned in alkaline solution at 60oC
according to the CAN/CSA S807-10 (2010). However, the second one was reinforced with
un-conditioned GFRP stirrups No. 15 which was considered as reference specimen. Those
specimens were a part of an extended experimental program for investigating the behaviour
of GFRP-RC barriers under static and impact loading (Ahmed et al. 2009). This program
was conducted at the University of Sherbrooke in collaboration with the Ministry of
Transportation of Quebec (MTQ).

300 mm

400 mm

Steel stirrups 10 @ 75 mm
to prevent splitting
Debonding tube

300 mm
l d = rb + d b

db

P
FRP stirrups
Concrete block

rb = bend radius
d b = bar diameter
ld = embedment length

Fig. 1. Testing of FRP stirrups to determine the bend strength


310

2. TEST PROTOTYPES
The experimental program included two full-scale concrete bridge barriers type MTQ 210
reinforced with GFRP bars. One of them was reinforced with GFRP bars without any
conditioning (210-GFRP-1) whereas the curb of the second one was reinforced with preconditioned GFRP stirrups (210-GFRP-2). The slabs of the barrier barriers were reinforced
with GFRP bars No. 20 spaced at 75 placed transversally in the top layer while the
transverse bottom reinforcement was GFRP bars No. 20 spaced at 125 mm. The top and
bottom longitudinal reinforcement of the slab was GFRP No. 20 spaced at 185 mm. On the
other hand, the curb of the barriers was reinforced with GFRP stirrups No. 15 spaced at 300
mm and 5 GFRP bars No. 15 placed longitudinally. Figure 2 shows the reinforcement
configuration of the MTQ 210 barriers and the geometry of the GFRP stirrups that were
employed in this investigation.
Var.

275
450
35

395

415

8
1

275
GFRP No. 15 (5)

CHAMFER
15 x 15 (TYPE)
75 Clear cover

275

GFRP Stirrups

GFRP No. 15 @ 220

GFRP No. 20 @ 125


1000

1000

225

GFRP No. 20 @ 185


38 149 38

114 225 15

280

GFRP No. 20 @ 75 mm

500

Fig. 2. GFRP reinforcement of the MTQ 210 barriers and stirrups configuration
The conditioning was conducted in accordance with the Specification for Fibre-Reinforced
Polymers CAN/CSA S807-10 (2010). The GFRP stirrups were kept submerged in alkaline
solution at 60oC for 3 months (14 weeks exactly). The specimens were placed in the
alkaline solution at the 60oC on May 15, 2009 and were removed from the alkaline solution
on September 1, 2009. Thereafter, they were instrumented and utilized in fabricating the
MTQ 210 barrier prototype (210-GFRP-2). The reference barrier prototype was cast on
April 2009 while the conditioned prototype was cast on September 2009. Figure 3 shows
the conditioning of the GFRP stirrups in the environmental chamber at the University of
Sherbrooke. Figure 4 shows the GFRP stirrups after being conditioned and instrumented
them with electrical resistance strain gauges.

311

The test prototypes were constructed using ready-mixed normal weight concrete type
MTQ-V with a target compressive strength of 35 MPa after 28 days. The concrete strength
was determined at the day of the test by testing of 150300 mm cylinders. The concrete
strength of 2010-GFRP-1 and 2010-GFRP-2 was 37.8 and 50.0 MPa, respectively. Figure 5
shows the fabrication of the 210-GFRP-2 barrier which was reinforced with conditioned
GFRP stirrups.

Fig. 3. Conditioning of GFRP stirrups in alkaline solution at 60oC

Fig. 4. GFRP stirrups after being conditioned and instrumented

312

Fig. 5. Fabricating the 210-GFRP-2 using conditioned GFRP stirrups

2.1 Instrumentation
The barrier prototypes were instrumented with electrical resistance strain gauges to
measure strains in reinforcing bars and in concrete at critical locations as shown in Figure
6. Each MTQ Type 210 barrier prototypes were instrumented with 25 strain gauges
attached to the reinforcing bars and 5 gauges attached to the concrete surface. The strain
gauges attached to the reinforcing bars were arranged at three different transversal sections
and the strain gauges in one of those sections are shown in Figure 3. However, the concrete
strain gauges were attached in the critical locations to capture the compressive strains of the
concrete during the test as shown in Figure 6. The induced displacement and deflection in
the tested barrier prototypes were also measured using many Linear Variable Differential
Transformers (LVDTs) as shown in Figure 6.
Reinforcement strain gauge
Concrete strain gauge
LVDT

Fig. 6. Instrumentation of the MTQ 210 barriers

313

2.2 Test Setup and Procedure


The barrier prototypes were supported on two parallel supports spaced at 1.0 m with an
overhang of 1.0 m to simulate the actual barriers in real bridge decks. The slab of the
barrier prototype was tightened to the laboratory rigid floor by six 38-mm diameter
Dywidag bars (anchors) and nuts. A tensile force of 100 kN was applied on each anchor bar
to assure that no rigid body displacement occurs during testing. 45-mm thick square steel
plates (200200 mm) were used as bearing plates between the nuts and the slab. A clear
space of 114 mm was used between the bottom surface of the barrier and the laboratory
strong floor to allow for deflection and rotation of the barrier wall and slab during testing.
The load was applied horizontally to the center of the steel railing at a distance of 775 mm
from the slab surface. The load was transmitted from the hydraulic jack to the barrier
prototypes through a 2 m-long steel I-beam. Besides, the steel I-beam evenly distributes the
applied load over the length of the bridge railing to prevent any localized premature failure.
The load is applied using a 1000 kN-capacity hydraulic actuator connected to a manual
pump. Figure 7 shows the details of the test setup.

114

225

776

Load

1000

1000

500

Fig. 7. Test setup of the 210 barrier prototypes

3. TEST RESULTS AND COMPARISONS


3.1 Ultimate Capacity and Mode of Failure
Table 1 presents the ultimate capacities and modes of failure of the unconditioned (210GFRP- 1) and conditioned (210-GFRP- 2) barriers tested herein. It can be noticed that the
two prototypes had close ultimate capacities. The ultimate capacity of the conditioned
prototype (210-GFRP-2) was 87% of that of the unconditioned one (210-GFRP-1).
Table 1. Ultimate capacities and modes of failure of the MTQ 210 barrier prototypes
Ultimate
Barrier prototype
Mode of failure
capacity (kN)
Diagonal tension failure in the curb at
210-GFRP-1 (Unconditioned)
265.8
the connection between the bridge
210-GFRP-2 (Conditioned)
231.0
railing and the curb.

314

Figure 8 shows the failure of the 210-GFRP-1 and 210-GFRP-2 barriers. The 210-GFRP-1
prototype failed by diagonal tension failure in the curb at the connection between the bridge
railing and the curb. Similar mode of failure was observed for the conditioned prototype
(210-GFRP-2) as shown in Figure 8. From this figure, it can be noticed that there was no
difference between the two prototypes regarding the mode of failure.

(a) Unconditioned (210-GFRP-1)

(b) Conditioned (210-GFRP-2)

Fig. 8. Failure mode MTQ 210 barriers: (a) 210-GFRP-1 (Unconditioned); (b) 210-GFRP-2
(Conditioned)

3.2 Deflections
Figure 9 shows a comparison between the load-deflection relationships for the 210-GFRP-1
(unconditioned) and 210-GFRP-2 (conditioned) barrier prototypes. From the comparison, it
can be noticed that the deflection of the 210 conditioned barrier prototype (210-GFRP-2) is
in good agreement with that of the unconditioned one (210-GFRP-1). However, the
deflection of the conditioned prototype (210-GFRP-2) was slightly higher than that of the
unconditioned one (210-GFRP-1). Moreover, the maximum measured deflection, at the
maximum applied loads, was almost similar and ranged between 75 and 85 mm.
No. 1 (L)

No. 1 (R)

No. 2 (L)

No. 2 (R)

No. 3 (L)

No. 3 (R)

200

No. 4 (L)

No. 4 (R)

150

No. 5 (L)

No. 5 (R)

Load (kN)

250

100

300

No. 1 (R)

No. 2 (L)

No. 2 (R)

No. 3 (L)

No. 3 (R)

200

No. 4 (L)

No. 4 (R)

150

No. 5 (L)

No. 5 (R)

100

No. 1 (L)-Unconditioned
No. 1 (R)-Unconditioned
No. 1 (L)-Conditioned
No. 1 (R)-Conditioned

50

No. 1 (L)

250

Load (kN)

300

No. 3 (L)-Unconditioned
No. 3 (R)-Unconditioned
No. 3 (L)-Conditioned
No. 3 (R)-Conditioned

50
0

0
0

20

40

60

80

100

120

20

40

60

80

100

120

Deflection (mm)

Deflection (mm)

Fig. 9. Comparisons of load-deflection relationships for the 210-GFRP-1 and 210-GFRP-2

315

No. 1 (L)

No. 1 (R)

No. 2 (L)

No. 2 (R)

No. 3 (L)

No. 3 (R)

200

No. 4 (L)

No. 4 (R)

150

No. 5 (L)

No. 5 (R)

Load (kN)

250

100

300

No. 1 (L)

No. 1 (R)

No. 2 (L)

No. 2 (R)

No. 3 (L)

No. 3 (R)

200

No. 4 (L)

No. 4 (R)

150

No. 5 (L)

No. 5 (R)

250

Load (kN)

300

100

No. 4 (L)-Unconditioned
No. 4 (L)-Conditioned

50

No. 5 (L)-Unconditioned
No. 5 (R)-Unconditioned
No. 5 (L)-Conditioned
No. 5 (R)-Conditioned

50

No. 4 (R)-Conditioned

0
0

20

40

60

80

100

120

20

40

Deflection (mm)

60

80

100

120

Deflection (mm)

Fig. 9 (Contd). Comparisons of load-deflection relationships for the 210-GFRP-1 and 210GFRP-2

3.3 Strains in the Slab Reinforcement


Figure 10 shows a comparison between the applied loads-measured slabs strains
relationships of the unconditioned (210-GFRP-1) and conditioned (210-GFRP-2) barrier
prototypes. It can be noticed from Figure 8 that the maximum measured strains in the top
transverse reinforcement at failure was about 3600 microstrain for the 210-GFRP-1 barrier
prototype and about 3680 microstrain for the 210-GFRP-2 barrier prototype. From the
comparison, it can be noticed that there was a good agreement between the measured strain
in both prototypes (210-GFRP-1 and 210-GFRP-2).
300

300
No. 2

No. 3

No. 1

No. 4

250

250

200

200
Load (kN)

Load (kN)

No. 1

150
100

No. 1-Unconditioned
No. 2-Unconditioned
No. 1-Conditioned
No. 2-Conditioned

50
0
0

500

No. 2

No. 3

No. 4

150
100

No. 1 (R)-Unconditioned
No. 2 (R)-Unconditioned
No. 1 (R)-Conditioned
No. 2 (R)-Conditioned

50
0

1000 1500 2000 2500 3000 3500 4000


Strain (microstrain)

500

1000 1500 2000 2500 3000 3500 4000


Strain (microstrain)

Fig. 10. Comparison of reinforcement strains in the slabs (210-GFRP-1 and 210-GFRP-2)

3.4 Strains in the Curb Reinforcement


Figure 11 shows a comparison between the maximum measured strains in the curbs of the
GFRP-210-1 and GFRP-210-2 barrier prototypes. The maximum strain in the curb of the
unconditioned and conditioned prototypes (210-GFRP-1 and 210-GFRP-2) was 7900
microstrain at the failure load (265.8 kN) for the 210-GFRP-1 and was 9640 microstrain at
the failure load (231 kN) for the 210-GFRP-2. From the comparison in Figure 11 it can be

316

noticed that both conditioned and unconditioned barriers showed very close load-strain
relationships.
300
No. B2 (R)-Unconditioned

250

No. B2 (R)-Conditioned

Load (kN)

200
150

No. B1
No. B2

No. B1
No. B2

100
50
0
0

2000

4000

6000

8000

10000

12000

Strain (microstrain)

Fig. 11. Comparison of the maximum strains in the curbs (210-GFRP-1 and 210-GFRP-2)

4. SUMMARY AND CONCLUDING REMARKS


The main objective of this investigation was to evaluate the long-term performance of the
GFRP stirrups employed in MTQ 210 barrier prototypes. The curb of the first one (210GFRP-1) was reinforced with closed GFRP No. 15 stirrups. The curb of the second
prototype (210-GFRP-2) was reinforced with closed GFRP stirrups No. 15 that have been
subjected to alkaline solution at 60oC for 3 months (exactly 14 weeks) in accordance with
the CAN/CSA S807-10 (2010). Based on the test results and the comparisons between the
two prototypes the following concluding remarks can be drawn:

Both of the conditioned (210-GFRP-2) and the unconditioned (210-GFRP-1)


prototypes in showed the same mode of failure due to diagonal tension failure in the
curb at the connection between the bridge railing and the curb.
There was no significant difference between the two prototypes (210-GFRP-1 and
210-GFRP-2) in term of load-carrying capacity, slab reinforcement strains, stirrup
strains and the deflection at different locations.
Based on the comparative study of the two prototypes, it can be concluded that the
conditioning of the GFRP stirrups in alkaline solution at 60oC for three months did
not affect significantly the performance of the GFRP-RC barriers reinforced with
those conditioned stirrups.

5. ACKNOWLEDGEMENTS
The authors acknowledge the financial support received from the MTQ. The authors also
acknowledge Pultrall Inc. (Thetford Mines, Qubec), GFRP bar supplier; Beton Demix, the
concrete provider; and Les Coffrages Carmel Inc., fabricator of formworks, concrete
casting and curing. The authors would like to thank the technical staff in the structural
laboratory at the University of Sherbrooke.

317

6. REFERENCES
1.
2.
3.
4.

ACI Committee 440. 2004. Guide Test Methods for Fibre-Reinforced Polymers (FRPs)
for Reinforcing or Strengthening Concrete Structures. ACI 440.3R-04, American
Concrete Institute, Farmington Hills, Michigan, 40 p.
Ahmed, E., Dulude, C., and Benmokrane, B. 2009. Static and Dynamic Testing of 210
and 311 Barriers Reinforced with GFRP Bars. Technical Progress Report Submitted to
the Ministry of Transportation of Quebec (MTQ), September, 129 p.
Canadian Standards Association (CSA). 2002. Design and Construction of Building
Components with Fibre Reinforced Polymers. CAN/CSA S80602, Mississauga, ON,
Canada, 177 p.
Canadian Standards Association CSA. 2010. Specification for Fibre-Reinforced
Polymers. CAN/CSA S80710, Mississauga, Ontario, Canada, 27 p.

318

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

THERMAL ENDURANCE AND FATIGUE PERFORMANCE OF


PRESTRESSED CONCRETE PRISM REINFORCEMENT IN CIVIL
INFRASTRUCTURE
H.M. Vogel1 and D. Svecova2
1
2

Ph.D. Candidate, Dept. of Civil Eng., Univ. of Manitoba, Winnipeg, Canada


Associate Professor, Dept. of Civil Eng., Univ. of Manitoba, Winnipeg, Canada

ABSTRACT
This paper investigates the use of Prestressed high strength Concrete Prisms (PCPs) as a
reinforcing technique to take advantage of the high strength of fibre reinforced polymers
(FRP) in civil infrastructure. PCPs are small sections in which concrete is cast around a
concentrically prestressed FRP bar. The fatigue performance and thermal durability of
PCPs containing Carbon FRP (CFRP) were investigated and compared to that of beams
prestressed to the same level with CFRP reinforcement. Samples were designed in
accordance with guidelines for minimum concrete cover requirements to account for
differential swelling of FRP reinforcement in concrete subjected to thermal gradients. The
PCP reinforced beams were found to show more deterioration during the course of fatigue
testing. After the completion of fatigue testing, final static tests performed to ultimate did
not reveal any significant influence from repeated loading on the post-cracking behaviour.
Results also suggest a requirement for more research in the area of thermal weathering,
which was shown to have slightly beneficial influence on flexural performance.

1. INTRODUCTION
Studies have shown that the distinct mechanical and bond properties of FRP reinforcement
can significantly influence the analysis and design of reinforced concrete structures (ACI
440.1R 2001, Yost et al. 2003). Although higher tensile strengths allow the use of smaller
reinforcing areas and lower costs in civil infrastructure, the flexural stiffness of concrete
beams reinforced with FRP becomes a concern. The project presented in this paper uses
high strength concrete in Prestressed Concrete Prisms (PCPs) as an alternate reinforcing
technique to improve the flexural stiffness beyond cracking of the beam and reduce
deflections as well as crack widths under service loads. Although extensive research has
been performed on the behaviour of PCP reinforced concrete (Mirza et al. 1971, Zia et al.
1976, Chen and Nawy 1994), the durability of PCPs containing FRP reinforcement still
needs to be investigated. Consequently, the endurance of this reinforcing technique was

319

examined by studying the behaviour of PCP reinforced beams subjected to a combination


of repeated loading and thermal gradients expected in the Canadian climate.

2. EXPERIMENTAL PROGRAM
2.1 Section Geometry and Test Procedure
As illustrated in Fig. 1, all beams in this paper were designed to a width of 125mm and a
depth of 250mm with the reinforcement centred at 50mm from the bottom. Two 6mm
diameter steel reinforcing bars were provided at 30mm from the top of each beam to serve
as compression reinforcement. The beams were 2550mm in length. Four beams were pretensioned with a 9mm diameter CFRP tendon while four beams were reinforced with
50mm high by 50mm wide PCP reinforcement. The PCP reinforcement was also pretensioned with a 9mm diameter CFRP tendon. Steel stirrups were used to tie the
reinforcement in a cage as well as provide the required shear resistance during testing.
125m
30mm

125m
30mm

6mm
Steel

170mm

170mm
50 mm

6mm
Steel

9mm CFRP
[Aslan 200]

50mm
6mm Steel 9mm CFRP
Stirrup
[Aslan 200]
(b) Prestressed Beam

6mm Steel 50mm x 50mm


PCP
Stirrup
(a) PCP Reinforced Beam

Fig. 1. Beam Geometry


All beams in this paper were tested under three point loading with the load centered
between the supports. A 10mm high by 10mm deep notch was introduced over the width of
the beam at midspan to ensure the first crack initiates at the point of maximum moment.
Two 50mm long electrical resistance strain gauges (ESGs) were placed on the surface of
the beam prior to testing, at the reinforcing level and at 50mm on either side of the crack at
midspan. The strains were used as a measure of the reinforcements ability to transfer load
to the surrounding concrete.
2.2 Material Properties, Prestress Application and Thermal Weathering

The CFRP bars used in the research were manufactured by Hughes Brothers. The bars
revealed an ultimate tensile strength of 2,563MPa with an elastic modulus of 171,962MPa.
After 28days of moist curing, the high strength concrete used in the PCPs reached a
compressive strength of 131MPa with an elastic modulus of 45,150MPa and a splitting
tensile strength of 6.8MPa. For similar curing age and conditions, the normal strength

320

concrete used for the beams reached a compressive strength of 50.1MPa, an elastic
modulus of 28,949MPa and a splitting tensile strength of 4.2MPa.
All reinforcing bars were prestressed to 27% of their ultimate strength using a hydraulic
jack as well as the coupler mechanism shown in Fig. 2(a) to prevent damage of the
reinforcement incurred by the use of steel chucks. A load cell was inserted between the
jack and the bulkhead to determine the load applied to the tendons during prestress
application. The prestressing bed in Fig. 2(b) was built on the top flange of a W920x420
steel section with threaded rods for adjustable prestressing heights. One of the 6mm ESGs
from each sample was connected to a data acquisition (DAQ) system to monitor elongation
and verify the applied load during prestressing. Strain gauges revealed average total losses
of 12% for the PCPs and 17% for the prestressed beams at the time of test.
9mm

Steel Coupler

20 mm 13mm

50 mm

L51x51x6.4
50 mm

35mm FRP Anchor Steel Chuck

Threaded Overlaid
Rod
Plyform

Steel Spacer
95mm

W920x420

55mm
(a) Steel Coupler Mechanism
(b) Prestressing Bed Dimensions
Fig. 2. Prestressing Application Details
After prestress release, two PCP reinforced beams and two CFRP prestressed beams were
subjected to thermal weathering. Standardized methods for evaluating the performance of
reinforcement in a concrete environment subjected to temperatures in the Canadian climate
do not exist. The specimens were therefore subjected to 50 cycles ranging from 23oC and
+40oC, based on the requirements of ASTM E1512 (2001) for the durability of anchors
under freezing and thawing.

3. TEST RESULTS AND DISCUSSION


After thermal cycling, weathered beams were prepared for testing along with those kept at
ambient temperature. Table 7 provides a description for sample names with respect to the
weathering conditions, loading scheme and reinforcing scheme adopted.
Samples tested for fatigue were subjected to loading cycles ranging between 2 and 15kN.
The maximum load was chosen to lie between the cracking load of the beam (12kN) and
the cracking load of the PCPs (18kN) to preserve the flexural stiffness required to minimize
deflections at service conditions. The minimum load was chosen to reduce movement at the
bearing points. Loading was repeated for 2,000K cycles, with intermediate static tests
performed at 250K cycle intervals to evaluate changes in bond performance, crack width
321

and flexural stiffness. After completing the cycles, a final static test to ultimate was
performed to estimate the residual flexural performance of the samples.
Table 7. Sample Description
Beam Name

Weathering [oC]

Loading

Reinforcing

PCP-B2-1T
PCP-B2-2
PCP-B2-3TF
PCP-B2-4F
PC-B2-1T
PC-B2-2
PC-B2-3TF
PC-B2-4F

-25 to 40
20
-25 to 40
20
-25 to 40
20
-25 to 40
20

Static
Static
Fatigue
Fatigue
Static
Static
Fatigue
Fatigue

PCP
PCP
PCP
PCP
CFRP Prestressed
CFRP Prestressed
CFRP Prestressed
CFRP Prestressed

3.1 Bond Performance Trend under Fatigue


The bond performance was monitored during each fatigue test using ESGs instrumented at
50mm on each side of the crack at midspan. If the bond performance of the reinforcement
deteriorates, the transfer of forces will require longer anchorage and lower readings will be
recorded at the sensor location. Results at each of the intermediate static tests are plotted in
Fig. 3Fig. 3(a) for the PCP reinforced beams at an applied load of 15kN. Results from the
prestressed beams are shown in Fig. 3Fig. 3(b) for comparison. The level of deterioration
ranges from 17% (PCP-B2-3TF) to 38% (PCP-B2-4F) for the PCP reinforced beams and
from 14% (PC-B2-3TF) to 25% (PC-B2-4F) for the prestressed beams. These ranges
suggest higher uncertainty in the bond performance for PCPs. Although the result can be
associated with the difficulty of achieving a surface finish that is as durable for PCP
reinforcement as it is for commercially available reinforcement, further research on a larger
set of samples is required for confirmation. Results from Fig. 3(a) also suggest that the PCP
reinforced beam subjected to thermal weathering shows lower deterioration than that
obtained for the beam kept under ambient conditions. The same result was observed for the
prestressed beams, but to a lesser extent.

(a) PCP Reinforced Beams


(b) Prestressed Beams
Fig. 3. Concrete Strain Trend with Repeated Loading
322

3.2 Crack Width Trend under Fatigue


The progression in crack width during the course of fatigue testing is shown Fig. 4(a) for
the PCP reinforced beams at an applied load of 15kN. The readings were recorded at the
reinforcing level using a PI-gauge and are plotted with respect to the number of cycles at
the intermediate static tests. Results from the prestressed beams are shown in Fig. 4(b) for
comparison. The figures confirm the results obtained for bond performance. The reduction
observed in strains at the surface of the concrete corresponds to larger slip of the
reinforcement and an increase in crack width. Results continue to suggest larger
deteriorations in the parameter for PCP reinforcement. The extent of deterioration ranges
between 48% (PCP-B2-3TF) and 84% (PCP-B2-4F) for the PCP reinforced beams and up
to 33% for each of the prestressed beams. Results also suggest more variability in crack
width for the PCP reinforced beams and less deterioration with thermal weathering.

(a) PCP Reinforced Beams


(b) Prestressed Beams
Fig. 4. Crack Width Trend with Repeated Loading

3.3 Flexural Stiffness Trend under Fatigue


The bond performance and crack width considered in the previous sections are parameters
whose measurements were localized with respect to the midspan section. The flexural
stiffness of the member, on the other hand, is influenced by the contribution of properties
over the complete length of the member. It can therefore provide a global indication of the
flexural performance over the service life of the member.
The progression in flexural stiffness is shown in Fig. 5(a) for an applied load of 15kN at
each of the intermediate static tests for the PCP reinforced beams. The values in this figure
were obtained by fitting a linear model through the load-deflection data using a least
squares regression method at each of the intermediate static tests. When comparing results
to those obtained for the prestressed beams in Fig. 5(b), it is apparent that a gradual
reduction in flexural stiffness was observed with repeated loading. The deterioration in the
parameter appears to be more pronounced for the PCP reinforcement. The amount of
deterioration ranges from 29% (PCP-B2-3TF) to 32% (PCP-B2-4F) for the PCP reinforced
beams and up to 15% (PC-B2-4F) and 16% (PC-B2-3TF) for the prestressed beams.

323

3.4 Final Static Tests and Residual Performance


Load-deflection responses obtained from the final static tests performed after fatigue testing
are shown in Fig. 6(a) for all PCP reinforced beams and in Fig. 6(b) for all prestressed
beams. With the exception of beam PC-B2-1T, the figure suggests that deteriorations
measured during repeated loading were localized to the midspan section alone and that
thermal weathering did not appreciably affect the post-cracking behaviour of the beams.

(a) PCP Reinforced Beams


(b) Prestressed Beams
Fig. 5. Flexural Stiffness Trend with Repeated Loading

PCP-B2-4F
PCP-B2-1TS

PC-B2-1T
PC-B2-4F

PCP-B2-2
PCP-B2-3TF

PC-B2-2

PC-B2-3TF

(a) PCP Reinforced Beams


(b) Prestressed Beams
Fig. 6. Load-Deflection Response
The average bond performance for each type of reinforcement was also evaluated to
confirm the absence of deterioration observed in Fig. 6. The average bond strength was
established by considering force equilibrium of a bar segment between each set of cracks
formed during the final static test. The parameter depends on the force gradient between
cracks and the surface area of the reinforcement within the crack spacing. Results are listed
in Table 1(a) for the PCP reinforced beams and Table 1(b) for the prestressed beams. The
table includes information for the length of beam over which the cracks were developed.
Values for PCPs are comparable to that listed in the CSA A23.3 (2004) code for estimating
shear resisted by an interface separating concrete cast at different times. The value of
0.25MPa listed by the code is slightly higher than the average of 0.23MPa obtained during
testing but suggests that PCPs have adequate bond. Lower bond strength was observed for
beams subjected to repeated loading with increases for samples exposed to weathering.
324

Bond strength values for the prestressed beams were found to average 2.67MPa. Over the
range of loading and weathering conditions considered, the standard deviation suggests less
variability in bond strength for the PCPs. However, a larger set of samples needs to be
tested to clarify the result obtained for beam PC-B2-1T and confirm the post-cracking
behaviour observed in Fig. 6. The bond strength was found to decrease for beams subjected
to repeated loading and increase for those exposed to thermal weathering.
Table 8. Average Bond Strength
(a) PCP Reinforced Beams
Beam Label
Cracked Length [mm]
PCP-B2-1T
1225
PCP-B2-2
1275
PCP-B2-3TF
1525
PCP-B2-4F
1450
AverageSD
N/A
(b) Prestressed Beams
Beam Label
Cracked Length [mm]
PC-B2-1T
1275
PC-B2-2
1100
PC-B2-3TF
1100
PC-B2-4F
1175
AverageSD
N/A

Number of Cracks
10
11
14
12
N/A

Bond Strength [MPa]


0.23
0.24
0.22
0.22
0.230.01

Number of Cracks
7
6
6
7
N/A

Bond Strength [MPa]


2.98
2.58
2.60
2.51
2.670.21

4. CONCLUSIONS AND RECOMMENDATIONS


This paper presents results from tests performed on 4 PCP reinforced beams and 4
prestressed beams to investigate the influence of thermal weathering and repeated loading.
The following conclusions can be drawn from the results:
1.
2.
3.
4.
5.

6.

Deterioration in bond performance during fatigue ranged from 17 to 38% for the PCP
reinforced beams and 14 to 25% for the prestressed beams. The deterioration was
lower for beams subjected to thermal weathering.
The increase in crack width ranging from 48 to 84% for the PCP reinforced beams
compared to 33% for each of the prestressed beams subjected to repeated loading. Less
increase in crack width was observed for beams subject to thermal weathering.
A reduction in flexural stiffness ranging between 29 and 32% was observed for the
PCP reinforced beams subjected to repeated loading, compared to 15 and 16% for the
prestressed beams. Thermal weathering did not influence deterioration.
The deteriorations observed during fatigue testing did not appear to affect the postcracking behaviour during the final static tests.
When analyzing the stabilized crack pattern, bond strength values were found to
average 0.23MPa for the PCPs. The value is comparable to that given by the CSA
A23.3 (2004) design code for shear resisted by an interface separating concrete cast at
different times.
Bond strength values for the prestressed beams averaged 2.67MPa. The standard
deviation suggests less variability in the parameter for PCPs but more testing is
325

required to confirm this result and the durability improvement for beams subjected to
weathering.

5. ACKNOWLEDGMENTS
The authors would like to thank the Natural Sciences and Research Council of Canada
(NSERC) for their funding. The technical assistance of Mr.Chad Klowak and Mr.Grant
Whiteside is also greatly appreciated.

6. REFERENCES
9.
10.
11.
12.
13.
14.
15.
16.

ACI Committee 440. 2001. Guide for the design and construction of structural concrete
reinforced with FRP bars. ACI 440.1R-01 American Concrete Institute, Farmington
Hills, Michigan, USA, 41 pp
Yost, J.R., Gross, S.P. and Dinehart, D.W., 2003. Effective moment of inertia for
GFRP reinforced concrete beams. ACI Structural Journal, 100(6): 732-739.
Benmokrane, B., Chaallal, O. and Masmoudi, R. 1996. Flexural Response of Concrete
Beams Reinforced with FRP Reinforcing Bars, ACI Structural Journal, 93(1): 4655.
Mirza, J.F., Zia, P. and Bhargava, J.R. 1971. Static and Fatigue Strength of Beams
Containing Prestressed Concrete Tension Elements. Final Report NCSU Project FRD110-69-2, North Carolina State University, North Carolina, USA.
Zia, P., Mirza, S. and Rizkalla, S. 1976. Static and Fatigue Tests of Composite Tbeams Containing Prestressed Tension Elements. PCI Journal, 21(6): 76-93.
Chen, B. and Nawy, E.G. 1994. Structural Behavior Evaluation of High Strength
Concrete Beams Reinforced with Prestressed Prisms Using Fiber Optic Sensors. ACI
Structural Journal, 91(6): 708-718.
ASTM. 2001. Testing Bond Performance of Adhesive-Bonded Anchors. ASTM
E1512-01, Annual Book of ASTM Standards, v04.11, Easton, MD, USA, pp. 664-668.
Canadian Standard Association (CSA). 2004. Design of Concrete Structures. CSA
A23.3-04, Toronto, ON, Canada, 358 p.

326

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

FRP EXTERNALLY BONDED SYSTEMS FOR A SUSTAINABLE


STRENGTHENING OF MASONRY STRUCTURES
F. Ceroni1, B. Ferracuti2, M. Pecce3 and M. Savoia4
1

Assistant Professor, Engineering Dept., Univ. of Sannio, Benevento, Italy.


Ph. D., Dept. of Civil, Environmental and Materials Engineering, Univ. of Bologna, Italy.
3
Professor, Engineering Dept,. Univ. of Sannio, Benevento, Italy.
4
Professor, Dept. of Civil, Environmental and Materials Engineering, Univ. of Bologna, Italy.
2

ABSTRACT
According to recent design Codes and Recommendations, verification against debonding of
FRP-reinforcement is imperative in the design of strengthening interventions. In FRP bonded
masonry elements, end debonding must be avoided by checking the maximum force which
can be carried out by reinforcement. In this paper, design criteria against debonding in
masonry elements externally strengthened by FRP sheets are proposed. For different masonry
supports, the results of a statistical analysis of bond tests reported in the literature are
collected and used for calibrating prediction formulas. Starting from the experimental
database, formulas for fracture energy of interface bond law and bond strength are assessed.
In order to provide design formulas according to the Partial Safety Factor Method for
Ultimate Limit State verification, mean and characteristic values of the bond strength have
been determined by means of statistical methods.

1. INTRODUCTION
Strengthening techniques by means of externally bonded FRP systems are often adopted for
masonry structures, especially in historical buildings. Their reversibility is in fact considered
a key aspect in the framework of sustainable interventions.
In masonry elements externally bonded with Fiber Reinforced Polymers (FRPs), the bond
mechanism is a central topic for the effectiveness of the strengthening. Debonding is a very
brittle failure mode usually occurring in the masonry support few millimetres below the
adhesive layer, due to the low tensile strength of masonry; debonding does not allow to reach
the full strength of the reinforcement and, thus, should be avoided. The bond mechanism is
complex, depending on the mechanical properties of the masonry block, adhesive, and FRP
reinforcement. It is worth to notice that the term masonry refers to a wide variety of
materials, which can be grouped in two main categories: natural stones and artificial clay
bricks. The natural stones are comprehensive of several materials depending on the different
327

geological origin and composition in function of the quarry location and, thus, they can have
very dissimilar mechanical properties (compressive and tensile strength, Youngs modulus,
porosity, surface consistency); conversely, the clay bricks are artificially made and
characterized by less scattered mechanical properties. The large property variability as a
function of the support material can determine very different results in terms of bond
behaviour and bond strength.
In this paper, the results of several bond tests on different types of masonry blocks have been
examined in order to check the effectiveness and the sustainability of the strengthening
technique with FRP materials. In order to reduce the scatter of the experimental data, the
results of bond tests on natural stones and artificial clay bricks have been examined
separately. Design rules for bond strength have been then calibrated by statistical methods
according to the Partial Safety Factor Method for Ultimate Limit State verification.

2. EXPERIMENTAL RESULTS OF BOND TESTS ON FRP SHEETS BONDED ON


MASONRY BLOCKS
2.1 Influence of Set-up in Bond Tests
In order to evaluate both bond strength and bond stress distribution, experimental tests have
been performed by several researchers according to different set-ups: beam-tests, single and
double shear tests. In the beam test set-up, usually adopted for characterization of bond
behaviour of steel rebars in concrete elements according to the RILEM standards, the tension
force is applied to the FRP reinforcement adopting a flexural scheme. However, the loading
pattern of the beam test can cause a shear failure in the blocks for small length/height ratios
[1] that does not allow to obtain the actual bond strength.
By contrast, in single and double shear tests, the tensile force is directly applied to the
external FRP reinforcement, glued on one side only (single scheme) or on two opposite sides
of the masonry block (double scheme). A further distinction within the shear tests is related to
whether the masonry block is subjected to a tensile or compressive force due to the restraint
in the loading application. Due to the low tensile strength of the masonry block, the scheme
providing compression to the block is the most commonly used for both single or double
configuration. Moreover, the single shear test is expected to give more reliable results rather
than the double configuration, being less influenced by imperfections of the alignment of
FRP reinforcement in the testing machine.
Dissimilar set-ups could lead to quite different debonding load and, thus, result from different
experimental programs are not always directly comparable. For these reasons, the
experimental database presented in this paper is related only to single and double shear tests
where the masonry blocks are loaded in compression.

2.2 The Database of the Experimental Tests


Being this study aimed at defining the maximum load that the FRP reinforcement can
sustain at debonding, the results of specimens with bonded length, Lb, smaller than the
effective ones have been neglected. The effective length is the bonded length needed to
328

completely transfer the maximum load [2, 3, 4] and depends mainly on the tensile strength
of the support, the Youngs modulus and thickness of the reinforcement. From observation
of experimental distribution of strains during the loading application, the researchers
confirmed that the effective length for masonry elements is never smaller than 150 mm;
thus, only results of experimental tests with bonded lengths larger than 150 mm have been
considered (see Table 1). Moreover, only tests where failure was due to debonding at the
FRP masonry interface were selected. In Table 1, the range of variability of geometrical
parameters of FRP and masonry blocks considered in the experimental database are
indicated separately for the natural stones and clay bricks.
For the natural stones, a wide variability of mechanical properties is expected since various
materials have been employed for the tests: yellow and grey Neapolitan Tuff [1, 5], Leccese
stone [6], Sicilian calcarenite stone [7]. Yellow and grey tuff are volcanic stones made of
consolidated ash ejected from winds during volcanic eruptions. Tuff is a typical stone of the
Southern Italy, and has been diffusely used as building material in the past for its good
mechanical properties in compression and the low unit weight. Calcarenite stones are
limestones, formed by the percolation of water through a mixture of calcareous shell
fragments and quartz sand causing the dissolved lime to cement the mass together.
Limestone is very common in architecture, especially in Europe and North America. The
Leccese stone is a white calcarenite stone characterized by high porosity and very good
mechanical properties; this stone is made of regular calcareous silts mixed with small
particles of clay minerals, is a very compact stone, but low cemented. This stone is very
common in the Puglia region (Southern Italy).
The compressive strength of the natural stones of the collected database [1, 5, 6, 7] varies in
the ranges: 2.7-5.0 MPa for the yellow tuff, about 4.0 MPa for the grey tuff, 2.2-11.3 MPa for
the Sicilian calcarenite, and 24 MPa for the Leccese stone. Moreover, the compressive
strength of the clay bricks [8, 9, 10, 11, 12, 13, 14] varies in the range 7-42 MPa.
The compressive strength of masonry blocks is usually an experimental data given by the
researchers, on the contrary, the tensile strength is not always known by direct experimental
results and, thus, it has been assumed always as the 10% of the compressive strength in the
present study in order to use homogenized data in the statistical analysis
Because the natural stones used in the database may have very different mechanical
(strength, Youngs modulus) and physical (porosity, surface texture) properties, a further
distinction within the natural stones has been done. In particular, the tuff stones (yellow and
grey) have been considered separately from the limestones (Leccese stone and Sicilian
limestone). About the geometrical and mechanical properties of the FRP reinforcements, the
values given by the manufacturers have been usually reported in the collected papers, and,
thus, these values were assumed (see Table 1). Both glass and carbon fibres were used in the
tests and all specimens were strengthened with only 1 layer.
Table 1. Geometrical parameters of specimens and mechanical properties of FRP sheets.
Support
Lb [mm]
bf [mm]
tf [mm]
Ef [GPa]
bf/bm
Natural stones
150200
50100
0.330.80 0.1650.350 82246
Clay bricks
150305
40152
0.21
0.0970.480 65240
329

3. BOND STRENGTH FOR FRP SYSTEMS EXTERNALLY BONDED ON


BRITTLE SUPPORTS
If the failure of the external FRP reinforcement is due to debonding in the masonry support,
the maximum transferable load, depending on the bond strength, is related to the tensile
strength of masonry block and can be evaluated using the fracture mechanics approach.
For FRP reinforcement bonded on a brittle support the maximum transmissible force, in
case of anchorage with infinite length, and the fracture energy depend on the bond-slip law
at the interface area, b(s), and can be expressed as follows:
Fmax b f

0 b xdx

0 b s ds

(1)

Being respectively bf the width of the reinforcement, b and s the bond shear stresses and
the slip along the FRP-support interface. It can be demonstrated the following relation
holds between the maximum force and the fracture energy:

Fmax b f 2 E f t f F

(2)

being tf and Ef thickness and Youngs modulus of the FRP reinforcement. Eq. (2) is valid for
a general bond-slip law expressed by a sufficiently regular function [15].

3.1 Bond strength for FRP bonded to concrete support


In order to determine the maximum shear stress of an interface bond law b(s), a concrete
failure criterion is needed. Adopting the Mohr-Coulomb criterion, the maximum shear stress,
max, is a function of both compressive and tensile strength of concrete, but the expression is
different in case of failure in pure shear loading condition [16] or in presence of normal stress
component [17]. The presence of normal compressive stresses parallel to the bonding surface
in the concrete surrounding the bonded area has been evidenced by numerical studies [18]
and by the visual inspection of the debonded concrete surface after experimental tests [19]. In
the present paper, the maximum shear stress is calculated as [17]:

max kb f c f ct

(3)

This term is twice the cohesion in the Mohr-Coulomb criterion and is motivated by the
presence of a confinement normal stress. In Eq. (3), kb is a shape factor modelling the effect
of the plate-to-concrete width ratio, bf/bc; various corrective shape factors are suggested by
the literature, as the following ones:
kb 1.06

2 b f / bc
1 b f / 400

(a)

or w

2 b f / bc
1 b f / bc

(b)

(4)

The first expression has been adopted by the Fib bulletin 14 [2] and Italian Guidelines [3],
while the second one has been suggested in [4]. The shape factor models the following
330

phenomenon: when the FRP width is smaller than the concrete support width (bf/bc < 1), the
bond stresses may extend over a width larger than the FRP width bf and a larger volume of
concrete is involved in the fracture phenomenon. Moreover, as bf/bc is close to 1, this effect
vanishes. Therefore, taking into account the shape factor, kb, the fracture energy of the b(s)
law can be written in the form:

F kG kb

f ck f ctm

(5)

where kG is a coefficient which must be calibrated thanks to experimental results of bond


tests. Substituting Eq. (5) in Eq. (2), the maximum transferable load is defined as a function
of the FRP axial stiffness, the concrete strength and the plate-to-concrete width ratio.

3.2 Bond strength for FRP bonded to masonry supports


For masonry elements externally strengthened with FRP composites, the Italian guidelines
[3] proposed a design formula for the maximum transmissible load similarly to that in the
case of concrete support (see Eq. (2)). The characteristic value of the fracture energy for
composite bonded to masonry bricks was defined similarly to concrete as follows:

Fk c1 f mk f mtm

(6)

being fmk and fmtm the characteristic compressive and the mean tensile strength of the masonry
blocks, respectively.
At that time (2004) very few experimental results were available and the masonry supports
considered in the statistical analyses were very different to each other. This led to a design
formulation less reliable than the case of concrete support. Lacking experimental
evaluations of the tensile strength, it can be assumed as fmtm = 0.1fmk. The coefficient c1
should be assessed according to experimental data analysis and, if no experimental tests are
available, the value c1 = 0.015 mm is suggested in [3].
If Eq. (6) is introduced in Eq. (2), the debonding load depends, as in the case of concrete
elements, on the FRP axial stiffness and on the strength of the masonry support, while the
effect of the plate-to-masonry block width ratio, bf/bm is neglected.
In order to verify the validity of the formulation reported in [3], it has been applied to the new
experimental database described in section 2. Note that the value of the coefficient c1 = 0.015
mm in Eq. (6) gives the 0.05 percentile of the fracture energy, so this provision is expected to
give a conservative result with respect to the experimental values. The comparison with
results of bond tests on natural stones and clay bricks are reported separately in Fig. 1a and
Fig. 1b, respectively. Moreover, the natural stones have been divided in tuff and limestone.
The mean values of the experimental compressive and tensile strengths of masonry blocks are
used in Eq. (6).

331

20
Tuff stones [1, 5]

Fmax,exp
[kN]

30
Fmax,exp
[kN]

Limestones [6, 7]

10

clay bricks [8-14]

15

0.5 0.5

Fmax,CNR,5% = bf (2 Ef tf (0.015 (fm fmt) ))

Fmax,CNR,5% = bf (2 Ef tf (0.015 (fm fmt)0.5))0.5 [kN]

0
0

10

[kN]

0
0

20

(a)

15

30

(b)

Fig. 1. Experimental data vs. characteristic debonding load (5% percentile) according to
[3]: a) natural stones (tuff stones and limestones); b) clay bricks.
These comparisons show a different safety degree, depending on the stone type, of the
experimental results with respect to the 0.05 percentile force value. The experimental results
are underestimated especially in the case of tuff and clay bricks. This leads to suppose that c1
should depend on the stone type and that the debonding load must be related not only to the
strength of the stone, but also on the type. In fact, even if the compressive strength is the
same, the natural stones used for constructions can be very different in terms of density,
porosity, surface texture due to dissimilar geological processes. These factors influence
differently the bond behaviour which is related to the fracture energy of the material. Hence,
with reference to Eq. (6), the coefficient c1 must be calibrated separately with reference to
homogeneous sets of support materials. Moreover, CNRs formulation does not take into
account the effect of the FRP-to-masonry width ratio, that is expected to be significant in
masonry elements too [20].
In the following, a new design formula taking the different typologies of masonry and the
effect of the FRP-to-masonry block width into account is proposed. In order to calibrate the
new formulation, the wide database of experimental results concerning bond tests on masonry
elements externally bonded with FRP sheets previously described has been used.
The first step to assess the new formulation is the definition of the shape factor. Eqs. (4) are
commonly used for concrete elements, but both of them must be limited to 1 when the ratio
bf/bc is close to1. Results of bond tests on clay bricks strengthened with carbon FRP sheets
with variable width and, thus, variable bf/bm ratios have been used to calibrate the following
expression for the shape factor (Reluis project internal report):
kb

3 b f bm

(7)

1 b f / bm

This expression is a continuous function defined in the range [0, 1] and gives values of kb
ranging from 1.73 (limit for bf/bm approaching 0) to 1 (for bf/bm = 1).

332

4. THE PROPOSED BOND STRENGTH MODEL


The bond strength model is defined by the following expression:

~
F kG kb

~
Fmax,th b f 2 E f t f kG kb f m f mt

f m f mt [F/L]

[F]

(8)

where kG is a parameter to be determined experimentally and the other symbols have the
meaning previously indicated. Note that fm and fmt are the mean experimental compressive
and tensile strength of the masonry block, while in Eq. (6) the characteristic compressive
strength is requested. Moreover, the shape factor kb is introduced as defined in Eq. (7).
~

The strength model of Eq. (8) should be fine-tuned by calibrating the parameter kG
minimizing the difference between the theoretical, Fmax,th,i, and the experimental, Fmax,exp,i,
debonding force for the i-th test. According to the Design by testing Eurocode [21]
approach, a coefficient a is defined by means of the regression line through the origin of the
graph Fmax,th,i vs. Fmax,exp,i. Its slope gives the least-square coefficient, a, defined as:
a

F
F

max, exp, i
max, th , i
2
max, th , i

(9)

Thus, the corrected strength model becomes:

~
Fmax,th,m a Fmax,th a b f 2 E f t f kG kb f m f mt b f 2 E f t f kG ,m kb f m f mt

~
F , m a 2 kG kb

f m f mt k G ,m k b

f m f mt

(a)

(b) (10)

assuming a 2 kG kG , m and being Fmax,th,m and F,m the mean provisions for the debonding
load and fracture energy. Eq. (10a) is then written in a form very similar to Eq. (5) adopted
for FRP-concrete bonding. The values of coefficient kG,m are reported in Table 2 for FRP
bonding on the three different kinds of masonry supports. In order to measure the error of
the model represented by Eq. (10), a random variable is defined as the ratio of the
experimental debonding load Fmax,exp,i to the theoretical mean value Fmax,th,m,i. In the
probabilistic framework, the strength can be assumed as a random variable and, in general,
the 0.05 percentile (named characteristic value) of its frequency distribution is used for
design purposes. A Gaussian distribution can be adopted for the natural logarithm of the
strength value; then a new variable is defined as ln() .
In order to evaluate the percentile values of the debonding load, the coefficient of variation
of the variable Fmax,th has been calculated, neglecting the variance of the mechanical
properties of materials and considering only the variance of the variable . The
characteristic values of kG, debonding load and fracture energy can be calculated as [21]:

k G ,k kG ,m exp k s 0.5 s2

(a)

333

Fk kG , k kb

Fmax,th, k b f 2 E f t f kG , k kb f m f mt (b)

f m f mt (c)

(11)

where s is the standard deviation of the variable , k is the coefficient of the Gaussian
distribution to calculate the 0.05-percentile. If the number of experimental data n > 100, k
= 1.645. The condition n > 100 is verified for the bond test data on limestones and clay
bricks, while for tuff stones, being n = 20, the value kn = 1.680 is assumed instead of k.
The 0.05 percentile provisions of parameter kG,k for the three types of stone are reported in
Table 2. In Figures 2 a-c, the experimental debonding loads are plotted versus the mean and
the 0.05 percentile predictions given by Eqns. (10) and (11), respectively. In order to compare
the bond behaviour of different masonry supports, the experimental values of the fracture
energy, obtained assuming in Eq. (2) the experimental values of the debonding forces, are
reported in Fig. 2d. The characteristic values of the fracture energy given by Eq. (11c) is also
reported in Fig. 2d. The comparison confirms bond strength of FRP over limestones is
significant smaller than in the case of bonding over tuff stones and clay bricks.
Table 2: FRP-bond tests on different masonry supports: number of tests adopted for
calibration procedure; mean and 0.05-percentile values of kG.
Supports
n kG,m [mm] kG,k [mm]
Artificial stones Clay bricks
145
0.093
0.031
Tuff stones
17
0.097
0.032
Natural stones
Limestones and Leccese stones 65
0.022
0.012
Fmax,exp
[kN]
40

15

Clay bricks [8-14]


mean
5% percentile

Yellow tuff [1, 5]


Grey tuff [5]
mean
5% percentile

Fmax,exp
[kN]
10

20

5
Fmax,th,m [kN]

Fmax,th,m [kN]

0
0

15

20
Leccese stone [6]
Sicilian limestone [7]
mean
5% percentile

Fmax,exp
[kN]
10

40
F

(a)

10

15

bricks: 5% percentile
(b) Clay
Tuff stones: 5% percentile

[N/mm]

2.0

Limestones: 5% percentile
Clay brick: exp. data [8-14]
Tuff stones: exp. data [1,5]
Limestones: exp. data [6,7]

1.0

5
Fmax,th,m [kN]
0

(c)

10

(f m f mt )0.5 [N/mm2]

0.0

15

15

(d)

30

Fig. 2: Experimental loads vs. mean and 0.05 percentile predictions: a) Clay bricks; b) Tuff
stones; c) Limestones; d) Fracture energy versus strength for different masonry supports.

334

5. CONCLUSIONS
With reference to FRP externally strengthened masonry elements, the paper firstly points out
the clear correlation between FRP-masonry debonding phenomena and the main concepts of
fracture mechanics, which are the basis of the design criteria against debonding introduced by
Italian Recommendations CNR DT200/2004 [3]. Several experimental data available in the
scientific literature, concerning bond tests on different types of masonry blocks, have been
used to calibrate a prediction formula for the debonding load according to the Design by
Testing procedure suggested by the Eurocode [21]. In the new formula, different design
rules have been defined for the bond strength over natural and artificial stones.

6. ACKNOWLEDGEMENTS
Financial supports of (Italian) Department of Civil Protection (Reluis 2010 Grant Task 3:
Technological innovation for seismic engineering) is gratefully acknowledged.

7. REFERENCES
1.

Ceroni F. and Pecce M. 2006. Bond tests on concrete and masonry blocks externally
bonded with CFRP. Proceedings of the 3rd Int. Conf. on FRP Composites in Civil
Engineering, CICE 2006, Miami, Florida, 13-15 December, pp. 17-20.
2. fib. 2001. Externally Bonded FRP Reinforcement for RC Structures. fib Bulletin 14,
Technical Report, Task Group 9.3 - FRP Reinforcement for Concrete Structures.
3. CNR-DT 200. 2004. Guide for the Design and Construction of Externally Bonded FRP
Systems for Strengthening, Council of National Research, Rome.
4. Chen J.F. and Teng J.G. 2001. Anchorage strength models for FRP and steel plates
bonded to concrete, Journal of Structural Engineering, ASCE, 127 (7): 784-791.
5. Aiello M.A. and Sciolti M.S. (2006). Bond analysis of masonry structures strengthened
with CFRP sheets, Construction and Building Materials, 20 (1-2): 90100.
6. Aiello M.A. and Sciolti M.S. 2008. Analysis of bond performance between CFRP
sheets and calcarenite ashlars under service and ultimate Conditions, Masonry
International, 21 (1): 15-28.
7. Accardi M., Cucchiara C., Failla A., and La Mendola L. 2007. Bond behavior between
CFRP strips and calcarenite stone. Proceedings of the 6th International Conference
Fracture Mechanics of Concrete and Concrete Structures, FraMCoS-6, Catania, 17-22
June, pp. 1203-11.
8. Capozucca R. 2010. Experimental FRP/SRPhistoric masonry delamination.
Composite Structures, 92 (4): 891-903.
9. Casareto M., Oliveri A., Romelli A., and Lagomarsino A. 2003. A bond behaviour of FRP
laminates adherent to masonry. Proceedings of the International Conf. Advancing with
Composites, Milan, May.
10. Faella C., Martinelli E., Nigro E., Paciello S. 2008. Adhesion of advanced composites on
masonry: experimental study and numerical analysis. Proceedings of the 4th International
Conference on FRP Composites in Civil Engineering - CICE 2008, Zurich, 22-24 July,
CD ROM.
11. Oliveira D. V., Basilio I., and Loureno P.B. 2011. Experimental bond behavior of
FRP sheets glued on brick masonry, Journal of Composites for Construction, ASCE,
335

15(1): 32-41.
12. Panizza M., Garbin E., Valluzzi M.R., and Modena C. 2008. Bond behaviour of CFRP and
GFRP laminates on brick masonry. Proceedings of the 6th International Conference on
Structural Analysis of Historical Constructions, Bath (UK), 2-4 July 2-4, eds. DAyala &
Fodde, pp. 763-770.
13. Seim W. and Pfeiffer U. 2010. Bonding of FRPs on masonry surfaces. Proceedings of
the 8th International Masonry Conference, Dresden.
14. Briccoli Bati S. and Fagone M.. 2010. An analysis of CFRP-brick bonded joints. XVIII
GIMC Conference, Siracusa, 22-24 September.
15. Ferracuti B., Savoia M., and Mazzotti C. 2007. Interface law for FRP-concrete
delamination. Composite Structures, 80 (4): 523-531.
16. Brosens, K. 2001. Anchorage of externally bonded steel plates and CFRP laminates for
the strengthening of concrete elements. Doctoral thesis, K.U.Leuven.
18. Freddi F., and Savoia M. 2008. Analysis of FRP-concrete debonding via boundary
integral equations, Engineering Fracture Mechanics, 75 (6): 1666-1683.
19. Mazzotti C., Savoia M., and Ferracuti B. 2008. An experimental study on delamination of
FRP plates bonded to concrete. Construction and Building Materials, 22 (7): 1409-1421.
20. Fedele R. and Milani G. 2011. Three-dimensional effects induced by FRP-frommasonry delamination, Composite Structure, doi:10.1016/j.compstruct.2011.01.022.
21. Eurocode. 2002. Eurocode: Basis of Structural Design. EN 1990:2002 E.

336

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

BRIDGE DECK-GUARDRAIL ANCHORAGE DETAILING FOR


SUSTAINABLE CONSTRUCTION
K. Sennah1, N. Nikravan2, J. Louie3, A. Hassan3, N. Al-Bayati4, M. El-Sayed3 and M.
Sayed-Ahmed2
1

Professor, Civil Engineering Department, Ryerson University, Toronto, Ontario, Canada


Graduate student, Civil Engineering Department, Ryerson University, Toronto, Ontario
3
Undergraduate student, Civil Engineering Department, Ryerson University, Ontario
4
Non-degree researcher, Civil Engineering Department, Ryerson University, Ontario
2

ABSTRACT
A GFRP bar bent was recently developed to overcome the problems of tensile stress
reduction at the bent. It is made of corrugated plastic tube formed with required shape of
the bent. The tube is then filled with GFRP material in liquid conditions and left to dry.
This developed GFRP bent was proposed to be used in one of the new bridges in Ontario.
This paper investigates the use of GFRP bar bents as stirrups at the joint between the steel
post of the bridge guard-rail system with the deck slab cantilever. In addition, GFRP bars
with headed ends are utilized for better anchorage at the post-deck slab joint. Two full scale
cantilever post specimens were erected and tested to-collapse. The first specimen cantilever
was reinforced with reinforcing steel bars, while the second one was reinforced with GFRP
straight bars, bent bars and those with headed ends as applicable locations. Results showed
that GFRP-reinforced specimen is as good as that reinforced with steel bars. All specimens
failed due to excessive torsional-shear cracks at the post-curb junction, resulting in spalling
of the concrete cover at the outer face of the curb at the steel post location.

1. INTRODUCTION
Fibre-reinforced polymers (FRPs), as non-corrodible materials, are considered as excellent
alternative to reinforcing steel bars in bridge decks to overcome steel corrosion-related
problems. FRP materials possess the necessary property of high tensile strength of 1100
MPa that makes them suitable as structural reinforcement for concrete. The special
ribbed surface profile of the GFRP bar shown in Fig. 1, ensure optimal bond between
concrete and the bar. Until recently, the installation of GFRP bars was often hampered by
the fact that bent bars have to be produced in the factory since GFRP bars cannot be bent at
the site. Also, bent GFRP bars are much weaker than straight bars, due to the redirection
and associated rearrangement of the fibres in the bend. As a result, number of bent GFRP
bars is increased and even doubled at such locations where bar bents are required. In case
of bridge deck slab and cantilevers, GFRP bars with headed ends are used as straight bars
337

with an end head to reduce their development length, thus avoiding the use of hooks. The
head is made of a thermo-setting polymeric concrete with a compressive strength far
greater than that of normal grade concrete. It is cast onto the end of the straight bar and
hardened at elevated temperatures. The concrete mix contains an alkali resistant Vinyl Ester
resin, the same material used in the straight bars, and a mixture of fine aggregates. The
maximum outer diameter of the heads is 2.5 times the diameter of the bar. The head of the
16 mm bar is approximately 100 mm long. It begins with a wide disk which transfers a
large portion of the load from the bar into the concrete. Beyond this disk, the head tapers in
five steps to the outer diameter of the blank bar. This geometry guarantees optimal
anchorage forces and minimal transverse splitting action in the vicinity of the head.

a) Headed bar
b) Bent bar
c) A slice in the bent bar
Fig. 1. Views of the GFRP bars with headed ends of bents
The Canadian Highway Bridge Design Code, CHBDC, (CSA, 2006) allows the use of
GFRP bent bars in bridges. Such bents reduce the tensile strength of the bar considerably
based on results from experimental testing. To reduce the effects of the bent on tensile
stresses, a newly developed GFRP bar bent was made of corrugated plastic tube formed
with required shape of the bent. The tube is then filled with GFRP material in liquid
conditions and left to dry. This developed GFRP bent was proposed to be used in one of the
new bridges in Ontario. These GFRP bents are used in the current research to strengthen the
deck-slab curb for torsional-shear.

2. RESEARCH APPROACH AND TEST SAMPLES


In bridge construction, as that shown in Fig. 2, concrete deck slab is cast in-place after the
girders are placed over the abutments and piers. Then, the sidewalk is cast over the deck
slab and exterior girder using dowels projecting from the slab. The steel posts and
guardrails are then assembled. This research investigates the joint between the steel post of
the bridge guard-rail system and the slab cantilever. Both GFRP bent bars, used as stirrups
for the curb under the post location, and the headed studs for bar anchorage were used to
reinforce the deck slab cantilever-steel post joint, as well as the top reinforcement of the
cantilever slab to eliminate the effects of de-icing salts used in the winter times in Canada.
Two full scale cantilever post specimens were erected and then tested to-collapse at the
structures laboratory of Ryerson University. Figure 3 shows dimensioning of the cantilever
slab specimens. The first specimen was a control specimen with the deck slab reinforced
338

using reinforcing steel bars. While the second specimen was identical to the first one except
that it was reinforced with GFRP straight bars, bent bars and those with headed ends. The
slab cantilever was 2.5x2.5 m in plan, with the steel post mounted near the tip of the
cantilever at the mid-width of the specimens.

Fig. 2. Schematic diagram of showing post-deck cantilever joint under consideration

Fig. 3. Dimensioning of the specimen


Figure 3 shows that there is a utility duct, embedded in the slab at 750 mm distance from
the outer face of the deck slab cantilever tip. This utility duct was made of 89-mm diameter
PVC tube. To facilitate applying horizontal load near the top of the steel post, the simulated
fixed end of the deck slab cantilever was enlarged and four, 50-mm diameter, PVC sleeves
were embedded in the enlarged end, as shown in Fig. 3, at 600 mm spacing. These ducts
were be used to apply tie-down system to ensure full fixity of the deck slab cantilever end.
Figure 4 presents the details the steel post, base plate, anchor bolts and anchorage system as
specified in AASHTO Specifications for guardrails (Sennah et al., 2011). Figures 5 and 6
show the bar details of the steel specimen and the GFRP specimen, respectively.

339

Fig. 4. Views of the fabricated post anchorage system

Fig. 5. Cross-section of steel-reinforced specimen showing steel bar arrangement

Fig. 6. Cross-section of FRP-reinforced specimen showing steel and FRP bar arrangements

340

a) Steel specimen
b) GFRP specimen
Fig. 7. Views of bar arrangement and formwork

Fig. 8. More views of GFRP bar arrangement and formwork for GFRP specimen-2

Fig. 9. Schematic diagram of the test setup

3. TEST SETUP, INSTUMENTATION AND TEST PROCEDURE


All specimens were fabricated in the structures laboratory of Ryerson University. First,
timber formwork was assembled. Then, reinforcing bars were assembled, followed by the
installation of the anchorage system and anchor bars. Concrete casting was performed using
341

ready mix concrete. Figure 9 shows schematic diagram of the test setup. Four 44.45 mm
diameter (1 ) all-threaded rods were inserted in the PVC embedded tubes at the enlarged
end of the slab cantilever to tie the specimen to the laboratory strong floor. Steel HSS
beams fixed to the laboratory floor and two hydraulic jacks, shown in Fig. 9, were used to
prevent rigid body movement of the specimen when applying horizontal load on the steel
post. The lateral load was applied to the steel post at 790 mm from the top surface of the
concrete slab. LVDTs were installed at predetermined locations in the specimens as shown
in Fig. 9 to record lateral and vertical displacement. These locations are (i) vertical
deflection of the cantilever tip, (ii) horizontal deflection of the steel post at 790 mm from
top surface of the concrete slab, (iii) the uplift at the location of the tie-down system, (iv)
the rigid body movement of the specimens at the inner side of the enlarged portion of the
deck slab, and (v) the horizontal displacement of the cantilever tip. Figures 10 and 11 show
views of the steel and GFRP specimens, respectively, before testing. Visual inspection was
conducted at each load increment during the test to record any change in structural integrity
of the specimen. Initial crack and crack propagation were recorded during testing. Failure
of the specimen was considered reached when the specimen could not absorb more load.

Fig. 10. View of test setup of steel specimen

Fig. 11. View of the GFRP specimen

Fig. 12. Views of crack pattern at 150 kN load and after failure for the steel specimen

Fig. 13. View of the failed steel specimen showing exposed steel bars after removing the
cracked concrete cover concrete
342

4. TEST RESULTS
Five concrete cylinders were tested in axial-compression to-collapse at the time of testing of
each specimen. The characteristics strength of concrete was calculated based on Clause
A14.1.2 of the Commentaries of the Canadian Highway Bridge Design Code. As such, the
characteristic strengths of concrete at the time of testing were 34.27 and 32.87 MPa for
steel specimen and GFRP specimen, respectively. All specimen exhibited first crack
starting near the rear corner of the base plate and propagated towards the front face of the
slab curb at an angle following the torsion-shear crack pattern. Other minor cracks
appeared at the rear end of the anchors in tension and the inner edge of the curb as expected
from the splitting forces in concrete resulting from pull-out of the anchors embedded in
concrete. With increase in applied load, cracks propagated towards the bottom tip of the
front face of the slab curb and cracks were widened to the extent that the concrete cover to
reinforcement, along the post base plate length, to spall. No sign of failure in GFRP or steel
reinforcement appeared after testing. However, steel stirrups in case of steel specimens and
GFRP bents and headed ends of the diagonal bars in case of the GFRP specimens were
exposed after spalling of the concrete cover.

Fig. 14. Views of the extended cracks at 160 kN load and at failure for the GFRP specimen

Fig. 15. View of the spalled concrete at the front side of the curb of GFRP specimen
Figure 12 show photos of the crack pattern along with the jacking loads at which these
cracks first started or propagated for the steel specimen. While Fig. 13 shows the spalled
concrete at the front face of the curb after failure. It can be observed that the first diagonal
crack appeared at the top surface of the curb at 80 kN, while the specimen failed at 151.76
kN. Figure 14 shows photos of the crack pattern along with the jacking loads at which
these cracks first started or propagated for the GFRP specimen. While Fig. 15 shows the
spalled concrete at the front face of the curb after failure. It can be observed that the first
343

diagonal crack appeared at the top surface of the curb at 90 kN. However, the specimen
failed at 169.11 kN. It is worth mentioning that flexural cracks on the top of the deck slab
cantilever appeared only at the ultimate load in the GFRP specimen as depicted in Fig. 14.
Such cracks did not appear in the steel-reinforced specimen. No indication of break in the
steel anchors was observed, however, steel anchors in the tension side appeared to be
deformed in bending. Figures 16 and 17 depict the jacking load-displacement relationship
for the steel specimen and the GFRP specimen, respectively. Displacement readings
showed that the tie-down system was effective in providing fixity at the enlarged portion of
the slab. In addition, insignificant rigid body movement of the specimens was observed. It
can also be observed that the deflections of the post were close to each other for two
specimens.

Fig. 16. Jacking load-displacement


relationship for steel specimen-2

Fig. 17. Jacking load-displacement


relationship for GFRP specimen-2

5. CONCLUSIONS
Results presented in the report may conclude that the GFRP-reinforced specimen is as good
as the steel-reinforced specimen. The primary failure mode is due to combined torsion and
direct compression resulting from the compressive stresses at the outer portion of the post
steel base plate.

6. REFERENCES
1.
2.

Canadian Standard Association (CSA). 2006. Canadian Highway Bridge Design Code.
CAN/CSA-S6-06, Mississauga, Ontario, Canada 733 p.
Sennah, K., Nikravan, N., Louie, J., Hassaan, A., El-Sayed, M., and Al-Bayati. 2011.
Experimental Study on Bridge Deck-Guard Rail Anchorage Incorporating Both GFRP
Bent Bars and GFRP with Headed Studs. Report submitted to Shoeck Canada Inc, 104
p.

344

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

MORE THAN 10 YEARS SUCCESSFUL FIELD APPLICATIONS OF


FRP BARS IN CANADA
B. Drouin1, G. Latour2 and H.M. Mohamed3
1

Pultrall Inc. Thetford Mines, QC, Canada


Trancels-Pultrall Canada Inc. Barrie, ON, Canada
3
NSERC-Postdoctoral Fellow, Research and Development, Pultrall Inc/ Department of Civil
Engineering, Faculty of Engineering, University of Sherbrooke, QC, Canada
2

ABSTRACT
Recently, there has been a rapid increase in using noncorrosive fibre-reinforced polymers
(FRP) reinforcing bars for concrete structures such as bridges, parking garages, and marine
structures in which the corrosion of steel reinforcement has typically led to significant
deterioration and rehabilitation needs. Many significant developments from the manufacturer,
various researchers and Design Codes along with numerous successful installations have led to
a much higher comfort level and exponential use with designers and owners across Canada.
After years of investigating, public agencies and regulatory authorities as the Public Works
and Government Services Canada (PWGSC) and Ministry of Transportation at different
provinces across Canada has now included FRP as a premium corrosion resistant reinforcing
material in its corrosion protection policy. In the last decade, the field applications indicated
that the V-ROD FRP rebar has been used as internal reinforcements in more than 100 bridges
across Canada. This paper presents a summary of different field applications of V-ROD FRP
reinforcements in different types of concrete structures. In addition, the new accessories and
continuing research and development in the field of FRP reinforcement will be presented.

1. INTRODUCTION
Electrochemical corrosion of steel is a major cause of the deterioration of the civil engineering
infrastructure. It is becoming a principal challenge for the construction industry world-wide.
Canada Construction Association (CCA) has estimated that the global infrastructure loss
would be in the vicinity of $900 billion. An effective solution to this problem is the use of
corrosion resistant materials, such as high-performance fibre-reinforced polymer (FRP)
composites, (Benmokrane et al. 2002). The applications of FRP reinforcements in the last 10
years have been approved that the cutting-edge technology has emerged as one of the most
cost-effective alternative solutions compared to the traditional solutions. The use of concrete
structures reinforced with FRP composite materials has been growing to overcome the
common problems caused by corrosion of steel reinforcement. The climatic conditions where
large amounts of salts are used for ice removal during winter months may contribute to
345

accelerating the corrosion process. These conditions normally accelerate the need for costly
repairs and may lead to catastrophic failure. Known to be corrosion resistant, FRP bars
provide a great alternative to steel reinforcement. FRP materials in general offer many
advantages over the conventional steel, including one quarter to one fifth the density of steel,
no corrosion even in harsh chemical environments, neutrality to electrical and magnetic
disturbances, and greater tensile strength than steel (Benmokrane et al. 2006; 2007; El
Salakawy et al. 2003a).
The objective of this paper is to show that V-ROD FRP rebar is on its way toward gaining
widespread acceptance in Canada. Clearly, the most tangible successes are in the area of
highway reinforced concrete bridge structures, especially at Ontario and Quebec provinces in
where the corrosion resistance of FRP reinforcements as well as their installation flexibility
are taken advantage of. In the following sections, development of codes and guidelines, field
applications across Canada of V-ROD FRP-bars throughout the last ten years are presented. In
addition, the new accessories and continuing research and development in the field of FRP
reinforcement will be presented.

2. DEVELOPMENTS OF CODES AND GUIDELINES


In North American, several codes and design guidelines for concrete structures reinforced with
FRP bars have been published from 2000 to 2010. In 2000, the Canadian Highway Bridge
Design Code (CHBDC) [CAN/CSAS6-00, (CSA 2000)] has been introduced including
Section 16 on using FRP composite bars as reinforcement for concrete bridges (slabs, girders,
and barrier walls). Design manual (ISIS-M03-2001) for reinforcing concrete structures with
FRP was presented by the Canadian Network of Centres of Excellence on Intelligent Sensing
for Innovative Structures (ISIS). In 2002, CAN/CSA-S806-02 has been published by the
Canadian Standards Association (CSA 2002) for design and construction of building
components with FRP reinforcements. The American Concrete Institute (ACI) introduced the
first and second guideline (ACI 440.1R) for the design and construction of concrete reinforced
with FRP bars in 2001 and 2003, respectively. As a result of the valuable, enormous and great
research efforts on different types of FRP-reinforced concrete structures in world wide during
the last decade, the aforesaid North American codes and design guidelines have been updated
and modified to encourage the construction industry to use FRP materials [ISIS-M03-2007;
CAN/CSAS6-06; CAN/CSAS6-06 edition 2010; ACI 440.1R-03; ACI 440.1R-06]. Currently
out for comment, a revised ACI 440.1R-06 and CAN/CSA-S806-02 is forthcoming later in
2011.
In order to establish stringent guidelines and values for FRP manufacturers and quality control
mechanisms for owners to ensure a high comfort level of product supplied, ISIS Canada
together with the manufacturer had initiated the Specifications for product certification of
FRPs as internal reinforcement in concrete structures. (ISIS Canada Corporation 2006) This
document was the basis for the new FRP specification (CSA S-807-10), which is now
available from the CSA website since 2010.

346

3. FIELD APPLICATIONS ACROSS CANADA


3.1 Highway Bridge Structures
The corrosion of steel reinforcing bars in concrete bridge decks, which leads to excessive
cracking, spalling, reduced strength, and loss of structural integrity, constitutes a major
problem when measured in terms of rehabilitation costs and traffic disruption. Concrete bridge
decks deteriorate faster than any other bridge components because of direct exposure to
environment, deciding chemical and ever-increasing traffic load. In Quebec and Ontario,
around half of the maintenance budget of the Ministry of Transportation (MT) is spent on
concrete structures damaged by corrosion of steel. Therefore, since the late 1990s, the
Structures Division of the MT at different provinces has been interested in building more
durable bridges with an extended service life of 75150 years. For example, the MT at Qubec
(MTQ) has carried out, in collaboration with the University of Sherbrooke, (Sherbrooke,
Qubec), several research projects utilizing the straight and bent non-corrodible FRP rebar in
concrete deck slabs and bridge barriers (El-Salakawy et al. 2003b). The use of FRP bars as
reinforcement for concrete bridge provides a potential for increased service life and economic
and environmental benefits.
In the last ten years, the V-ROD FRP bars have been used successfully in many bridge
structures across Canada. From 1994 up to date, there is more than 100 concrete bridges have
been reinforced with V-ROD FRP-bars. Straight and bent FRP bars (carbon or glass) were
used mainly as internal reinforcement for the deck slab and/or for the concrete barriers and
girders of the bridges. Figure 1 presents the successful growing and the increase of demand to
use V-ROD-FRP bars in reinforcing bridge structures across Canada. The figure shows that
the number of projects being designed and specified with FRP has exponentially increased
over the past couple of years. In 1994, Mederic Martin Bridge on Highway 15 Laval, QC was
the first bridge included with V-ROD FRP bars as reinforcement for the concrete barriers.
Taylor Bridge (Winnipeg, Manitoba, 1995) was the second field application of using FRP bars
as bridge reinforcements in Canada.
30

Number of Projects

25
20
15
10
5
0

Year

Fig. 1. Progress of using V-ROD FRP bars as reinforcement for bridges in Canada
347

In general, all the bridges that included with V-ROD reinforcements though the ten years ago
are girder-type with main girders made of either steel or prestressed concrete. The main
girders are simply supported over spans ranging from 20.0 to 90.0 m. The deck is a 200 to 260
mm thickness concrete slab continuous over spans of 2.30 to 4.0 m. Most of these bridges
have been reinforced with the glass FRP bars as a result of their relatively low cost compared
to other types of FRPs (carbon and aramid). The FRP bars were used mainly as reinforcement
to the deck slabs, barriers and girders.
The Joffre Bridge project, in Sherbrooke (Qubec, 1997), over the St-Franois River, was the
third bridge in Canada reinforced with FRP reinforcement. After the third successful field
application of the FRP as reinforcement to the deck slab, many other several demonstration on
real bridges reinforced with V-ROD rebar were carried out: Wotton Bridge in the municipality
of Wotton (Fig. 2), Magog Bridge (2002) on Highway 55 North (Fig. 3), and Cookshire-Eaton
Bridge (2003) on Road 108 and Melbourne Bridge (2005) in Quebec (El-Salakawy and
Benmokrane 2003 and 2004; Benmokrane 2004; El-Salakawy et al. 2003a, 2005; Benmokrane
et al. 2006). Also, Shell River Bridge was reinforced with FRP bars in 2004, Manitoba. All
these bridges were designed according to the first version of the Canadian Highway Bridge
Design Code (CSA S6-00).

Fig. 2. Wotton Bridge, QC (2002)

Fig. 3. Magog Bridge, Magoge, QC (2002)

348

In 2006, the new version of CHBDC CAN/CSA S6-06 was available and had been included
with more details about the design of concrete bridges with FRP bars. In the same year, two
bridges were constructed using FRP bars at Manitoba, (Maryland Bridge and 59th Ave Bridge
Floodway). Afterwards, in 2007 seven bridges were utilized with the V-ROD FRP
reinforcements, one in Quebec and another one in Saskatchewan, while the rest in Ontario
province. In 2009 the number of concrete bridges being designed and constructed with VROD FRP has increased to more than 22 bridges. Figures 4 and 5 show the use of GFRP bars
as reinforcement in concrete deck slab and parries for the recent two bridges at Alberta and
British Columbia, respectively. The Ministry of Transportation NW Region of Ontario has
been a leader in the use of FRP composite reinforcement in a number of cast-in-place and precast structures. In Ontario, the Noden Causeway Bridge opened in 1965 with 4.8 km long
including approaches, but was in need of some rehabilitation. The MTO opted with pre-cast
panels, and the project was divided into 3 phases. In 2008, V-ROD FRP rebar was used in
Phase 2 as reinforcement of the pre-cast pre-stressed concrete deck panels in area of
approximately 275 m length and 11m width. Pre-cast deck panels were reinforced using No. 5
and No. 6 GFRP bar with No. 4 carbon V-ROD for pre-stressing panels, see Fig 6.

Fig. 4. FRP decks and barriers Gateway Blvd/23rd Ave Alberta (2009)

Fig. 5. FRP decks/app slabs/ barriers Skagit River BC MOT (2009)

349

Fig. 6.a CFRP pre-stressing bars in forms,

Fig. 6.b Placing pre-cast panels

Fig. 6.c. Over view of the finished section of Phase 1 of Noden Causeway, ON
Eagle River Bridge (2009) is also an example of the MTO NW utilizing V-ROD in pre-cast
construction where greater quality control is possible along with simpler and quicker
installation on site. In this project, pre-cast pre-stressed box girders (87m length x 15m width)
were reinforced with V-ROD No. 5 and 8 along with CFRP bars for pre-stressing. Some No. 6
bars were also used in the approach slabs. Aside from GFRP bar which is typically used,
CFRP stirrups on some of the large girder were also used, See Fig 7. In 2010, V-ROD-FRP
straight and bent bars were utilized in more than 25 bridges in different province across
Canada. The most recent FRP-reinforced bridge in 2010 is Highway 410 Bridge located on
Highway 410 (Sherbrooke, Canada). The bridge is a girder type bridge consisting of nine steel
girders supported over one span with a total length of 46.6 m. The deck is a 225-mm thickness
concrete slab continuous over eight spans of 1.985 m each. The deck slab was reinforced with
two reinforcement mats using glass FRP bars at top and galvanized steel at bottom. In
addition, type 202 barrier walls reinforced with GFRP bars was used, See Fig 8.
Reinforced concrete bridge barriers as well as deck slab deteriorate faster than any other
bridge component because of direct exposure to aggressive environmental conditions; dicing
chemicals, in addition to impact and dynamic loads which they have to resist. The CHBDC
(CSA 2006) allows the use of GFRP bent bars in reinforcing the concrete barriers (types PL-2
and PL-3) including the connection between the slab and the barrier wall. This encourages the
use of concrete bridge decks totally reinforced with GFRP bars (deck slabs and barriers). In
2010, by the collaboration between the University of Sherbrooke and Pultrall Inc. it has been
350

completed all required MTQ's impact tests on barriers (PL-1 and PL-2) reinforced with VROD FRP straight and bent bars. This was the last stage in the perspective of responding to
all MTQ's standards. The field applications in the last ten years have demonstrated that the VROD GFRP bars were utilized in many concrete bridge barriers a cross Canada to eliminate
the corrosion problem of steel. Figure 9 shows the cast-in-place concrete barrier with V-ROD
FRP bent bars (Hwy 400/King Rd Bridge, ON).

a. CFRP stirrups in girders


b. Finished precast girders
Fig. 7. Eagle River Bridge MTO NW region (2009)

Fig. 8. Hwy 410 overpass in Sherbrooke, QC (2010)

Fig. 9. Cast in place concrete barrier with V-ROD FRP bent bars HWY 400/King Rd, ON

351

3.2 Parking Garage


The need for sustainable structures has motivated the Public Works and Government Services
Canada (PWGSC) in the use of FRP rebar as internal reinforcement in concrete infrastructure
applications. One of the most important successful applications is using FRP rebar in
reinforcing the parking garage. An agreement between PWGSC and the University of
Sherbrooke was reached to reconstruct the interior structural slabs of the LaurierTach
parking garage (Hull, Quebec) using carbon and glass FRP rebar, see Figure 10. The design
was made according to CAN/CSA-S806-02. This project allows direct field assessment and
long-term monitoring of FRP composite bars in a structure subjected to harsh environmental
and loading conditions. In 2010, new large parking garage (area 3000 m2) in Quebec City
(QC) was designed and constructed using the V-ROD FRP rebar. Similar demonstration
projects are planned for Montral and Qubec City as collaboration between the University of
Sherbrooke and Pultrall Inc using the V-ROD FRP rebar.

Fig. 10. Laurier-Tach Parking Garage, Ottawa/Hull, ON

3.3 Other Field Applications


In fact, the FRP rebar is being as alternative reinforcement to the conventional steel for
different structures. The V-ROD FRP rebar has been used in several and different industrial
project across Canada, USA, Middle East and Australia. It has been used as reinforcement in
retaining walls, water tank, approach slabs, marine structures, reinforcing block wall system
(see Fig 11) and pre-cast prestressed slabs.

Fig. 11. V-ROD reinforcing in Durisol block wall system, (Collingwood, ON)

352

4. FRP ACCESSORIES AND CONTINUING R&D


As the use of FRP reinforcing continues to increase dramatically, the manufacturer continues
investing in R&D to improve bar properties, expand its product range and develop accessories
to compliment the bar. As certain stage construction projects require bar connectors where
there is no room for lapping reinforcing bar, it became imperative to produce a FRP to FRP
connector, similar to connectors used for steel reinforcement except designed not to damage
the fibres. A lower performance crimped on connector tested to 345 MPa is available for
certain applications where it can for example have steel bar threaded into it as shown in
Fig.12. As with steel reinforcement, FRP bars are produced to accommodate various types of
bends, stirrups and ties required for development. As there are some limitations to FRP bends,
there is the option of utilizing headed reinforcement to achieve the required development.
This may be for example in congested reinforcement areas, long lengths of bars that cannot be
produced in FRP as one piece or as a design option for barriers for example. The headed
reinforcement would typically be used on the High Modulus (60 GPa) bar where maximum
spacing can be achieved and bends eliminated. It must be noted however that heads secured on
straight reinforcing bar will not address all reinforcement requirements. This will be
dependent on the types of structures being designed for example where continuity is required
and bends will be required. For higher performance, a 316 stainless steel connector has been
developed as a male/female assembly as shown in Fig 12.

Fig 12. Over view of straight and bent V-ROD FRP bars and accessories

5. CONCLUSIONS
The observations and the outcomes from the different field applications reported in this paper
can be summarised into the following: corrosion resistance is without a doubt the main motive
and attraction to use FRP over steel. Application of FRP reinforcement in different structures
has been proved to be very successful to date. From the construction point of view it was felt
by the construction personnel that the lightweight of the FRP reinforcements were easy to
handle and place during construction. Concrete bridges and parking garage structures provide
an excellent application for the use of FRP in new construction or rehabilitation of existing
slabs.

353

6. ACKNOWLEDGMENT
The authors would like to thank and express their sincere appreciation to the research group at
Department of Civil Engineering, Faculty of Engineering, University of Sherbrooke,
Sherbrooke, QC, Canada, led by Professor Brahim Benmokrane, for providing technical data
along with numerous testing and reports on FRP reinforcement for concrete infrastructure.

7. REFERENCES
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.

13.

ACI Committee 440. 2001; 2003; 2006. Guide for the design and construction of concrete
reinforced with FRP bars. ACI 440.1R-01;03; 06, American Concrete Institute (ACI)
Farmington Hills, Mich.
ACI Committee 440. (2004). Guide test methods for fiber reinforced polymers (FRPs) for
reinforcing or strengthening concrete structures, ACI 440.3R-04, American Concrete
Institute (ACI), Farmington Hills, Mich., 40 p.
Canadian Standards Association (CSA). 2000. Canadian highway bridge design code.
CAN/CSA-S6-00, Rexdale, Toronto.
Canadian Standards Association (CSA). 2002. Design and construction of building
components with fiber reinforced polymers. CAN/CSAS806-02, Rexdale, Toronto, 177 p.
Canadian Standard Association (CSA). 2006. Canadian Highway Bridge Design Code.
CAN/CSA S6-06, Rexdale, Ontario, Canada, 788 p.
Canadian Standards Association (CSA). 2006-Edition 2010. Canadian highway bridge
design codeSection 16, updated version for public review. CAN/CSA-S6-06, Rexdale,
Toronto.
Canadian Standards Association (CSA). 2010. Specification for fibre-reinforced
polymers. CAN/CSA S807-10, Canadian Standards Association, Rexdale, Ontario,
Canada, 44 p.
Benmokrane, B., Wang, P., Ton-That, T.M., Rahman, H., and Robert, J.F. 2002.
Durability of glass fiber-reinforced polymer reinforcing bars in concrete environment.
Journal of Composites for Construction, 6(3): 143153.
Benmokrane, B., El-Salakawy, E.F., Desgagn, G., and Lackey, T. 2004. Building a New
Generation of Concrete Bridge Decks using FRP Bars. Concrete International, the ACI
Magazine, 26(8): 84-90.
Benmokrane, B., El-Salakawy, E., El-Ragaby, A., and Lackey, T. 2006. Designing and
Testing of Concrete Bridge Decks Reinforced with Glass FRP Bars. Journal of Bridge
Engineering, 11(2): 217-229.
Benmokrane, B., El-Salakawy, E., El-Ragaby, A., and El-Gamal, S. 2007. Performance
Evaluation of Innovative Concrete Bridge Deck Slabs Reinforced with Fibre- Reinforced
Polymer Bars. Canadian Journal of Civil Engineering, 34(3): 298310.
El-Salakawy, E.F. and Benmokrane, B. 2003. Design and Testing of a Highway Concrete
Bridge Deck Reinforced with Glass and Carbon FRP Bars. ACI Special Publication, Field
Applications of FRP Reinforcement: Case Studies, SP-215-2, pp. 37-54. (Editors
invitation).
El-Salakawy, E., Benmokrane, B., and Desgagn, G. 2003a. FRP composite bars for the
concrete deck slab of Wotton Bridge. Canadian Journal of Civil Engineering, 30(5): 861
870.

354

14. El-Salakawy, E., Benmokrane, B., Masmoudi, R., Brire, F., and ric Breaumier, E.
(2003b). Concrete Bridge Barriers Reinforced with Glass Fiber-Reinforced Polymer
Composite Bars. ACI Structural Journal, 100(6): 815824.
15. El-Salakawy, E.F. and Benmokrane, B. 2004. Serviceability of Concrete Bridge deck Slabs
Reinforced with FRP Composite Bars. ACI Structural Journal, 101(5): 727-736.
16. El-Salakawy, E. F., Benmokrane, B., El-Ragaby, A., and Nadeau, D. 2005. Field
investigation on the first bridge deck slab reinforced with glass FRP bars constructed in
Canada. Journal of Composites for Construction, 9(6): 470479.
17. ISIS Canada. 2001; 2006; 2007. Reinforcing concrete structures with fiber reinforced
polymers. ISIS-M03-2001; 2006; 2007, The Canadian Network of Centers of Excellence
on Intelligent Sensing for Innovative Structures, Univ. of Winnipeg, Manitoba, Canada.

355

356

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

MECHANICAL RECOVERY OF EPOXY ADHESIVES SUBSEQUENT


TO EXCEEDING GLASS TRANSITION TEMPERATURE
O. Moussa1, A. P. Vassilopoulos2, J. de Castro3 and T. Keller4
1

PhD student, Composite Construction Laboratory (CCLab), EPFL, Switzerland


Research and Teaching Associate, Composite Construction Laboratory (CCLab), EPFL, Switzerland
3
Scientist, Composite Construction Laboratory (CCLab), EPFL, Switzerland
4
Professor and Director, Composite Construction Laboratory (CCLab), EPFL, Switzerland
2

ABSTRACT
Structural adhesives are susceptible to temperatures exceeding their glass transition
temperature, Tg, where a significant reduction in mechanical properties occurs. The question
arises to which extent, subsequent to cooling down, the mechanical properties can recover and
how long it takes to re-attain a reasonable mechanical performance.
An experimental investigation on the quasi-static thermomechanical response of a cold-curing
structural epoxy adhesive subjected to a wide range of temperatures (-35 to 150C) is
presented in this paper. The adhesive exhibited a full mechanical recovery subsequent to
cooling down from temperatures above Tg. An increase of mechanical properties, compared to
those before exposure to temperature, was achieved due to post-curing effects. A semiempirical model successfully simulated the thermomechanical recovery when taking the postcuring effects into account. The effect of exposure time on the mechanical response of the
examined adhesive was also investigated within the glass transition range. A significant effect
at temperatures above Tg and less pronounced effect at temperatures below Tg has been
observed.

1. INTRODUCTION
Adhesive bonding can offer substantial economic and performance advantages over other
conventional methods of joining. Cold-curing epoxy adhesives are primarily used in the
strengthening of existing structures. CFRP strips, for example, are adhesively bonded to
concrete or steel structures to increase structural resistance. Furthermore, adhesives are
increasingly used in new construction, e.g. to connect FRP bridge deck elements among
themselves or to main girders [1-2]. For practical reasons, resins used in civil engineering are
cold-curing resins, i.e., harden at ambient temperature. These adhesives still cure at low
temperatures, but allowable design values of mechanical properties and physical
characteristics are achieved only after longer curing periods [3].

357

During this curing period, the glass transition temperature also develops slowly, particularly at
low temperatures during winter, and it is not excluded that the latter is exceeded by the
ambient temperature. In this case, the elastic modulus may drop by up to three orders of
magnitude between the glassy and rubbery states. This is due to the movement of molecules at
elevated temperatures which tend to break the secondary bonds (e.g. hydrogen bonds and Van
der Waals interactions) joining the strong covalent intramolecular bonds [4]. A general
relationship exists not only between the adhesive mechanical properties and temperature but
also the time during which a given temperature is sustained.
In order to better understand the adhesive behavior with respect to the previously mentioned
applications, the effect of different temperatures and time periods on the mechanical
performance of a structural adhesive is investigated in this paper. The thermo-mechanical
recovery response and the influence of post-curing on physical performance and tensile
properties of the adhesive are also investigated.

2. EXPERIMENTAL INVESTIGATION
2.1 Adhesive Description
A commercial epoxy adhesive Sikadur-30 (from Sika Schweiz AG) was used in this study. It
is a two-component thixotropic, solvent free epoxy based resin, which is used for structural
joints between steel and concrete components. The resin and hardener are mixed at ambient
temperature at a ratio of 3:1 by weight of the respective constituents. The adhesive contains
silica quartz fillers with size between 200 and 500 m and weight fraction of approximately
55 %, as obtained from optical microscopy and resin burn-off respectively. According to the
manufacturers data sheets, the tensile strength and modulus of elasticity of the fully cured
material are 31 MPa and 11.2 GPa respectively (according to ISO 527). According to the
manufacturer, the glass transition temperature of the fully cured material is 62C (resulting
from a torsion pendulum test).

2.2 Experimental Set-Up and Procedure


Tension specimens were fabricated in separate aluminum molds with dimensions presented in
Fig. 1a (according to ASTM D638). Afterwards, specimens were cured at lab conditions for
two weeks.
Tensile experiments were carried out using a MTS Landmark 25-kN servo-hydraulic load unit
calibrated to 20% of its load capacity. All experiments were performed in a climate chamber
(ATS 3710) connected to a liquid nitrogen cooling system. A reference specimen (for
temperature measurement) was placed close to the loaded specimen with a thermal gage (Pt100) incorporated in the middle. Wooden tabs were glued at each side of the specimen
extremities (in contact with metal wedges) in order to avoid any temperature change at the
grips position thus ensuring a uniform temperature distribution along the full length of the
specimen. Longitudinal strains were measured using a MTS clip-on extensometer of 25 mm
gage length, as shown in Fig. 1b. Experiments were performed in displacement control at a
loading rate of 5mm/min. A minimum of three specimens were tested for each of the
considered cases.
358

W = 13 mm
R = 76 mm

WO = 19 mm

G
T = 4 mm

L = 57 mm
D = 115 mm

D = 115 mm
LO = 165 mm

Fig. 1. a) Specimens dimensions according to ASTM-D638 b) Experimental set-up


The experimental program of mechanical investigation was divided into two categories. The
first includes exposure to temperatures from -35 to 60C (see Table 1). As the Tg of specimens
cured at lab temperature for two weeks was found to be around 45C, the time during which
the specimen was exposed to temperature before loading was also investigated within the glass
transition range (40-50C). The second category (Table 2) involved tension experiments
subsequent to exposing specimens to temperatures, Tpc, exceeding Tg, then cooling down to
different temperatures, Tr, and thus investigating the recovery behavior of the adhesive.
Table 1: First experimental category
T (C)
-35
t (hr)
0.5
2
4

Tpc (C)
Tr (C)
tpc (hr)
0.5

20

30

40

50

60

Table 2: Second experimental category


60
100
20

40

50

20

40

50

60

150
50

60

For physical characterization of the adhesive, a heat-flux differential scanning calorimeter


(DSC-TA Q100) connected to a thermal analyzer was used to detect the heat released during
the cure reaction. The DSC is equipped with a liquid nitrogen cooling system providing an
inert atmosphere. Samples cured at lab temperature for two weeks were then post-cured at 60,
100 and 150C for 30 minutes and then quenched in liquid nitrogen. The residual cure and the
359

corresponding glass transition temperature were then obtained by running a dynamic scan at
5C/min heating rate. Three samples were investigated at each post-curing temperature.

3. EXPERIMENTAL RESULTS AND DISCUSSION


3.1 Effect of Temperature and Time
The stress-strain responses at temperatures between -35C and 60C are shown in Fig. 2.
Material stiffness and strength are reduced from 14.10.3 GPa and 45.04.4 MPa at -35C to
0.160.01 GPa and 5.270.13 MPa at 60C respectively which emphasizes the dependence of
mechanical properties on temperature.
The effect of time during which the temperature was sustained is presented in Fig. 3. For
specimens loaded at 50C (above Tg) a dependence on the time was found. After 0.5 hour the
tensile properties of the specimens dropped compared to properties before exposure to
temperature (at lab temperature). This is due to the fact that exposure temperature is higher
than the glass transition temperature attained by specimens cured at lab temperature for two
weeks (T>Tg). By prolonging the exposure time to 2 hours, a post-curing effect took place and
thus restoring the tensile properties. By prolonging the exposure time from 2 to 4 hours, the
increase in tensile properties was less pronounced. On the other hand the exposure time had a
less pronounced effect when exposing the specimens to 40C i.e. below Tg.
50

-35C
0C
20C
30C
40C
50C
60C

Stress, (MPa)

40

30

20

10

10

Strain, (%)

Fig. 2. Typical stress-strain curves of specimens exposed to temperatures between -35 and
60C

360

20

50C_0.5h
50C_2h
50C_4h
40C_0.5h
40C_2h

Stress, (MPa)

16

12

10

Strain, (%)

Fig. 3. Typical stress-strain curves for different time periods within the glass transition range

3.2 Thermomechanical Recovery and Post-Curing Effect


Curing of cold-curing adhesives is incomplete when cured at ambient temperature and below.
A fully cured state can be only achieved by post-curing. The effect of post-curing temperature
on the curing degree, , and Tg is shown in Fig. 4. As post-curing temperature increases, more
thermal energy is supplied to the system and more curing reactions can take place, which
results in higher degrees of cross-linking within the system. The more tightly knitted
molecular network requires more energy to permit chain motion and, correspondingly, Tg
increased. Accordingly, Tg increased from 45C (samples cured at lab conditions) to 62C
when post-cured at 150C for 0.5 hour. The corresponding curing degree increased from
approximately 95 to 100%.
65

1.00

Curing degree, (-)

60

Tg (C)

55

0.98

50

0.96
45

40

Curing degree
Tg
0

25

50

75

100

125

0.94
150

Post-curing temperature, Tpc (C)

Fig. 4. Effect of post-curing temperature on curing degree and Tg


The thermomechanical recovery behavior of the adhesive is shown in Fig. 5 subsequent to
exposing the specimens to 60, 100 and 150C during 0.5 hour. Specimens showed a recovery
behavior similar to that of the original curve (sigmoidal shape). An improvement of the
stiffness as a function of post-curing temperature was noticed compared to specimens tested
with no post-curing involved (black symbols).

361

A semi-empirical model was used to simulate the thermomechanical behavior of the adhesive
[5]. This model is derived from the distribution of the relaxation time over the transition phase
together with fitting the shape of the stiffness distribution along the transition. By
implementing the change in the glass transition temperature due to post-curing (Fig. 4), the
changes in the mechanical properties could be described. The modeling curves show a good
agreement with the experimental results. At 150C, a difference between modeling and
experimental results was found. Degradation of the adhesive might have been involved at this
temperature which would explain this difference.
15

Stiffness, E (GPa)

12

0
-40

Exp._Temp
Rec_60
Rec_100
Rec_150
Gibson
-20

20

40

60

80

100

Temperature, T (C)

Fig. 5. Recovery behavior and effect of post-curing temperature on stiffness (exposure


duration of 0.5 h)

4. CONCLUSIONS
The thermomechanical recovery behavior and the influence of post-curing when the adhesive
is subjected to temperatures exceeding Tg were investigated and the following conclusions
were drawn:

The mechanical performance of structural adhesives is highly dependent on service


temperature and the time during which this temperature is sustained, particularly within
the glass transition range.
A full recovery of stiffness was found subsequent to cooling down from temperatures
above Tg.
Stiffness was increased when specimens were post-cured (above Tg) due to an increase of
the curing degree.
An existing semi-empirical model was able to describe the stiffness behavior along the
glass transition range by implementing the change in Tg due to post-cure.

5. ACKNOWLEDGEMENTS
The authors would like to thank the Federal Roads Authority (FEDRO) for the funding of this
project; SIKA AG, Zurich for the adhesive supply; the Laboratory of Composite and Polymer
Technology (LTC EPFL) for the use of the DSC equipment.
362

6. REFERENCES
1.
2.

3.
4.
5.

Mays, G.C. and Hutchinson, A.R. 1992. Adhesives in civil engineering. Cambridge
university press, 348 p.
Blaschko, M. and Zehetmaier, G. 2008. Strengthening the Rslautal Bridge using
innovative techniques, Germany. Structural Engineering International: Journal of the
International Association for Bridge and Structural Engineering (IABSE), 18(4): 346
350.
Ellis, B. 1993. Chemistry and technology of epoxy resins. Blackie academic professional,
an imprint of Chapman & Hall.
Mahieux, C.A. and Reifsnider, K.L. 2001. Property modeling across transition
temperatures in polymers: a robust stiffness-temperature model. Polymer, 42(7): 3281
3291.
Gibson, A.G., Wu, Y.-S., Evans, J.T., and Mouritz, A.P. 2006. Laminates theory analysis
of composites under load in fire. Journal of Composite Materials, 40(7): 639658.

363

364

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

USE OF SILANE ADHESION PROMOTER TO ENHANCE FRP-TOSTEEL BOND PERFORMANCE


J. Sizemore1, J. Aidoo2 , K.A. Harries3 and J. Monnell4
1

Undergraduate Research Assistant, Dept. of Civil & Env. Engrg., Rose Hulman IT, USA
Assistant Professor, Dept. of Civil & Env. Engrg., Rose Hulman IT, USA
3
Associate Professor, Dept. of Civil & Env. Engrg., University of Pittsburgh, USA
4
Research Assistant Professor, Dept. of Civil & Env. Engrg., University of Pittsburgh, USA
2

ABSTRACT
The use of bonded fibre reinforced polymer (FRP) materials for the repair of steel
structures is an area of growing interest. Like concrete repair methods, FRP-to-steel
behaviour is dominated by the interface bond. Failures are usually adhesive failures at the
steel-to-FRP interface. FRP bond to steel requires a clean and sound substrate, and practical
field application requires a relatively simple procedure. The typical application involves
abrasive (grit) blasting followed within a few hours with a primer/conditioner to ensure that
corrosion product does not form and contaminate the newly exposed steel. The primer
chosen for epoxy adhesives is often a silane-based product which also serves as an
adhesion promoter. Previously reported research, has been inconclusive as to whether the
silane primer improves (promotes) bond performance or simply enhances bond
performance by inhibiting the formation of corrosion product between the time of surface
preparation and that of CFRP application. In this study, FRP-to-steel bond is assessed using
a notched four point bending (N4PB) test. Four surface conditioning methods are
demonstrated: no conditioning, corrosion inhibiting agent and two different silane
applications, one mixed in water, the other in methanol. The results of this study
demonstrate the superior bond performance of the silane-conditioned specimens in terms of
interfacial bond. A discussion of the effect of conditioning using silane with respect to
improving durability of FRP-repaired steel structures is presented.

1. INTRODUCTION
The use of fiber reinforced polymer (FRP) composite materials has become relatively
common in infrastructure applications. Most existing applications involve concrete-FRP
composite members. Nonetheless a relatively small and innovative body of work is
developing focusing on steel-FRP composite structural systems [1]. Steel-FRP composite
systems are almost exclusively aimed at retrofit methods for the underlying steel material
or structure. Whether strengthening, repairing fractures, relieving stress to enhance fatigue
365

performance or enhancing local or member stability, steel-FRP systems leverage the unique
material properties of each material in establishing a composite member or structure. FRPto-substrate bond is observed to be the dominate limit state in many applications. Whereas
FRP-to-concrete interface performance is dominated by cohesive failure in the concrete
substrate (the weakest component of the FRP-adhesive-substrate system); FRP-to-steel
interface behaviour is dominated by adhesive behaviour at the adhesive-steel interface. As a
result, provided the adhesive bond can be maintained, the FRP can be utilized more
efficiently than in concrete applications.

1.1 Behaviour of FRP-To-Steel Bond


In addition to conventional modes of failure (flexure, tension, buckling, etc.), FRPstrengthened steel members may exhibit debonding of the FRP laminate. In considering
debonding failures, the thickness of the adhesive layer plays a significant role in the failure
mode2. Typically a thin uniform adhesive layer is desirable. Adhesive layers of reasonable
thickness (less than 2 mm thick) will exhibit relatively ductile debonding failures within the
adhesive layer. Thicker adhesive layers exhibit brittle failures along the steel-adhesive
interface. Xia and Teng [2] have shown that FRP-steel interfacial behaviour is accurately
modeled using relatively simple load-slip relationships based on adhesive tensile properties.
Indeed, for thin adhesive layers, a bilinear load-slip relationship closely approximates
observed experimental behaviour.
Stratford and Chen [3] reported that interfacial stress analysis to predict shear and throughthickness peel stress distributions for FRP-steel adhesive joints can be accurately
accomplished using low-order linear elastic stress analysis such as that recommended by
Cadei et al.[4] Interfacial stress discontinuities occur at the termination of the adhesive
layer. A variety of termination details originally used in aerospace engineering may be used
to reduce these discontinuities including spew, convex or concave fillets, tapers and reverse
tapers, stepped FRP plates and external clamps [3].

1.2 Substrate Preparation


FRP Bond to steel requires a clean and sound substrate and practical field application
requires a relatively simple procedure. The typical method for FRP application4 involves
abrasive (grit) blasting followed within a few hours with a primer/conditioner to ensure that
corrosion product does not form and contaminate the newly exposed steel. Since epoxy
adhesives are typically used, a silane-based product can serve as a primer and adhesion
promoter. However published research [5-7] is inconclusive as to whether the silane primer
improved (promoted) bond performance or simply enhanced bond performance by
inhibiting the formation of corrosion product between the time of surface preparation and
that of CFRP application. Garden [8] reports an alternative method of corrosion mitigation
of a curved I-girder completely wrapped in CFRP; in this case, silica gel packs were
successfully used to protect the prepared surface from moisture and corrosion in the time
between preparation and CFRP installation.

366

2. ORGANOSILANE ADHESION PROMOTERS


Organosilanes are commonly used as adhesion promoters between polymeric and inorganic
materials to improve initial adhesion and bond durability in moist environments.
Applications of organosilanes include use as a surface primer or as additives in epoxy and
polyurethane adhesive formulations. For applications in civil infrastructure repair, use as a
primer is most practical. The use of organosilane adhesion promoters as prebond substrate
primers has been extensively studied and reported [9,10]. Organosilanes are most effective
on inorganic surfaces such as metals, glass, and stone [11]. Organosilanes are a family of
polymers that are composed of monomers that have a central silicone atom (Si), a
hydrolyzable organofunctional group (OR) and a second organofunctional group attached
through a small alkylene linkage (R): R Si (OR)3
The OR group is typically a methoxy, ethoxy, or acetoxy functional group that reacts with
water to form an intermediary silanol group (Si-OH). The silanol groups then condense
with each other to form polymeric structures with very stable siloxane (Si-O-Si) bonds.
They can also condense with metal hydroxyl groups on an inorganic substrate to form
stable oxane bonds (Si-O-Metal). The R is customizable and is often an amino, epoxy, or
methacrylate group, which attaches to the organic resin used to bond the repair material to
the substrate [11]. The application of organosilane as a primer thus results in the formation
of chemically bonded polysiloxane films on the metal substrate. A simplified illustration of
the coupling mechanism having three methoxy (OCH3) functional groups is shown in
Figure 1.
OCH3
R

Si

OCH3
OCH3

3H20

OH
R

Si

OH
OH

OH
3 R

Si

OH
OH

OH

OH

OH

inorganic surface

O Si O Si O Si O + 3
O

inorganic surface

Fig. 1. Organosilane adhesion promoter reacts with water to form silanols, which
react with the inorganic substrate.
Silanes are generally effective promoters of epoxy and polyurethane adhesion to metal
substrates although their effectiveness varies. Petrie [11] reports a spectrum of silanemediated bond effectiveness based on substrate material shown in Table 1. The
effectiveness of silane on a steel substrate is rated as slight to good and is dependent on
the metal oxide content (i.e.: presence of OH groups on the surface to bind to). On the FRP
side of the adhesive joint, silanes may be effective for GFRP materials but have little effect
for CFRP materials. Nonetheless, in the present application, the primer is only applied to
the steel substrate.

carbon black
graphite
barium sulfate
gypsum
chalk
lead
zinc
nickel
asbestos
steel
inorganic
oxides
talc
mica
aluminosilicates
inorganics
alumina
copper
aluminum
glass
quartz
silica

Table 1. Effectiveness of silane as an adhesion promoter on various substrates

None
Slight
Good
Excellent

367

2.1 Silane Interphase


The roughness of the surface to which the silane is applied has a significant influence on its
effectiveness as an adhesion promoter. Smooth substrates are excellent for silane
attachment, while rough, discontinuous substrates show very little benefit [11]. However, a
rough surface may demonstrate superior bond characteristics if the mechanical aspect of
bond contributes more to the system performance than the chemical adhesion promoted by
the silane. The hardness of the interphase provided by the adhesion promoter may be an
additional parameter that affects the systems overall mechanical properties. For example, a
soft interphase, will reduce stress concentrations and can significantly improve fatigue
performance [12]. A rigid interphase improves stress transfer to the substrate and can
improve interfacial shear strength. With these factors in mind, Adhesion promoters are
generally considered to increase the fracture energy required to initiate a crack [11].

3. NOTCHED FOUR POINT BENDING TEST


In this study, the notched four point bending (N4PB) test method [13] is used to assess
performance of CFRP-to-steel interfacial bond. This test method, shown in Figure 2 is
specifically suited to measuring the fracture resistance of bi-material interfaces and is well
suited to rapid preparation and simple testing of multiple specimens. Provided the crack is
located between the inner loading lines, the cracked ligament (dimension 2a in Figure 2) is
subject to constant moment conditions. Consequently, the strain energy release rate should
exhibit steady-state characteristics. Based on Euler-Bernoulli beam theory and plane strain
conditions, the steady state strain energy release rate, Gss, may be calculated based on the
loading to cause crack propagation of the specimen[13] (see Appendix at end of paper).
2a = 38.1 mm

2c
P/2b2

P/2b2

crack

h1

1
2a

b2

L2

L1

R1 R2

25.4 mm

625 mm
P/2

h2

2
L

strain gage (typ.)

b1 = 102 mm

31.8 mm

b1

b2 = 152 mm

P/2

406 mm

h2 = 6.35 mm

hadhesive = 1.00 mm
P/2

202 mm

P/2

hCFRP = 1.65 mm

(a) test set-up


(b) specimen dimensions
Fig. 2. Notched four point bending test (Harries and Webb13). Material 1 is the
CFRP, material 2 is the steel substrate
The crack ligament (dimension 2a) is formed by applying a 38.1 mm wide strip of adhesive
tape to the steel substrate prior to application of the CFRP. Although the CFRP bonds to the
tape, the interface between the tape and steel has essentially no resistance and thus the
ligament is assumed to be debonded. The crack through the CFRP is cut using a utility
knife following application of the CFRP. The resulting crack gap is approximately 0.9 mm
(the width of the blade). To ensure that the CFRP is completely cut through, the cutting
operation is continued until the substrate steel is scratched by the knife blade.

368

For convenience, the N4PB test is conducted in the position inverted from that shown in
Figure 2. A fixed test apparatus permitting accurate and repeatable specimen alignment was
used in a conventional closed-loop universal test machine. Four electrical resistance strain
gages were located along the centerline of the CFRP strip (Figure 2b) to identify the instant
of debonding. While debonding is brittle and debonding over the entire constant moment
region is essentially instantaneous, the gages record strain only at discrete locations. Thus
the debonding load, determined at a significant drop in strain, must be interpreted as the
load at which debonding propagated past the gage location. To ensure accurate data, the test
is conducted in at a load rate of 45 N/sec while data was acquired at a rate of 4 Hz.

4. EXPERIMENTAL SPECIMENS
Notched four point bending specimens shown in Figure 2b were used in this program. For
valid and repeatable N4PB tests, the steel substrate must remain elastic during the test.
Appropriate specimen dimensions (b1, b2, h2 and 2c) for the CFRP strip (hCFRP is fixed)
used were determined from a previous parametric study. The steel substrate was ASTM
A36 steel having an experimentally determined yield stress of 303 MPa. The CFRP strip
used is a commercially available preformed product having an ultimate capacity of 2790
MPa and a tensile modulus of 155 GPa. The commercially available adhesive is specified
for use with the CFRP on concrete or steel substrates; its tensile strength and modulus are
reported to be 22.6 MPa and 1.2 GPa, respectively.
All specimens were identical with the exception of the priming applied to the steel prior to
the application of the adhesive. Four priming regimes were used: 1) no treatment where the
adhesive is applied directly to the prepared steel surface; 2) degreaser: a commercially
available degreasing product chosen for its reported corrosion inhibiting properties; 3)
5% silane in methanol solution; and 4) 0.5% silane in aqueous solution. Three repetitions of
each test were conducted. Each test gives two comparable results (debonding to the left and
right of the crack); thus the sample size is six for each priming regime.

4.1 Corrosion Inhibiting Degreaser


The degreaser used was a commercially available water-based biodegradable product
containing approximately 1% by volume of both Dipropylene Glycol Monomethyl Ether
(CH3(OC3H6)2OH) and Sodium Metasilicate (Na2SiO3)[14]. The corrosion inhibiting
properties derive from the alkali formulation (pH = 12.5).

4.2 Organosilane
The organosilane used in this study was -glycidoxypropyltrimethoxysilane
(CH2OCHCH2O(CH2)3Si(OCH3)3)[15]. The formulation is shown in Figure 3. This
product has three methoxy (OCH3) functional groups that will form oxane bonds to the
steel substrate (Si-O-Fe).

369

OCH3
CH2 CH CH2 O CH2 CH2 CH2 Si
O

OCH3
OCH3

Fig.3. -glycidoxypropyltrimethoxysilane.
Two methods of depositing silane on the steel surface were used. Both methods were
outlined in the product technical data sheet [15]. In the first preparation, 50 parts silane is
diluted into 950 parts methanol, resulting in a 5% silane solution. In the second preparation,
0.5% silane in an aqueous solution was prepared by stirring 5 parts silane into 995 parts
water adjusted to a pH between 3.5 and 4.5 using acetic acid. The resulting solution is
reported to be stable for 24 hours. Both solutions were applied to the steel specimens within
an hour of their preparation.

4.3 Steel Surface Preparation


The bonded surface of the steel substrate was sanded using a 1500 sfpm (surface feet per
minute) belt sander and a 40 grit zirconia alumina belt resulting in a sound, slightly striated
surface for bonding the CFRP strips. Sanding the steel surface removes mill scale and
oxidation product (rust), leaving positively charged iron (Fe2+ or 3+) ions which bond to the
negatively charged hydroxyl (OH) ions of the silanol (Si-OH) resulting in the oxane bonds
described above and shown in Figure 1. The iron bond with the oxygen from the silane may
explain the immediate appearance of brown discoloration (thought to be corrosion product)
on the steel plates following the application of the silane. This discoloration, however, was
only noted for the 0.5% silane aqueous solution and not for the 5% silane in methanol.

5. EXPERIMENTAL RESULTS
A summary of experimental results is shown in Table 2. CFRP strains at various load levels
are shown in Figure 4. At lower load levels (4.5 and 9 kN), there is little evidence of the
effect of primer application. As the loads increase (13.5 kN), greater local damage to the
adhesive layer is anticipated [12] and some differentiation becomes evident. At the
debonding limit state, the performance of each primer regime is evident; in order of
performance from poorest to best: degreaser, no treatment, 5% silane in methanol and 0.5%
silane in water.
Table 2. Summary of averagea results
no
5% silane in
degreaser
treatment
methanol
kN
15.6 (15%)
14.7 (13%)
15.2 (16%)

0.5% silane
in water
19.2 (8%)

load at debonding
CFRP
strain
at

1018 (45%)
785 (18%)
1198 (30%) 1416 (39%)
debonding
steady state strain
energy release rate13, J/mm 1.13 (32%)
0.92 (34%)
0.97 (32%)
1.40 (11%)
Gss
a
average (COV in %) of data from both sides of crack from three specimens (i.e.: n = 6)

370

Figure 5 shows the three performance criteria normalized by the performance of the no
treatment regime. The poor performance, approximately 84% of that of the no treatment
regime, of the degreaser is evident. Similarly the improved performance, approximately
130%, of the 0.5% silane in an aqueous solution is clear. The 5% silane in methanol
solution resulted in higher debonding strains (118%) although was not as resistant to
debonding (86%).
1.6
load at debonding

2500

CFRP strain ()

2000

CFRP strain at 4.5 kN applied load


CFRP strain at 9.0 kN applied load
CFRP strain at 13.5 kN applied load
strain at debonding (see Table 2)

1.4

CFRP debonding strain


steady state strain energy release rate

1.2
1.0

1500

0.8

debond

0.6

1000
0.4

13.5 kN
9.0 kN
500
no treatment
0

0.2

4.5 kN

degreaser

5% silane
in methanol

0.0

0.5% silane
in water

Fig. 4. CFRP strains at different load levels.

no treatment

degreaser

5% silane in
methanol

0.5% silane in
water

Fig.5. Effect of treatments normalized to


behavior of specimens having no treatment.

6. DISCUSSION AND CONCLUSIONS


The use of a commercial degreaser, despite its corrosion inhibiting properties, resulted in
poorer bond performance than providing no treatment at all. This is thought to result from
the deleterious effects of the deposition of sodium metasilicate on the steel surface. This
result should be noted by researchers who regular report degreasing steel substrates prior
to FRP installation.
The use of 0.5% silane in aqueous solution improves bond performance by about 20%,
while 5% silane in methanol has little (or a slightly deleterious) effect on bond
performance. The aqueous solution provides the water necessary to hydrolise the silane
into its intermediary silanol form permitting the immediate formation of the polysiloxane
film. The methanol carrier, on the other hand, does not result in hydrolization and perhaps
allows the silane to evaporate with the methanol between the time of priming and FRP
application (about 22 hours in this case).
This study has indicated the efficacy of silane in promoting adhesion of FRP-to-structural
steel surfaces. The 0.5% silane aqueous solution is believed to be very practical to use in
field applications: it is stable for 24 hours, may be easily sprayed or painted onto large
surfaces and is more environmentally benign, having only 5 parts per 1000 organosilane.

371

7. REFERENCES
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.

13.
14.
15.

Harries, K.A. and El-Tawil, S. 200. Review of Steel-FRP Composite Structural


Systems. Proceedings of the 5th International Conference on Composite Construction,
July, Colorado.
Xia, S.H. and Teng, J.G. 2005. Behaviour of FRP-to-Steel Bonded Joints. Proceedings
of the International Symposium on Bond Behaviour of FRP in Structures, December,
Hong Kong.
Stratford, T.J. and Chen, J.F. 2005. Designing for Tapers and Defects in FRPStrengthened Metallic Structures. Proceedings of the International Symposium on
Bond Behaviour of FRP in Structures, December, Hong Kong.
Cadei, J.M.C., Stratford, T.J., Hollaway, L.C. and Duckett, W.G. 2004. Strengthening
Metallic Structures Using Externally Bonded Fibre-Reinforced Polymers. CIRIA Pub.
No. C595, 233 p.
Bourban, P.E., McKnight, S.H., Shulley, S.B., Karbhari, V.M. and Gillespie, J.W.
1994. Durability of Steel/Composites Bonds for Rehabilitation of Structural
Components. Proceedings of the 3rd Materials Engineering Conference, pp. 295-303.
McKnight, S.H., Bourban, P.E., Gillespie, J.W. and Karbhari, V.M. 1994. Surface
Preparation of Steel for Adhesive Bonding in Rehabilitation Applications. Proceedings
of the 3rd Materials Engineering Conference, pp. 1148-1155.
Karbhari, V.M. and Shulley, S.B. 1995. Use of Composites for Rehabilitation of Steel
Structures Determination of Bond Durability. Journal of Materials for Civil
Engineering, ASCE, 7(4): 239-245.
Garden, H.N. 2001. Use of Composites in Civil Engineering Infrastructure. Reinforced
Plastics, 45(7/8): 44-50.
Plueddemann, E.P. 1991. Silane Coupling Agents, Plenum Press, New York, 272 p.
Mittal, K. L., Silane and other Coupling Agents, Conference Series: Vol. 1 (1992);
Vol.2 (2000); Vol. 3 (2004); Vol. 4 (2007); Vol. 5 (2009) VSP, Utrecht.
Petrie, E.M. 2005. Organosilanes as Adhesion Promoting Additives for Epoxies and
Polyurethanes. SpecialChem4Adhesives, 3 August, (accessed July 31, 2010).
ONeill, A., Harries, K.A. and Minnaugh, P. 2007. Fatigue Behavior of Adhesive
Systems used for Externally-bonded FRP Applications Proceedings of the 3rd
International Conference on Durability & Field Applications of FRP for Construction
(CDCC 2007) Quebec City, eds: Benmokrane, B. & El-Salakawy, E., May.
Harries, K.A., and Webb, P. 2009. Experimental Assessment of Bonded FRP-to-Steel
Interfaces Proceedings of ICE: Structures and Buildings, 162(5) 233-240.
LPS Precision Clean. 2008. Technical Data Sheet, LPS Laboratories, June.
XIAMETER OFS-6040 Silane. 2009. Technical Data Sheet, Dow Corning, January.

372

APPENDIX Calculation of Steady State Strain Energy Release Rate for N4PB Test
[13]
Steady state strain energy release rate per unit width of the specimen is calculated as:
M 2 (1 2
Gss
2 E2

) 1

I2 Ic

(1)

M = PL/2b2 is the applied moment per unit width calculated from the total applied load, P;
E2 and 2 are the Youngs modulus and Poissons ratio for steel (material 2);
I2 = h23/12 is the moment of inertia per unit width of the steel.
Ic is the transformed composite moment of inertia per unit width of the CFRP and steel:
3

Ic

h1
h
h h (h h2 )2
2 1 2 1
12
12
4(h1 h2 )

(2)

E2 (1 12 ) E1 (1 2 2 ) is the modular ratio.


For the FRP-to-steel system considered, it is also necessary to transform the CFRP and
adhesive layers into a single component of the bimaterial material [13] (i.e.: material 1). In
this case the thickness h1 is calculated: h1 = hCFRP(I1/ICFRP)0.33. Where I1 is calculated using
Eq. (2) replacing h1 with hCFRP and h2 with hadhesive. Similarly, the modular ratio, , is
calculated replacing E1 with ECFRP and E2 with Eadhesive, and 1 and 2 with CFRP and
adhesive, respectively.

373

374

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

STRUCTURAL HEALTH MONITORING OF CORRODED STEEL


REINFORCEMENT IN FRP REPAIRED BEAMS USING
MULTIPLPEXED FIBRE BRAGG GRATING SENSORS
R. Al-Hammoud1, K. Soudki2 and T. Topper3
1

PhD Candidate, University of Waterloo, Waterloo, Ontario, Canada


Professor and Canada Research Chair, University of Waterloo, Waterloo, Ontario, Canada
3
Professor Emeritus, University of Waterloo, Waterloo, Ontario, Canada
2

ABSTRACT
Fibre optic sensors (FOS) are considered to be state of the art technology in structural
health monitoring. FOS are widely accepted due to their facility for monitoring strains in
structures that are subjected to a harsh environment, their ability to be installed internally,
and their adaptability to multiplexing (multiplexed Fibre Bragg Grating sensors FBG).
In this study, eight reinforced concrete beams were subjected to a corrosive environment in
the laboratory. The steel reinforcement in these beams was instrumented with FBG sensors
and encapsulated strain gauges. Each beam had two instrumented bars. Both of the
reinforcing bars had an axial gun-drilled hole. One of the bars had an FBG sensor installed
in the hole and the other had encapsulated strain gauges installed before the concrete was
cast. The beams were exposed to different corrosion levels using impressed current in a
humid environment and then wrapped with carbon fibre reinforced polymers (CFRP)
before being tested under repeated loading. The main purpose of the study was to examine
the effect of CFRP repair on the bond behaviour between corroded steel and concrete under
repeated loading. This paper discusses the experience of installing FBG sensors in gundrilled bars and the handling of the FBG sensors in the laboratory and describes the
performance of the FBG sensors in beams subjected to a corrosive environment. The paper
gives recommendations concerning the use of FBG sensors for monitoring strain in
corroded steel reinforcement in concrete.

1. INTORDUCTION
The use of new advanced materials such as carbon fibre reinforced polymers (CFRP) has
been increasing in reinforced concrete structures through the past decades to solve
corrosion problems. The use of such new materials in structures requires the use of nondestructive evaluation (NDE) techniques that help monitor the structures real-time
behaviour. Monitoring the structural behaviour, helps to detect any decrease in
performance, and to plan maintenance if needed [1, 2].
375

Fibre Bragg grating (FBG) sensors have been used recently for NDE. Their main
advantages are their lightweight, small physical dimensions (which helps minimize their
effect on the structural resistance), that they are water and corrosion resistant, and their
ability to be multiplexed (measurements can be done at more than one point on the
specimen with one lead). Multiplexing helps minimize the number of wires required to get
the signals from transducers to a reading device which in turn reduces the effect the
presence of those wires has on the strength of a structure. Another important advantage of
the FBG sensors is that measurements are made using a change in wavelength, which is an
absolute measurement. Hence FBG sensors measure the strain by calibrating the shift in the
wavelength that the sensor undergoes due to a change in strain of the attached
reinforcement [3-7].
The major disadvantage of the FBG sensors is that they are brittle and hence care is needed
while installing them and handling their wires. In addition, the FBG sensors and their
required reading equipment are expensive, however due to the progress that is taking place
in this field the sensor and the reading instrument are becoming more affordable although
still expensive [8].
The main purpose of this study was to examine the effect of wrapping corroded reinforced
concrete beams with CFRP on the bond behaviour between the corroded steel
reinforcement and the concrete under repeated loading. Measurements of strains along the
steel-concrete interface were required to monitor the process of bond failure. In an earlier
study, strain gauges installed in a premanufactured groove in the bars before casting failed
in the corroded beams. This paper discusses the use and handling of multiplexed FBG
sensors in the laboratory and describes a technique in which strain can be measured in
corroded beams. The strains measured using the FBG sensors were compared to strain
readings from encapsulated strain gauges.

2. EXPERIMENTAL PROGRAM
The experimental program consisted of testing eight corroded reinforced concrete beams
repaired with CFRP sheets under repeated loading. The test variables were the anchorage
length (200 mm or 350 mm), the wrapping condition, the corrosion level (5% or 15% mass
loss) and the maximum fatigue load level since the minimum fatigue load level was fixed at
10kN for all test beams (Table 1).
All beams were of the same size with a rectangular cross-section (254152 mm) and a
length of 2000 mm. Each beam was reinforced with two 20M (19.5mm in diameter) Grade
400 steel deformed bars as tension reinforcement. Two 8 mm smooth bars were used as
compression steel reinforcement and 8 mm stirrups were used as shear reinforcement. A
hollow 8 mm diameter stainless steel bar, Grade 304, was placed at a distance of 80 mm
from the bottom of the beam (Figure 1). This steel bar was used as a cathode in the
accelerated corrosion process. The stainless steel bar and the two 20M deformed bars were
extended 150 mm from one end of the beam to provide for the electrical connections
necessary for accelerated corrosion. The 20M deformed bars were hooked from the other
end to ensure that bond failure occurs at the other instrumented end of the beam (anchorage
zone which varies between 200 mm and 350 mm). Salted concrete was placed in the
376

anchorage zone section of the beam to help initiate corrosion and an electric current
equivalent to 150A/cm2 was induced to speed up the corrosion process of the reinforcing
steel bars [9]. The 20 M deformed bars were gun-drilled with a hole diameter of 10 mm to a
depth of 550 mm. The bars were then cleaned from the inside with acetone and air pressure.
Table 9. Test Matrix
Anchorage FRP
Corrosion
length
Wrapped Level (%)
(mm)
200
No
5
200
Yes
5
200
Yes
5
200
Yes
5
200
Yes
15
350
Yes
5
350
Yes
5
350
Yes
5

Beam notation
F-200-U-5%-54kN
F-200-W-5%-82kN
F-200-W-5%-90kN
F-200-W-5%-100kN
F-200-W-15%-98kN
F-350-W-5%-136kN
F-350-W-5%-145kN
F-350-W-5%-150kN

Maximum
fatigue load
level (kN)
54
82
90
100
98
136
145
150

Failure
mode
Bond
Flexure
Flexure
Bond
Bond
Flexure
Flexure
Flexure

600

P
50 75

125

125

125

200

P
200

200

200

200

125

125

125

75 50

152

2 No. 8
Stainless
steel bar
254
45

150

45 62
175

stainless
steel bar

2000

8mm stirrups

concrete pocket

2 No. 20

175

Fig. 1. Longitudinal and cross-sectional details of the beams (All dimensions are in mm).
Fibre optic sensors were installed in half of these bars and encapsulated strain gauges
(figure 2) were installed in the other half. Fibre optic sensors and encapsulated strain
gauges were used to measure the strains along the length of the bar in the corroded repaired
beams. Due to corrosion, the bar surface was expected to damage the adhesion of the strain
gauges placed on the surface of a bar. Accordingly, the strain gauges and fibre optic sensors
were placed inside the gun-drilled bars. Sikadur 52 was used to attach the fibre optic
sensors and the encapsulated strain gauges to the inside of the rebar. The fibre optic sensors
used were multiplexed fibre Bragg grating sensors and had the following characteristics:
custom FBG sensor array, FBG center wavelengths: 1520nm 1540nm 1560nm
1580nm (+/- 0.5 nm tolerance) 5000 Rated FBG, >50%R, <3nm BW, >15dB isolation, 1
each fusion-spliced 8m jumper (with heavy duty 3mm type buffer cladding), 1 each
FC/APC.
To place the FBG sensors, a small weight was introduced at the end of the lead wire to
make sure that the FBG sensors reached the bottom of the hole in the rebar. This confirms
that the strain readings of the arrays will be taken at the points specified along the length of
377

the bar. The encapsulated strain gauges were placed as follows. To ensure the position of
the strain gauge along the length of the bar, the wires of three encapsulated strain gauges
just close to the strain gauge were glued to a 1.5 mm diameter steel rod. The steel rod was
then placed in the gun-drilled hole inside the bar. The holes for both the FBG sensors and
the encapsulated strain gauges were then filled with epoxy. The spacing between the arrays
of the FBG sensors was 50 mm and 70 mm for the beams with anchorage zones 200 mm
and 350 mm respectively (Figure 3). The spacing between the encapsulated strain gauges
was about 80 mm and 150 mm for the beams with anchorage zones 200 and 350 mm,
respectively.
Initially, the plan was to test 16 beams with FBG sensors. While placing the FBG sensors in
the gun-drilled holes, two sensors broke, and during handling the beams, another 6 sensors
were damaged due to breaking of the output wire. Accordingly, only 8 beams had working
FBG sensors and are reported in this study.

Fig. 2. Encapsulated strain gauge.

Fig. 3. Details of array placement in Multiplexed Fibre Bragg grating sensors [10]
Linear variable differential transducers (LVDTs) with an accuracy of 0.01 mm (0.0004 in.)
were used to measure the slip between the steel bar and the concrete as well as the beam
deflection at midspan. The readings of the strain gauges, the LVDTs and the load cell were
recorded using a National Instrument SCXI data acquisition system at a sampling rate of 20
readings per second and saved in a computer. The readings from the multiplexed FBG
sensors were recorded using a sm130 Optical Sensing Interrogator module through
ENLIGHT, Micron Optics Sensing Analysis software in another computer. The data
from the two computers were then synchronized with the help of a fixed time stamp
recorded on both computers.

378

3. EXPERIMENTAL RESULTS
One beam was corroded unrepaired, and the other 7 beams were repaired with CFRP sheets
along the anchorage zone. The failure mode for the unrepaired beam was by slip of the bar
followed by spalling of the bottom part of the concrete. The repaired beams failed by either
FRP rupture after the slip of the bars or by bar rupture or by concrete crushing (Beam F350-W-5%-150 kN). The gun-drilled hole had a cone end that introduced a stress raiser in
the bar causing the bar to crack and eventually rupture due to the applied fatigue loading
due to a crack that propagated from the inside outwards. The corrosion pits in the bar along
the anchorage zone had similar effect by introducing a stress raiser leading to rupture of the
bar due to fatigue loading. As a result five beams failed in flexural-fatigue and three beams
failed in bond-fatigue. Figure 4-a shows a typical flexural fatigue failure. Figure 4-b shows
a typical fatigue-bond failure with a FRP rupture.

(a)

(b)

Fig. 4. Beams turned upside down after failure: a) beam F-200-W-5%-90kN showing bar
rupture due to corrosion pits, b) beam F-200-W-15%-98kN showing FRP rupture after bar
slip.
Figure 5 shows the fatigue life versus the load range applied data for all beams in this
study. The 3 beams with an anchorage length of 350 mm failed in flexural-fatigue. Up to a
maximum fatigue load of 145 kN (load range = 135 kN) the beams failed by bar rupture. In
an attempt to achieve a fatigue-bond failure, the maximum load was increased to 150kN
(load range = 140 kN) which caused the reinforcing bars to yield leading to failure of the
beam by concrete crushing (Beam F-350-W-5%-150 kN). It was concluded that wrapping
the beams with a 350 mm anchorage length (with 5% corrosion) will prevent fatigue-bond
failure in such beams.
For the beams with an anchorage length of 200 mm, corroding a beam to a 5% corrosion
level caused the beam to fail at 254,700 cycles in bond-fatigue at a load range of 44 kN.
After being repaired with CFRP sheets the 5% corroded beams failed by flexural-fatigue up
to a load range of 80 kN. The load range was increased up to 90 kN resulting in a fatiguebond failure of the beam at a life of 63,961 cycles. One beam was then corroded to 15%
mass loss and repaired with a CFRP sheet. This beam was tested at a load range of 88 kN
resulting in a bond failure at a fatigue life of 70,553 cycles (Figure 5). It is clear that
repairing the beams with CFRP sheets increased the fatigue life of the beams. Exposing the
379

beams to further corrosion and then repairing them did not result in a further reduction in
the fatigue lives of the beams. This is due to the fact that corrosion causes the concrete-bar
interface to crack which leads to spalling of the bottom part of the concrete. Therefore,
repairing the beams with CFRP sheets confined the concrete around the reinforcing bars
increasing the fatigue-bond life, until the failure was by rupture of the FRP sheet.
Irrespective of the corrosion level, the fatigue-bond life of a beam is dependent on the
strength of the FRP sheet.

Fig. 4. Load Range (kN) versus Fatigue Life (Cycles)


The strains along the length of the bar were plotted for different percentages of the fatigue
life. Figure 6a shows the strain results from the encapsulated strain gauges while Figure 6b
shows the strain results from the FBG sensors for the unrepaired beam. The strain values
show an approximately linear behaviour along the length of the bar which suggests a
uniform shear stress along the length of the bar in the anchorage zone. The strain values
increase as slip starts to occur (Figure 7). From figure 7, the increase in slip from 0 to 0.5
mm from the first cycle to 10% of the fatigue life explains the increase seen in the strain
readings (Figures 6).
The readings from the FBG sensors were in close agreement with the readings from the
encapsulated strain gauges for the first cycle before slip occurs. The difference in the strain
readings between the strain gauges and the FBG sensors increased when slip varied from
one bar and the other. At failure slip was higher in the bar that had the FBG sensors
installed (Figure 7) than the other bar as load was transferred to the bar with the
encapsulated strain gauges. This resulted in an increase in the strains decreasing in the
former and increasing in the latter. It is important to note that the FBG sensors were not
affected by corrosion and they continued to work even after the failure of the beam. The
corrosion levels of the beams varied between a 5% and a 15% mass loss. Protecting the
sensors from corrosion by installing them inside the bars that were corroded allowed the
380

recording of strain readings along the length of the bar for both sensor types. The variation
of strain along the length of the bar was essential in determining the shear stress at the
steel-concrete interface which was used to evaluate the bond behaviour. This new technique
of installing the sensors inside the reinforcing bars eliminated the transducer failure due to
corrosion that had occurred in previous tests.

(a)
(b)
Fig. 6. Strain Profile along beam length: a) encapsulated strain gauges' readings, b) FBG
sensors readings.

Fig. 7. Slip versus percentage of fatigue life


4. CONCLUSIONS
FBG sensors and encapsulated strain gauges placed inside gun-drilled steel bars were not
affected by a corrosive environment. The technique of gun-drilling the bars enabled the
strains along the length of corroded bars to be recorded under fatigue. The strains along the
length of the bar are critical in calculating the shear stress at the steel-concrete interface
which helps in understanding the bond behaviour under repeated loading. The advantages
381

of the FBG sensors are their light weight, small physical dimensions, water and corrosion
resistance and their ability to be multiplexed, which reduces the amount of wires required.
These advantages together with installing them in gun-drilled holes make them suitable for
monitoring the strain in reinforced concrete subjected to harsh environments. However,
great care needs to be taken while installing and handling the FBG sensors to avoid
damaging them and their wires.

5. REFERENCES
1.

Grattan, S., Basheer, P., Taylor, S., Zhao, W., Sun, T. and Grattan, K. 2007. Fibre
Bragg grating Sensors for Reinforcement Corrosion Monitoring in Civil Engineering
Structures. Journal of Physics: Conference series 76.
2. Benmokrane, B.; Rahman, A., Mukhopadhyaya, P.; masmoudi, R.; Zhang, B.; lord, I.
and Tadros, G., 2001. Fiber-Optic Sensors Monitor FRP-Reinforced Bridge. Concrete
International, 23, (6): 33 38.
3. Majumder, M.; Gangopadhyay, T.; Chakraborty, A.; Dasgupta, K. and Bhattacharya,
D. 2008. Fibre Bragg Gratings in Structural Health Monitoring Present Status and
Applications, Sensors and Actuators, A 147, pp. 150 164.
4. Grattan, S., Basheer, P., Taylor, S., Zhao, W., Sun, T. and Grattan, K., 2007. Corrosion
Induced Strain Monitoring Through Fibre Optic Sensors. Journal of Physics:
Conference series 85.
5. Tahir, B.; Ali, J. and Abdul Rahman, R., 2005. Strain Measurements Using Fibre
Bragg Grating Sensor. American Journal of Applied Science, (special issue), pp. 40
48.
6. Tjin, S., Wang, Y., Sun, X., Moyo, P. and Brownjohn, J., 2002. Application of QuasiDistributed Fibre Bragg Grating Sensors in Reinforced Concrete Structures.
Measurement Science and Technology, 13: 583 589.
7. Merzbacher, C.; Kersey, A. and Friebele, E., 1995. Fiber Optic Sensors in Concrete
Structures: A Review. Smart Material Structures, 5: 196 208.
8. Degrieck, J.; De Waele, W. and Verleysen, P., 2001. Monitoring of Fobre Reinforced
Composites with Embeded Optical Fibre Bragg Sensors, with Application to Filament
Wound Pressure Vessels. NDT & E International, 34: 289 296.
9. Al-Hammoud, R.; Soudki, K. and Topper, T., 2010. Bond Analysis of Corroded
Reinforced Concrete Beams Under Monotonic and Fatigue Loads. Cement and
Concrete Composites, 32(3): 194 203.
10. Micron Optics, Inc., 2007. Personnel Conversation, Atlanta, GA, USA.

382

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

SHORT TERM AND LONG TERM PROPERTIES OF NEWLY


DEVELOPED BENT GFRP REINFORCING BARS
A. Weber1 and C. Witt
1
2

Head of R & D, Division of GFR, Schoeck Bauteile GmbH, Baden-Baden, Germany


President, Schoeck-Canada Inc., Kitchener, ON

ABSTRACT
Bent FRP rebars generally are made in a manufacturing process using the same material
composition, wetting and consolidation process as that of straight bars. An interconnected
bending process and a modified curing process are added. All deformations have to be
finished before curing.
The paper presents in a short analysis the general state of stress for the bent section of a
rebar under load. Solutions are presented to reduce local stress concentration and to
increase the overall performance. The result is a bent bar with a modulus of 50GPa with
nearly no restrictions on the bending geometry.
A test approach to evaluate short term and long term material properties is presented. The
test setup used therefore is presented and discussed. The basic properties short term
strength and modulus of elasticity are given as well as the bond properties of the straight
portion and the bent portion of the bar. Safe long term values are evaluated from long term
tests under sustained load.
1. INTRODUCTION
The lack of the possibility of bending a straight bar is seen as one of the main drawbacks of
FRP rebars. But in most construction projects even steel stirrups are delivered pre-bent to
the construction site. But what is the difference between bent steel and bent FRP
reinforcing bars?
2. ANALYSIS
Bending of a steel rebar, for example with a typical bending diameter of 5 times the
diameter of the bar, leads to deformations of more than 15% on both sides (compression
and tensile section) of the bar. This high value can not be achieved by composite materials.
In addition, this deformation is achieved in steel by (non elastic) yielding.

383

If this bending is not possible for any cured (elastic) composite, the bending process has to
be completed before curing. Can the fibres alone manage this high strain, or do they slip to
compensate this high local strain?
The answer is given by several researchers (Ishihara et al. 1997; Morphy 1999). The
analysis of them showed that there is no compensation and no high elastic strain in the
fibres. If the fibres are not held together in a parallel way, the behaviour will be as follows:
as the outer fibres are not being stretched while bending, the whole deformation takes place
by compression of the fibres at the inside section of the bend. For a bending diameter of 7
times the diameter the compression strain is now in the order of 20%. Here buckling is the
only possibility for the fibres to achieve this deformation as shown in Figure 1.

20% compression strain

Fig. 1. Length of a sinus curve in real scale


Assuming such a bend of 7 times the bar diameter and a sinusidal buckling form with 20%
compression strain, the amplitude is 15% of the wavelength. Meaning if the fibre rovings
buckle every 10 mm the Amplitude is 1.5 mm, off axis angles of more than 30 are
achieved. Such laminates are much weaker and less stiff than parallel unidirectional
laminates.

Fig. 2. Cuts of a stirrup bigger (left) and smaller (right) bending diameter

384

In Figure 2 details of two different stirrups are shown. Both with the same diameter and
fibre volume fraction. The first on the left side with a bend radius of 7ds; where ds is the bar
diameter, and the second on the right side with a bend radius of 4ds. On the left side, the
fibres are more or less parallel through compression of the envelope. On the right side
bigger sections with buckled fibres are to be seen.
From Figure 3, it can be noticed that smaller off axis angles lead to higher strengths. A
good compromise between mechanical behaviour and applicability is a bending diameter of
7 times the bar diameter (Bank 2006).
1
Strength 0.6

strength as percentage of 0

0,9

Strength 0.4

0,8

Strength 0.2

0,7
0,6
0,5
0,4
0,3
0,2
0,1
0
0

10

20

30

40

50

60

deviation angle

Fig. 3. Influence of angle misalignment on strength of UD laminates for different fibre


volume percentages (Menges 1980)
3. MANUFACTURING
If consolidating and curing is performed in two different process steps, a third step in which
the bar is deformed can be included. In the case of a closed mould pultrusion this is not
possible because both steps are performed in the closed heated tool. In addition, the
grinding of ribs which is a simple well automated and well controllable process for straight
bars, seems to be extremely complex for curved geometries. As a consequence a totally
different process had to be developed.
The main thoughts in this new process are:
- hold all fibres at their position while bending
- achieve a bond behaviour which is comparable to that of the straight bar
- use simple tools and machineries
The solution is the close envelopment of the fibres using a corrugated pipe. This pipe is
flexible in the bending direction but stiff in the diameter direction.

385

4. TEST METHOD
As these bars are different in manufacturing, all basic properties have to be determined. The
most important values are: Strength and stiffness of the bend and the straight portion of the
bar as well as long term properties.
In the codes different setups have been recommended to test the properties of bent bars. In
Ahmed et al. (2010) the B.5 (ACI 440.3R-04) setup (which is virtually the E setup of CSA
S806-02) has been found to be superior to the simple B12 setup. The authors agree that the
test setup should represent the reality as closely as possible and recommend the following
setups as a quarter setup B5 Annex E (Figure 4).

Fig. 4. Quarter concrete block setup similar to setup CSA S806-02 Annex E
Six additional tests according the full Annex E setup B have been performed at the RWTH
Aachen University with similar results to the smaller test setup.
The longterm strength is determined according the time to failure approach with a specimen
similar to the setup top right strength test, bent bar.
4.1 TEST PARAMETER
For the comparison of the bond behaviour two different series have been produced. Both
with exactly the same fibre and resin content as well as bending diameter. For the
Parameter A the bond was normal, for parameter B the bond was reduced by purging the
resin between the core of the bar and the ribs.
386

5. RESULTS
The results from the tests at the straight portion are mostly related to the fibre content of the
bar. The calculated Modulus (fibre percentage by modulus of the fibre) is very close to the
measured value. 50 GPa is the measured value from different sources. The measured value
of the strength is in the range of 1000 MPa. For both parameters A and B more or less the
same values have been reached.
5.1 Comparison Bent Portion A and B
Figure 5 shows the normal failure of a specimen Type A (normal bond). It has been taken
of a specimen after a failure under a constant load of 500 MPa after 20h of load at 60C.
The section is broken with no signs of bending or interlaminar shear. The bond breaker is
on the left side.
Strength
(MPa)
727,0
812,0
770,0
672,6
690,3
816,0
796,0

Force (N)
164302
183512
169500
76000
78000
92208
89948

Test
CSA E
CSA E
CSA E
CSA E/4
CSA E/4
CSA E/4
CSA E/4

AC
AC
AC
UW
UW
SB
SB

Fig. 5. Tensile failure of a bent bar


The specimen in Figure 6 is a specimen without bond (Type B). Figure 6. shows a
specimen of type B broken directly after loading with 240 MPa. Delaminations and bending
failure clearly can be seen.
Force (N)
54300
45100
102600

Strength
Test
(MPa)
240,3 CSA E
199,6 CSA E
454,0 CSA E

Fig. 6. Bending failure of a bent bar (no bond)

387

AC
AC
AC

5.2 Longterm Values


The longterm strength is determined according the time to failure approach at 60C. With a
shift to 40C by a factor of 100 this method is on the safe side. Calculations have shown
that in Canada and Europe the temperature only in thin insulated faades can reach values
in a range of up to 38C (Weber 2010).
Time to Failure Tests in Wet Concrete
Temperature Shift ComBAR Stirrup 12mm + 20mm

sustained load in MPa

1000
900
800

,9

Extrapolation 12 and 20mm

700
600
500
400

1000
900
800

,0

700

,8

,7

,6

,5

600
40C

60C

500
100a

350

350

50a

300

300
250
200

400

,4

,3

250

300
1

10

100

1000

10000

100000

1000000

time to failure in h

Fig. 7. Time to failure results stirrups at 60C

The characteristic/guaranteed values of the Type A specimens are shown in Table 1.


Table 1. Table with characteristic/guaranteed values (parameter A)
ComBAR bent bars 2011

d=12

d=20 source
>900 Schck, Uni Waterloo
Schck, Uni Waterloo, RWTH
>550 AC
Schck, Uni Waterloo, RWTH
50 AC

Strength straight portion

MPa

1000

Strength bent portion

MPa

650

Modulus Elasticity

GPa

50

Strength long term

MPa

250

bond straight portion

MPa

bond bent portion

MPa

10

250 Schck
10 Schck, Uni Waterloo
Uni
12 Waterloo

388

6. SUMMARY AND DISCUSSION


The presented test results show that with new methods it is possible to achieve high
strength and high stiffness not only in pultruded straight glass fibre rebars, but also in bend
profiles using a new bending technique.
The presented stirrup is a significant step forward compared to conventional stirrups made
like straight bars. Stirrups with short term strength values of more than 700 MPa are now
state of the art.
It has been shown, that there are many possibilities to increase the strength of a unidirectional composite in a bend.
The first point is the parallelity of the fibres: Ondulated fibres reduce the strength beginning
at angles as small as 5. For laminates with an off-axis value of 30 a strength of only 10%
is reached. A solution for this problem is a greater bending angle in the range of 7times the
bar diameter and a rigid envelope, which holds the fibres at their place.
The second important point is the bond in the bent portion. A slip in this portion leads to a
bending moment in the beginning of the curvature. This can have a knock-down effect on
the already reduced strength of this section. For a smooth section without bond only one
third of the strength of a section with good bond was measured.
The third point is the durability: In short term tests a part of the load will be transduced by
ondulated fibres through the resin. In long term the creep of the resin leads to a
redistribution towards the straight fibres. The load for the fibres will increase and the time
to failure will be much shorter than for a uniformly loaded section.
A bend can be tested according the time to failure concept giving us secure design values.
The extrapolated value for 10,000 hours at 60C is a good estimation for the 1 mio. h value
for 40C. In the case of the tested stirrups, a value of 250 MPa is determined as the
characteristic value of the long term strength.
This is sufficient but only slightly more than the limit value for shear reinforcement
according to the ACI 440, S806-02 or the CHBDC S6-06..
With this in mind it should be checked if first generation stirrups, according the existing
codes, can sustain the prescribed forces in the critical sections also in the longterm.
7. REFERENCES
1.
2.

Ahmed, E. El-Sayed A, El-Salakawy E. and Benmokrane B. 2010. Bend Strength of


FRP Stirrups: Comparison and Evaluation of Testing Methods. Journal of Composites
for Construction, ASCE, 14(1): 3-10.
ACI Committee 440. 2004. Guide test methods for fiber reinforced polymers FRPs for
reinforcing or strengthening concrete structures. ACI 440.3R-04, American Concrete
Institute (ACI), Farmington Hill, MI, 40 p.
389

3.

ACI Committee 440. 2006. Guide for the design and construction of concrete
reinforced with FRP bars. ACI 440.1R-06, American Concrete Institute (ACI),
Farmington Hill, MI, 44 p.
4. Bank, L. 2006. Composites for construction Structural Design with FRP Materials.
Chapter 6 FRP Shear Reinforcement, Wiley, ISBN: 978-0-471-68126-7, 560 p.
5. Canadian Standards Association (CSA). 2002. Design and construction of building
components with fibre reinforced polymers. CAN/CSA S806-02, Mississauga, ON,
177 p.
6. Canadian Standards Association (CSA). 2006. Canadian Highway Bridge Design
Code. CAN/CSA S6-06, Mississauga, ON.
7. Ishihara, K., Obara, T., Sato, Y., Ueda, T., and Kakuta, Y. 1997. Evaluation of
ultimate strength of FRP rods at bent-up portion. Proc., 3rd Int. Symp. on Nonmetallic
(FRP) Reinforcement for Concrete Structures, Vol. 2, Japan Concrete Institute JCI,
Sapporo, Japan, 2734.
8. Menges, G. Kunststoffverarbeitung Umdruck zur Vorlesung (printed lecture notes ed.
Menges G. IKV ) Aachen 1980
9. Morphy, R. D. 1999. Behaviour of fiber reinforced polymer FRP stirrups as shear
reinforcement for concrete structures. MS thesis, Univ. of Manitoba, Winnipeg,
Manitoba
10. Weber, A. and Witt, C. 2010. The New Safety Approach for the Durability of
advanced Composites in Concrete. Proceedings of the 8th short and Medium Span
Bridges (SMSB) Niagara Falls, ON, Canada, 3-6 August, 8 p.

390

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

SUSTAINABILITY PROBLEMS ASSOCIATED WITH THE HEAT


AND FIRE PROTECTION OF FIBER REINFORCED POLYMERS,
(FRP)
H. Ibrahim1 and I. Mahfouz2
1
2

Ph.D. Graduate Student, University of Illinois, Urbana, Il, USA


Professor of structural Engineering, Banha University, Egypt, Director of the Egyptian FRP
Standing Code Committee & Member of the ACI 440 committee

ABSTRACT
One of the main problems that hinder sustainability and the efficient utilization of FRP in
the construction fields are the problems associated with excessive heat and fire resistance of
FRP systems. Recognizing the fact that CFRP composites cannot sustain the exposure to
elevated temperature above the glass transition temperature of the epoxy resin, Tg, which
for most commercially available epoxy resins varies from 60 to 80oC, excessive heat
exposures above these values can adversely affect the efficiency of the CFRP strengthening
works, while fire will defiantly result in a total loss of CFRP strengthening works if not
properly protected.
It should be pointed out that the provisions dealing with the adverse effects of fire and heat
on FRP works in the FRP guidelines and codes do not fully satisfy the sustainability
principles and must therefore be revised.
Several shortcomings in the FRP practice, as well as, in the FRP guidelines and codes that
would hinder the satisfaction of the sustainability principles were recognized and presented
in this work along with means of rectifying such shortcomings.
In this work three case studies for the heat and fire protection systems of FRP strengthened
projects that fully meet the sustainability requirements are also presented. The case studies
will include test results that demonstrate the efficiencies of heat and fire protection systems
used in providing full protection of FRP strengthening works.
1. INTRODUCTION
Sustainability in the most basic level requires that the interdependence between our
economic activities, our social values, and our planet's capacities should be taken into
consideration at all stages of any construction works. It is about finding ways to meet our

391

needs in one area without diminishing or damaging another, today and for future
generations. It can be stated that, the beneficial overall effects of the implementation of the
environmental, economical and social sustainability requirements for the design and
construction of new projects have been widely recognized and have become essential and
integral requirements for the construction of new structures. Accordingly, sustainability
requirements have become essential design principles incorporated as standard design
practices and mandatory design requirements in building codes and regulations worldwide.
However, despite its importance and beneficial impacts on the construction industry at
large, the same kind of attention has not yet been paid to the sustainability requirements in
other sector of the construction fields involving strengthening, repair and retrofitting of
deficient structures in general and those involving the use of FRP in particular. This is
attested to by the examination of the durability provisions of the ACI 440.2R-08 guidelines
as well as other FRP codes and guidelines and in particular those related to FRP
strengthening works subjected to excessive heat and fire that clearly reveal that such FRP
guidelines and codes [1-8] do not satisfy the basic sustainability requirements. In fact, it is
important to point out that the satisfaction of the sustainability principles should be equally
applied to both new structures and strengthening works. Accordingly, such shortcomings in
the engineering practice have to be rectified.
The main objective of this paper is to highlight the importance of incorporating
sustainability design principles in the design of FRP strengthening applications with
emphases on the applications related to the fire and heat protection of FRP strengthening
works. The other objective of the paper is the presentation of field applications
incorporating the sustainability principles in cases of structures strengthened using FRP and
subject to excessive heat and fire.
2. COMPLIANCE OF FRP STRENGTHENING SYSTEMS WITH THE
SUSTAINABILITY REQUIREMENTS
The FRP applications can be classified with respect to fire and heat resistance into the
following two groups:
1- FRP Applications where the use of fire or heat protection systems are compulsory.
Such cases include:
FRP systems subjected to excessive heat that exceeds the Tg temperature of the
resin used.
FRP applications that require strengthening that exceed the strengthening limits
specified in the ACI 440.2R-08 FRP guidelines.
For the preceding two cases FRP systems cannot be used without insuring that the
FRP systems are properly and fully protected against heat or fir, a matter that will
hinder the use of FRP at large in the case of the unavailability of such systems. It is
noted that such cases will satisfy the sustainability requirements only and only if the
fire and heat protection system are proven to be effective, durable and its efficiency
is guaranteed for extended life spans.
2- FRP Applications where the FRP guidelines allows the use of FRP systems without
fire or heat protection systems:

392

Such systems include the cases where the designs of FRP applications are governed
by the strengthening limits specified in the FRP guidelines and codes. As a
consequence of adopting such design principle, the FRP strengthening work will be
the totally lost but without structural collapse when exposed to fire and excessive
heat exceeding the Tg temperature of the resin
Despite that fact that the safety of the structure will be maintained in the cases of
exposure to excessive heat and fire but the economical, environmental and social
aspects of sustainability principles will be violated. As a result, the use and the
enclosure of such design practices and principles in the FRP guidelines and codes
need to be reconsidered. In addition the implementation of such practices in real life
applications should be prohibited or at least discouraged.
That is, in matters pertaining to structural strengthening it is recommended that the
FRP guidelines and codes should not only be concerned with the satisfaction of the
strength and serviceability limit states but should also consider the sustainability
and durability requirements as one of the limit states of design.
3. FRPS VITAL CONTRIBUTION TO SUSTAINABLE DEVELOPMENT
FRP systems have become one of the most widely used systems for the strengthening and
retrofitting of structures in the world. Accordingly, FRP systems have a critical role to play
in the construction fields and subsequently must be further developed to meet the
sustainability requirements in order to be able to contribute in the future success of
sustainable development of structural strengthening.
Accordingly, a special attention has to be paid to the enhancement of the FRP strengthening
systems when exposed to fir and excessive heat since the FRP systems in their current
status do not meet the basic sustainability requirements for the reasons presented in the
preceding section. In this regard, the following measures and enhancements must be
considered:
1- The low glass transition temperature of the commercially available resins used in FRP
composites is one of the major factors that need to be enhanced in order to achieve the
sustainability objectives.
2- The lack of adequate fire protection systems that can be relied upon to fully protect FRP
strengthening works is the main contributor in not satisfying the sustainability
objectives. Such development, will also fully guarantee the safety of the structure in the
cases where it becomes essential to exceed the strengthening limits imposed by the
guidelines. This will also result in opening the way for other applications using FRP not
allowed by the current FRP guidelines and codes.
In general it can be stated that a major revision has to be made in FRP guidelines and codes
that will ensure the full incorporation of the sustainability principles and subsequently can
lead to the rectification of a number of shortcomings and scopes of applications associated
with FRP practices.

393

4. IMPLEMENTATION OF FRPS SUSTAINABLE DEVELOPMENT


In the following, real life field applications where the sustainability principles were
incorporated and implemented are presented.
4.1 Sustainable Heat Insulation Material for CFRP Works
4.1.1 Strengthening of a Box Girder Bridge, Egypt
Considering the advantageous of the CFRP technology, and in particular the advantageous
of being able to complete the entire work in the shortest time possible compared to other
methods and without the need to shore the structure or interrupt the flow of traffic were the
decisive factors that influenced the adoption of CFRP technology as the strengthening
method most suitable to use for the strengthening of the upper deck of a box girder bridge.
Recognizing the fact that CFRP composites cannot sustain the exposure to elevated
temperature above the glass transition temperature of the epoxy resin, Tg, which for most
commercially available epoxy resins varies from 60 to 80oC, the use of the proposed FRP
system became questionable since FRP strengthening works will be subjected to elevated
temperature that exceeds 120oC during the applications of asphalt.
The problem was not in finding a heat protection layer by itself, but in recommending a
sustainable FRP strengthening system that can provide adequate protection against elevated
temperature while possessing an early age high compressive strength capable of resisting
the applied loads from pavement works as well as from truck loads during the life span of
the bridge. As a result, most if not all of the commercially available heat insulation
materials did not meet such design criteria. In this regard, field and laboratory research
program were carried out. The program comprised the testing of a number of heat
insulation materials modified to suit the required compressive strength.
The ambient temperature, the temperature of the asphalt layers as well as those of the
CFRP laminated strips were recorded using pre-installed thermocouples. The temperature
of the hot asphalt layers were raised to 140oC and kept at the temperature values
corresponding to the temperature resulting from the actual applications of asphalt layers
during the pavement works. The test results are shown in Figure 1.
Examination of the test results reveals that the results obtained from all mixes were
successful and met the specified design criteria for both strength and heat insulation. The
70 mm thick VC type mortar gave the best heat resistance performance. However, it was
decided to use the PRC type mortar since the compressive strength of the VC mortar did
not satisfy the specified characteristic compressive strength.

394

HEAT EXPOSURE TEST


160

140

TEM
PER
A
TU
R
E( C
)

120

AMBIENT TEMPERATURE
ASPHALT TEMPERATURE
"VC" MORTAR - 50 MM THICK
"VC" MORTAR - 70 MM THICK
"PRC" MORTAR 50 MM THICK
"PRC" MORTAR 70 MM THICK
"LC" MORTAR - 70 MM THICK
" LC" MORTAR - 50 MM THICK

100

80

60

40

20

0
0

30

60

90

120

150

180

210

240

TIME ( MINUTES )

Fig. 1. Elevated heat exposure test results.

Fig. 2. Installed CFRP Laminated Strips on the Bridge and Monitoring system

AMBIENT
TEMPERATURE
HEAT INSULATING
MATERIAL
ASPHALT

13.00

12.00

11.00

10.00

9.00

8.00

CFRP STRIPS

7.00

130
120
110
100
90
80
70
60
50
40
30
20
10
0

6.00

TEM
PERATURE, (C)

It is noted that during the application of the asphalt layers continuous monitoring of the
temperatures at a number of locations of the CFRP laminated surfaces were made using
previously installed thermocouples on the surface of the CFRP strips (Figure 2). The
monitoring system was used to regulate the rate of application of the asphalt layers such
that the temperature of the CFRP layers does not exceed 60oC; the Tg temperature of the
epoxy resin of used. It is noted that the recorded temperature during the application of the
asphalt layers did not exceed the temperature of 40oC, previously obtained during the preinstallation tests, (Figure 3). However, the temperature continued to rise and reached a peak
value of 51oC after 6 hours, but was followed by a gradual drop of temperature. The
temperature recorded in the following day showed stable condition within the acceptable
limits.

TIME ( HOURS )

Fig. 3. Measurements of Temperature During The Application Of Asphalt Layers.

395

4.2 Sustainable Fire Protection of CFRP Works


4.2.1 FRP Strengthening of the Main Conference Hall of Bibliotheca Alexandria
The main conference hall has a continuous two span folded plate roof of 44 m and 12 meter
long, respectively, along with a paneled beam floor supported by the folded plate roof by
means of a number of hangers, as shown in Figure 4.

Fig. 4. The Main Conference Hall Alexandria Library.


The paneled beam floor is used for utility purposes and carries the loads of a number of
generators, a fire extinguishing system, the conference hall curtains, the air conditioning
units and systems, the artificial ceiling of the conference hall, live loads on the paneled
beam floor. As a result of the renovations that were planning to take place that involved
major changes in the utility systems and types of unites supported by the paneled beam
floor that will result in significant increase in the magnitudes of applied loads, a complete
structural safety assessment became essential. The results obtained revealed that the folded
plate roof after the application of the planed renovations did not satisfy the safety
requirements of both the Egyptian and the American codes of practices [8; 10]. In this
regard it was found that the area of main tensile reinforcement of folded plate roof was
inadequate. Various strengthening options were studied from economical, constructability
and technical standpoints. Recognizing the superior properties and advantages of FRP
systems, it was found that the use of externally bonded Carbon Fiber Reinforced Polymers,
(CFRP) laminated strips and sheets were the most suitable strengthening method to use for
such applications.
It is noted that the required increase in the strengthening was 15% which was below the
strengthening limits of the FRP guidelines and code. Accordingly, the code's provision that
allows the use of CFRP without fire protection could have been used. However,
recognizing the fact that such approach did not meet both the economical and social
sustainability principles, it was decided to employ a fire protection system capable of
guaranteeing full protection of the FRP strengthening works and the structure for the 2
hours fire resistance rating of the project. A number of fire protection systems were
presented by the suppliers, none of the official systems provided by the FRP suppliers were
proven to be effective to provide required fire protection for the FRP works and they were
all rejected. This matter that was about to result in a complete rejection of the use of FRP
strengthening system in this project.
396

A fire protection proposal was approved that relies on the use of a fire extinguishing system
along with heat insulation material that will provide a sustainable fire protection system for
the project. In this regard, a fire extinguishing system will be automatically activated at a
temperature of 57oC, (which is below the Tg temperature of the epoxy resin) was used,
along a heat insulation layers capable protecting the CFRP laminated strip against excessive
heat which can reach up to 300oC, assuming that in the incident of fire the prevailing
temperature after the fire has been extinguished will not exceed 300oC. The system was
implemented with recommendation of the need of carrying out continuous maintenance and
tests of the fire extinguishing system in order to guarantee a long term and sustainable
efficiency.
4.2.2 El-Harram Project, Makah, KSA : FRP Strengthening of El-Haram Columns
In general, it should be pointed out that for any FRP structural strengthening that exceeds
the strengthening limits specified in the guidelines which could be in the range of 40%, the
CFRP strengthening works must be fully protected against fire exposures, otherwise a total
loss of CFRP strengthening works shall take place resulting in a total collapse of the
element. As such, any fire protection system that cannot meet the preceding important
condition shall be totally unacceptable.
Due to the ongoing renovations and new extension in El- Haram in Makah, the need to
increase the load carrying capacities of over four hundred columns to double their current
capacities became very essential and required immediate attention and quick solutions to be
implemented. Under such circumstances, FRP was ideally suited strengthening method that
can fulfill such requirements. Needless to say that such increases in the ultimate strengths
of such columns exceed the strengthening limits imposed by the guidelines and code a
matter that required that all FRP strengthening woks should be effectively and fully
protected against fire.
In this regard, extensive survey and studies on the fire protection systems currently
employed in the FRP fields was performed. However, Based on extensive tests that were
carried out at Warringtonfiregent Laboratory, Gant, Belgium proved that non of the
commercially available fire protection systems were capable of meeting the 2 hours fire
resistance rating specified in the project. It was realized that such fire protection systems
can provide only adequate protraction to the concrete and steel reinforcement without
providing any protection to FRP .Accordingly, such cases may be suitable only for the
cases where the required strengthening level does not exceed the strengthening limits
specified in the guideline,which does not represent our case, as shown in Figure 5

397

1200

COMPARISSON BETWEEN THE FIRE PROTECTION SYSTEMS

1000

TEM
PERATURE(C)

800

UNE 23-093-81
ASTM 119

600

TCR SYSTEM 50 MM ( USA TEST)


PROPOSED SYSTEM ( 30 MM)

400

TCR SYSTEM 50 MM ( GANT, BELGIUM


TEST )
TYFO SYSTEM( 57 MM)

200

0
0

20

40

60

80

100

-200
TIME (min)

Fig. 5. Test Results of the Fire Exposure Tests on Commercially available Fire
Protection Materials
Accordingly, a new fire protection system has to be developed in order to guarantee not
only structural safety but also, sustainability since it will not be acceptable to lose all FRP
works in the case of fire a matter that would require the re-application of the FRP
strengthening works which cannot be allowed due to the nature and importance of the
structure. As a result of the extensive theoretical and experimental research works, a fire
protection system was developed and tested at Warringtonfiregent Laboratory, Gant,
Belgium. Initially a set of trial fire tests were performed on a set of FRP wrapped concrete
cylinders and protected using the developed fire protection system. As a result of the
successful results obtained, a full scale wrapped column using 5 layers of SikaWrap-530C
and protected using 50 mm of the fire protection layers was tested in accordance with
ASTM 119 standard fire test (Figures 6&7). During the entire duration of the fire test the
column was subjected to an axial compressive load of 2119 KN which is greater than the
ultimate compressive load of the unwrapped column. It should be pointed out that the
duration of the fire test was 135 minutes during which the fire protection layers were
capable of providing full protection to the CFRP wrapping sheets. In this regard, the
column did not suffer any form of distress or failure, as well as, the CFRP wraps remained
intact and did not suffered any forms of debonding, delamination or any other forms of
deterioration (Figures 6 & 7).
The results obtained revealed that the proposed fire protection systems are capable of
providing full fire protection satisfying fire resistance rating of at least 120 minutes [11]. It
is noted that the proposed system was approved and implemented for the fire protection of
the project.

398

a)CFRP Wrapping b) Column Before Test


c) After Test (After Cooling Down)
Fig. 6. Full Scale Fire Exposure Test of CFRP Wrapped R.C. Column

Fig. 7. Test Results of the Full Scale Fire Exposure Test


5. CONCLUSIONS
1- Sustainability principles should be satisfied not only for the design of new structure but
also for strengthening, retrofitting and repair works of deficient structures.
2- Successful implementations of the sustainability principles in real life field applications
for the strengthening of three major structures using FRP have been presented in this
paper.
3- The sustainability design principles should be included as essential and integral parts of
FRP guidelines and code.
399

4- Provisions that accept the total loss of FRP works when exposed to fire should be
omitted from FRP codes and guidelines.
5- Heat and fire protection measures must represent an essential and integral part of the
design of any CFRP strengthening works in order to satisfy the sustainability design
requirements
6- Emphases should be placed on research works that aim at the enhancement of the
properties of the commercially available resin and the development of fire protection
systems capable of fully protecting FRP strengthening works.
6. ACKNOWLEDGMENTS
The contributions of Engineer Wassim Mahfouz and the R.& D. department of the
Kabbani Group, Jeddah, K.S.A during the Fire test of El Haram Project is highly
appreciated.
7. REFERENCES
1.
2.
3.
4.
5.
6.

7.

8.
9.
10.
11.

HBRC. 2005. The Egyptian Code of Practice for the Use of Fiber Reinforced
Polymers, (FRP) in Construction Field. (ECP 208), Cairo, Egypt.
Canadian Standard Association. 2002. Design and Construction of Building
Components with FRP. CAN/CSA-S806-02, Mississauga, Ontario, 177 p.
Canadian Standard Association. 2000. Canadian Highway Bridge Design Code.
CAN/CSA-S6-00, Mississauga, Ontario.
ACI Committee 440. 2008. Guide for the Design and Construction of Externally
Bonded FRP Systems for Strengthening Concrete Structures. ACI 440.2R.08,
American Concrete Institute, Farmington Hill, MI, 76 p.
fib (CEB-FIB). 2001. Externally Bonded FRP Reinforcement for RC Structures.
Technical Report, Bulletin 14.
Japanese Society of Civil Engineers (JSCE), (1997), Recommendation for Design
and Construction of Concrete Structures Using Continuous Fiber Reinforcing
Materials, Concrete Engineering Series 23, Edited by A. Machida, Tokyo, Japan,
325 p.
Intelligent Sensing for Innovative Structures (ISIS). 2001. Strengthening Reinforced
Concrete Structures with External Bonded Fiber Reinforced Polymers. (ISISM04-01),
Canadian Network of Centers of Excellence on Intelligent Sensing for Innovative
Structures, University of Manitoba, Winnipeg, MB, Canada.
ISO TC/71 SC-6 Committee. Non-traditional reinforcing materials for concrete
structures.
HBRC. 2007. The Egyptian Code for Design and Construction of Concrete Structures
ECCS 203, Cairo, Egypt.
ACI Committee 318. 2008. Code Requirements for Structural Concrete. ACI 318R08 and Commentary, American Concrete Institute, Farmington Hill, MI.
Test report. 2010. Warringtonfiregent Laboratory, Gent, Belgium, August.

400

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

EFFECTS OF AGING ON GLASS TRANSITION AND CREEP


BEHAVIOR OF EPOXY USED IN FRP STRENGTHENING
A. Jaipuriar1, J. Flood2, Y. Jeong2, C.E. Bakis1 and M.M. Lopez2
1
2

Department of Engineering Science and Mechanics, Pennsylvania State University, USA


Department of Civil and Environmental Engineering, Pennsylvania State University, USA

ABSTRACT
Fiber reinforced polymer structural strengthening systems typically utilize ambienttemperature cured resins. As aging can be expected to occur in ambient-cured epoxy resins
during service, it is important to understand the effects of aging on the mechanical
properties of the material. In the present investigation, glass transition temperature (Tg) and
creep behavior of a commercially available resin system were studied for aging at room
temperature for up to 100 days. A significant increase in Tg and decrease in creep rate
occurred with increasing age. Evidence suggested that both phenomena were caused by
rapid physical aging during the first 30 days of age, rather than post curing. Additional
results showed the beneficial effects of post curing and the mixed effects of short-term
moisture saturation on Tg.
1. INTRODUCTION
Epoxy resins are predominantly used as the adhesive and matrix material in FRP
strengthening systems. Like other thermosetting polymers, epoxy is made by mixing
ingredients (liquid epoxide and a curing agent) and allowing irreversible cross-linking of
the molecular network to proceed over time. Cooler temperatures slow the curing process,
while warmer temperatures accelerate it. The desired result of curing is a polymer with
rigid, or glassy, characteristics throughout the expected service temperature range. At
sufficiently high temperatures, polymers lose their desired glassy characteristic and behave
as rubbery materials. This phenomenon is known as the glassy-rubbery transition, and
various methods exist for the assignment of a characteristic temperature to it [1]. Below this
so-called glass transition temperature, Tg, the time dependent deformation of a polymer is
much less than that occurring above the Tg. The Tg of epoxy increases as cross-linking
progresses with cure (a type of chemical aging). The Tg of epoxy also increases due to
physical aginga process of equilibration and consolidation of the material that occurs
below the Tg at ever-slowing rates as temperature decreases [2]. Other chemical and
physical processes typically associated with degradation, such as simple moisture
absorption, hydrolysis, and photolysis, can decrease the Tg.

401

In strengthening civil infrastructure, structures are back in service within a few days or
weeks after installing the FRP system. Ambient-cured epoxies, at the age of a few weeks or
less, have glass transition temperatures relatively close to the environmental temperature
and therefore can exhibit magnified creep under mechanical loading. Several research
groups [3-5] have investigated creep at the FRP-concrete interface. However, little
emphasis has been placed on the effects of aging on the behavior of the epoxy and the FRP
strengthening system under sustained loading.
The objective of present investigation is to evaluate the Tg and creep of a room-temperature
cured epoxy resin as a function of aging time up to 100 days. The influence of moisture
absorption on the Tg is evaluated as well. The information obtained in this investigation is
useful for modeling the bond behavior of FRP/concrete joints under long-term loading.
2. EXPERIMENTAL PROGRAM
The experimental investigation consisted of dynamic mechanical analysis (DMA),
differential scanning calorimetry (DSC) and tensile creep tests on epoxy resin. The epoxy
resin system selected for the study is a commercially available, two-part, ambient-curing
resin system specifically formulated for saturating fibers and for bonding wet-layup or
pultruded FRP plates to concrete, masonry, and timber substrates. Prior to testing, the
materials were stored at room temperature (201C) in a sealed desiccator with silica gel.
One specimen was tested from each of three separately mixed batches. Thus, three
specimens were used for each type of test in order to obtain qualitative information about
batch-to-batch scatter. A few specimens were conditioned in a bath of de-ionized water
prior to Tg testing with the dry specimens.
2.1. Dynamic Mechanical Analysis
For the assignment of Tg, epoxy films of 0.3 mm thickness were molded between two
precisely machined flat steel surfaces. The films were cut to planar dimensions of
approximately 305mm. DMA scans were done using a constant sinusoidal strain
excursion of 0.05% in tension and a loading frequency of 5 Hz. The Tg test sequence was
determined according to ASTM E1640 [6].Specimens were equilibrated at -50C and
temperature was ramped at the rate of 3C per minute until 120C was reached. Three
sequential heating/cooling cycles were run for each specimen to study the effect of
temperature excursions on the Tg. Further details regarding assignment of the Tg by various
methods such as storage modulus, loss modulus, and Tan can be found in the literature [1,
6, 7].
2.2. Differential Scanning Calorimetry
DSC tests were utilized to study the glass transition temperature and the progress of epoxyamine crosslinking reaction. ASTM D3418-08 was followed to assign the Tg [8]. A sample
mass of about 5 mg was used in each scan. Large volume stainless steel pans of 60 l
capacity with a lid and rubber seal capable of sustaining high internal pressure was used in
the scans to prevent any leakage of volatiles. Scans were done at 3C/min from 0C to
250C after equilibrating the samples at 0C.
402

2.3. Tensile Creep Testing


Epoxy sheets were cast in open molds to a thickness of approximately 3.2 mm. Coupons
were machined to planar dimensions of about 15013 mm. A constant tensile stress of
about 10 MPa was applied for roughly 5000 minutes (or 83 hours) at room temperature
(201C) using dead weights. Two 12.7-mm extensometers were used to measure strain on
opposite faces of the specimen to compensate for bending strain. Based on the measured
creep strains, the creep compliance (strain divided by stress) was computed.
3. RESULTS
3.1. DMA Characterization of Tg
DMA tests were carried out on specimens aged in a desiccator at room temperature for 7,
30 and 100 days. The three batches of material used for these tests are denoted A, B and C.
Typical plots of storage modulus, loss modulus and Tan for 7-day-old specimens are
shown in Figs. 1a, 1b and 1c, respectively. The Tg assigned according to the storage
modulus, which is an indication of onset of loss in stiffness, is around 40C for material
aged 7 days. The effect of temperature excursions up to 120C is evident in the consecutive
DMA scans (Run 1, Run 2 and Run 3) on the same specimen. The Tg increase of roughly
20-30C is hypothesized to be due to an increase in the crosslink density of the polymer at
the elevated temperatures. A summary of Tg values for the various batches and ages based
on the first DMA scan is shown in Fig. 1d. A 10-15C increase in first-run storage modulus
Tg during the first 100 days of aging at room temperature can be seen. The source of this
increase is explained in the section on DSC results.
A preliminary study of the effect of moisture on Tg of epoxy was carried out. Thin film
specimens of 0.2-0.3 mm average thickness were molded and cured at room temperature
for 7 days. After 7 days of curing, DMA specimens were immersed in room temperature
de-ionized water. A saturation moisture content of approximately 2.5% by weight was
obtained after 3 days in water. DMA scans were done at ages of 30 days (23 days in water)
and 100 days (93 days in water). The average Tg for three batches of epoxy are reported in
Fig. 2a alongside results for unhydrated epoxy specimens of similar total age. It can be seen
that the Tg of 23-day-soaked specimens is comparable to the equivalently aged unhydrated
specimens based on storage modulus, whereas it is higher than the unhydrated specimens
based on loss modulus and Tan . An increase in Tg with moisture content has been
reported previously. According to Wu et al. [9], the presence of water (up to 2 weight %)
accelerates the curing rate and the evolution of Tg in epoxy resins. Further studies need to
be done to determine if the increase in Tg is due to the progress of a chemical reaction or
some other mechanism in the current epoxy resin. For the 93-day-soaked specimen, a
reduction in Tg was visible based on storage and loss moduli, but the Tan Tg indicated an
increase. The Tan curve had an obscured peak at around 65C (Fig. 2b). Yang et al. [10]
have reported the splitting of Tan peaks in aqueous solution. The splitting can be
attributed to plasticization of epoxy in water or non uniform drying during DMA testing,
with the obscured peak associated with the Tg of a wetter phase and the higher peak being
Tg of a drier phase. It is also interesting to note that the Tg assigned according to the Tan

403

peak can be deceptive, as the stiffness of the material decreases considerably at 20C below
the Tan Tg, according to the storage modulus Tg.

Run 1
Run 2
Run 3

0.45
TgR1=43C
TgR2=50C
TgR3=59C

3
2
1

0.40
0.35

Loss Modulus (GPa)

Storage Modulus (GPa)

TgRun1=49C
Run 1
Run 2
Run 3

TgRun2=67C
TgRun3=76C

0.30
0.25
0.20
0.15
0.10
0.05
0.00

0
-50 -30 -10

10

30

50

70

-0.05

90 110 130

-50 -30 -10

10

30

50

(a)

1.4
1.2

TgR1=60C

Run 1
Run 2
Run 3

TgR3=92C

Tan

0.8
0.6
0.4
0.2
0.0
-40 -20

20

40

60

Storage Modulus
Loss Modulus
Tan

80

TgR2=85C

1.0

90 110 130

(b)

Glass Transition Temperature (C)

1.6

70

Temperature (C)

Temperature (C)

70
60
50
40
30
20
10

A B C
7 days

80 100 120 140

Temperature (C)

A B C
30 days

A B C
100 days

Age of Specimen

(c)

(d)

Fig. 1. Tg datatypical repeated scans of (a) storage modulus, (b) loss modulus, (c) Tan
for 7-day aged epoxy; (d) summary of first-run DMA Tg data for epoxy cured for various
numbers of days at room temperature, from batches A, B, and C.

404

23 days exposure

70

Tg (C)

0.8

Unhydrated specimen
Soaked in deionized H2O

60
50
40
80

SM

70

Tg (C)

LM

TD

93 days exposure

60
50

0.8

Storage Modulus
Loss Modulus
Tan

0.7
0.6

0.7
0.6

Hidden Peak

0.5

0.5

0.4

0.4

0.3

0.3

0.2

0.2

0.1

0.1

0.0

40

SM

LM

20

40

60

80

100

Tan

80

(Storage Modulus (GPa))/10


Loss Modulus (GPa)

90

0.0
120

Temperature (C)

TD

(a)
(b)
Fig. 2. DMA Tg results for water-saturated epoxy (a) comparison of Tg for soaked and
unhydrated specimens at different ages using storage modulus (SM), loss modulus (LM)
and Tan (TD); (b) raw DMA data for a typical specimen soaked for 93 days.
3.2. DSC Characterization of Tg
Three batches of epoxy (A, B and C) were aged in a desiccator for 7, 30, and 100 days at
room temperature for the DSC Tg tests. Figure 3 shows the evolution of the DSC Tg with
age. As was seen with DMA, the DSC Tg increases by roughly 10C or more in the first
100 days of age.

Glass Transition temperature (C)

70

7 days
30 days
100 days

60
50
40
30
20
10
0

Batch A'

Batch B'

Batch C'

Fig. 3. Summary of DSC Tg data for epoxy cured in dry conditions for various numbers of
days at room temperature, from batches A, B, and C.
Since DSC scans can provide the enthalpy of reaction of the specimen being tested as well
as the Tg, the test results can be used to determine if there is any correlation between the
degree of cure and the Tg of the epoxy at various ages. For this purpose, an epoxy
specimen that had been previously aged in a desiccator at room temperature for 9 days was
transferred to an oven with 40C circulating air to accentuate the possibility of post-curing
during aging. DSC samples were cut from the specimen and tested after 96, 120, and 240
hours in the oven.

405

The unreacted enthalpy of reaction, Hunreacted is calculated from the area under the
exothermic peak of the DSC scan. For the area integration, a baseline was constructed by
fitting a cubic spline across the peak of interest [11]. The degree of cure, ,was then
evaluated as = (Hfresh - Hunreacted)/(Hfresh), where Hfresh is the enthalpy of reaction of
freshly mixed epoxy.

Normalized Heat Flow (J/g)

Exo Up

2.4
1.8

Init+120 hrs @ 40C

75 J/g

1.2

0.0
-0.6
-1.2

64 J/g

Tg=62C

0.6

Tg=44C
Tg=64C

Tg=61C

Init+96 hrs @ 40C

66 J/g
72J/g
Init

-1.8
Init+240 hrs @ 40

-2.4
0

40

80

120

160

200

Temperature (C)

Fig. 4. Effect of 40C exposure of various durations on Tg and unreacted enthalpy of


reaction of epoxy initially aged in dry conditions at room temperature for 9 days.
Results of the post-curing investigation shown in Fig. 4 reveal an increase in Tg from 44C
in 9-day-old material to about 65C following the 10-day exposure to 40C. This
substantial increase exceeds that observed in room temperature aging for up to 100 days.
However, the residual enthalpy of reaction, Hunreacted, does not change appreciably in this
time period. The increase in Tg without further increase in degree of cure can be explained
by the phenomenon of physical aging, which is typically most rapid at temperatures just
below the Tg of the material [2]. Since exposure to 40C did not advance the degree of cure
of an epoxy specimen previously aged 9 days at room temperature after mixing, it can be
concluded that the increases in Tg observed in epoxy aged at room temperature (Figs. 1 and
3) are also likely to be due to physical aging.
3.3. Tensile Creep Test of Epoxy Coupons
Room-temperature creep testing was done on three separate batches (A, B and C) of
resin after 7, 30, and 100 days of aging at room temperature. The resulting creep
compliance curves are shown in Fig. 5a. The creep rate decreases significantly with the age
of the specimen at the start of creep testing. It is known that the rate of physical aging is
dependent on (Tg-T), where T is the conditioning temperature, which is room temperature in
this case [2]. The rate of aging is highest when the conditioning temperature is a few
degrees less than the Tg. As Tg increases further beyond the ambient temperature with
physical aging, the rate of aging (and further Tg increase) decreases considerably, which is
reflected in the larger difference in creep characteristics of 7- and 30-day-old specimens as
compared to the difference in creep characteristics of 30- and 100-day-old specimens.

406

0.006

7B''

7A''
30A''
100A''

0.0025

Compliance (1/MPa)

7C''

7A''

0.004
0.003
0.002

30A''

30B''

30C''

0.001

Compliance (1/MPa)

0.005

0.0020

0.0015

0.0010

0.0005
100C'' 100B''

0.000
0

100A''

1000 2000 3000 4000 5000 6000 7000 8000

Time (Minutes)

Log Time (Minutes)

(a)
(b)
Fig. 5. Creep compliance versus time(a) unshifted data; the line labels indicate the time
of aging in dry, room-temperature air prior to testing and batch identification; (b) master
creep compliance curve for epoxy of 100 day age.
A master creep compliance curve for 100-day-old epoxy resin incorporating the effects of
physical aging can be generated by horizontally shifting short term creep data obtained with
epoxy aged 7 and 30 days to coincide with short term creep data for epoxy aged 100 days,
as shown by the arrows in Fig. 5b. This procedure, known as the time/aging-time
superposition principal, allows for the prediction of extended duration creep behavior in
material aged 100-days based on the results of short-term creep tests [2].
4. CONCLUSIONS AND RECOMMENDATIONS
Glass transition, aging, and creep characteristics of an ambient-curing epoxy system used
for the bonding of FRP strengthening systems to structural elements have been investigated
for various aging conditions. After aging the material at room temperature for 7 days in a
dry environment, the Tg based on the dynamic storage modulus was roughly 40C. At 100
days of age at room temperature, the Tg increased by 10-15C based on the storage
modulus. Short exposures to temperatures of 120C rapidly increased the Tg even more
by roughly 20C based on the storage modulusdue to post-curing. Using DSC, it was
found that the evolution of Tg with aging at temperatures up to 40C is mainly a
manifestation of physical aging rather than post-curing. Since ambient-cured epoxies have
Tg values only slightly higher than the ambient temperature, the rate of physical aging can
be high at early age. Therefore, creep is rapidly reduced at early age as well. Tensile creep
test data confirmed a rapid decrease of creep with respect to room-temperature aging time
for epoxy aged up to 30 days. For epoxy aged at room temperature between 30 and 100
days, the additional reduction of creep with respect to aging time was less marked. A
preliminary investigation of the effect of moisture on the evolution of Tg showed
contradictory trends for different measures of Tg by DMA and non-monotonic trends over
time for a given measure of Tg. Further research is needed to better understand the
complicated interactions among moisture, aging, Tg, and creep behavior. In addition, the
implications of creep on the bond strength of installed FRP strengthening systems need to
be investigated as well.

407

5. ACKNOWLEDGEMENTS
The authors gratefully acknowledge the financial support of the National Science
Foundation through grant CMMI-0826461 and an REU supplemental grant.
6. REFERENCES
1.

2.
3.
4.
5.
6.
7.

8.
9.
10.
11.

Bakis, C.E., 2008. Elevated Temperature Capability of Epoxy Resins: Issues and
Developments. Proc. US-South America Workshop on Innovative Materials for Civil
Infrastructure, Research and Education, M.M. Lopez, Ed., National Science
Foundation, 13-15 October, Santiago, Chile, 4 p. (CD ROM).
Struik, L.C.E. 1978. Physical Aging in Amorphous Polymers and Other Materials,
Elsevier.
Choi, K.K., Meshgin, P., and Reda Taha, M.M. 2007. Shear Creep of Epoxy at the
Concrete-FRP Interfaces. Composites Part B: Engineering, 38:772-780.
Diab, H., and Wu, Z. 2007. Nonlinear Constitutive Model for Time-Dependent
Behavior of FRP-Concrete Interface, Composites Science and Technology, 67:23232333.
Gullapalli, A., Lee, J.H., Lopez, M.M., and Bakis, C.E. 2009. Sustained Loading and
Temperature Response of Fiber-Reinforced Polymer-Concrete Bond. Transportation
Research Record: J. Transportation Research Board, 2131:155-162.
ASTM Standard E1640. 2010. Standard Test Method for Assignment of the Glass
Transition Temperature by Dynamic Mechanical Analysis. Annual Book of ASTM
Standards, ASTM International.
Jaipuriar, A., Flood, J., Bakis, C.E., Lopez, M.M., and He, X., 2010. Glassy-Rubbery
Transition Behavior of Epoxy Resins used in FRP Structural Strengthening Systems.
Proc. 5th Intl. Conf. FRP Composites in Civil Engineering, CICE 2010, Eds. L. Ye, P.
Feng, and Q. Yue, Tsinghua University Press, Beijing, 1: 397-400.
ASTM Standard D3418. 2010. Standard Test Method for Transition Temperatures
and Enthalpies of Fusion and Crystallization of Polymers by Differential Scanning
Calorimetry. Annual Book of ASTM Standards, ASTM International.
Wu, L., Hoa, S.V., and Ton-That, M. 2004. Effects of Water on the Curing and
Properties of Epoxy Adhesive used for Bonding FRP Composite Sheet to Concrete.
Journal of Applied Polymer Science, 92:22612268.
Yang, Q., Xian, G., and Karbhari, V. 2008. Hygrothermal Ageing of an Epoxy
Adhesive used in FRP Strengthening of Concrete. Journal of Applied Polymer
Science, 107: 2607-2617.
Hhne, G.W.H., Hemminger, W.F., and Flammersheim, H.J. 2003. Differential
Scanning Calorimetry, 2nd edition, Springer.

408

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

MORE THAN TWENTY YEARS OF FIELD INSTALLATIONS


OF COMPOSITES WITH DURABILITY TESTING
AND FIELD OBSERVATIONS
E. R. Fyfe 1, R. Watson 2 and M. McCullagh 3
1

Fyfe Co. LLC - San Diego, California, USA


R.J. Watson, Inc. - Buffalo, New York, USA
3
Fyfe Co. LLC - San Diego, California, USA
2

ABSTRACT
The Tyfo SEH (primary glass fiber) and Tyfo SCH (primary carbon fiber) composites
have been tested 10,000 hours by the Aerospace Corporation for Fyfe Co. LLC for the
Department of Transportation, California. A summary of these results are given.
Field applications, which are the real test of durability, have been in place since 1990 which is
over 20 years of experience. Applications for seismic, marine, and corrosion have been
installed. The paper discusses early applications in New York City, New York thruways near
Buffalo, Pennsylvania, and Nikko Hotel in Los Angeles.
The key field applications have been visited recently and durability results will be presented.
There are now over 9,000 installations of the Tyfo Systems. No actual durability problems
have been reported, even when in high ultraviolet areas.
1. INTRODUCTION
In 1987, the first North American work on composite strengthening was started in San Diego.
The Tyfo System was proposed for use by the Department of Transportation in California. A
test program was started on twelve large-scale columns at the University of California, San
Diego.
The testing was very successful, and demonstrated that the strength and ductility of circular
and rectangular columns could be increased by wrapping with high strength fiber composites.
Rectangular columns were wrapped in the shape of the rectangle with beveled or round
corners, and although not as efficient as a circular wrap, can obtain increases in strength and
ductility of up to ten or more.

409

In one test, a circular column that had been severely damaged by testing without a jacket
was repaired and retrofitted with the Tyfo SEH System, and achieved a ductility of 10 (the asbuilt specimen achieved a ductility of 3 upon re-testing.
Once the testing had shown great promise, the question of how long it would last in the field
became the main technical barrier to extensive use for repair of civil structures.
The Department of Transportation in California (Caltrans) arranged with the Aerospace
Corporation to do long-term (10,000 hour) durability testing. This is required of each
company with a composite system to obtain Caltrans approval. Other durability tests include
those conducted at the University of Toronto.
2. DURABILITY LABORATORY TESTING
In 1996, the Aerospace Corporation, consulting with SAMPE and Caltrans, developed a series
of tests to determine environmental durability. The test results, in general, are shown in
Table 1. Both the Tyfo SEH and SCH Systems were submitted for the 10,000 hour testing.
Table 1. Test Results
Modulus of elasticity (ksi)
Test
SEH 51S (glass)
SCH 41S (carbon)
1,000
3,000
10,000
1,000
3,000
10,000
hrs.
hrs.
hrs.
hrs.
hrs.
hrs.
Control
3.96
9.15
100% Humidity/38C
4.04
3.94
3.93
9.37
9.55
9.24
Salt Water
4.03
4.02
3.88
9.64
9.42
9.79
pH9.5 CaCO3
3.85
4.00
3.88
9.45
9.50
9.50
Dry Heat at 60C
3.85
4.05
9.63
9.29
Freeze/Thaw (20 cycles)
4.02
9.88
References Aerospace Corporation Testing
In the field, the systems are installed according to the same procedures. An initial
thickened epoxy layer, which is allowed to tack, is placed prior to the composite
application and a finish fiber layer and/or a thickened epoxy finish is applied to the
installed composite. Paint is also applied for ultraviolet resistance.
It is important to note that for both the Tyfo SEH and SCH Systems, the design modulus
was not affected by exposure testing. The design criteria developed (refer to ICBO ES
ACI25 for criteria) is based on strain compatibility, which is directly related to the modulus
(see Hooke's Law f = sE). The strain limitations imposed on the design (no design shall be
for strains beyond 0.004) insure both the client and the engineer of record that the composite
will provide the required additional capacity for the service life of the structure.
We are aware that some extreme exposure conditions will affect the strength of the composite
and effectively "shorten" the stress-strain curve (line in the case of advanced composites).
However, the strain-based design insures that these high stresses will not be reached (during
the specified design loading) in the installed system.

410

The results are simple and show that both the Tyfo SEH and the Tyfo SCH Systems work
well, and that both, with the proper reduction factor, will last for many years (100 years is
the design target).
Note that the strength of SEH-51A was reduced after 10,000 hours by 36% in the 100%
Humidity tests. Further 15,000 hour (5,000 extra hours) testing showed no additional loss of
tensile strength. No loss of strength was observed in the other tests. The strength of the
SCH-41 was reduced after 10,000 hours by 16% in the 100% Humidity test. No loss of
strength was observed in the other tests. No loss of stiffness (tensile modulus) was observed
in either the SEH-51A or the SCH-41 Systems.
3. APPLICATIONS DEMONSTRATING DURABILITY
In order to extrapolate laboratory results, field application data should be considered. It could
be that a composite system that shows losses of strength in the laboratory could last 100 years
in the field with no problem because it never sees extreme conditions as set up in the laboratory.
Corrosion and non-corrosion reinforced concrete applications need to be treated
differently. It can be expected in corrosion applications that the composite jacket will be
stressed over a period of time by the latent corrosion continuing in the system.
3.1 Nikko Hotel - Non-Corrosion - Seismic Application
The Nikko Hotel, in Beverly Hills, California (design authority - City of Los Angeles), had 34
short rectangular columns damaged in the Lander earthquake. So, it was in a hazard
situation. The Hart Consultant Group (Dr. Gary Hart) investigated all possible retrofit
solutions and selected the Tyfo SEH System to repair the columns in the hotel. The design
of the composite jackets was carried out by Dr. Nigel Priestley (See Figure 1.)
In October of 1993, the retrofit was completed. This completion was just in time for the
Northridge earthquake. The system worked as designed and allowed for large
displacements with no damage to the columns or the Tyfo SEH jacket. It should be noted that
the hotel was 14 miles from the epicenter in Northridge and one mile closer than the bridge
that collapsed at Highway 10 and La Cienaga.

(a)
(b)
Fig. 1. Kikko Hotel: (a) Column Wrapping, Oct. 1993; (b) Completed Column Jan. 1994

411

The February 2011 visual inspections show no sign of deterioration of the Tyfo SEH System
in place for eighteen years.
3.2 New York City Bridge - Corrosion Application
In May of 1994, a corrosion repair was completed on an overpass over FDR Drive near the
Manhattan Bridge.
The center bent has three 600 mm by 1200 mm rectangular columns, which had
deteriorated down to the steel reinforcing. The Tyfo S System was applied on two of the three
columns. Recent inspection shows that the retrofitted columns are in good condition, while the
patched column needs further repairs (See Figure 2).

(a) City of New York Structure

(b) After Tyfo S System application

FDR Drive

(Right column not wrapped)

(c) Reference column was


patched once after 10 years and
currently requires patching again

(d) Tyfo S System application

Fig. 2. New York City Bridge

412

3.3 New York Thruway in Buffalo (1-90) - Corrosion Application


Any resident of New York State can tell you that the reinforcing bars on many of the freeway
structures are corroding with extensive cracking and spalling of the concrete columns. It is
not unusual to see bridges in repair with extensive patching, injection, or replacement projects.
In October of 1994, two 1000 mm diameter columns on a structure in Buffalo (Figure 3), at
Exit 52, were retrofitted using the Tyfo SEH System. The columns are salt sprayed all
winter in this harsh environment.

Fig. 3. City of New York Thruway in Buffalo

Fig. 4. Recent visual inspections show no deterioration of the system after seventeen winters
(March 2011)

Fig. 5. Corrosion and spalling continue on other unwrapped elements (March 2011)
413

Recent visual inspections show no deterioration of the system after seventeen winters as
shown in Fig. 4 while Fig. 5 shows the corrosion and spalling continue on other unwrapped
elements.
3.4 Pennsylvania
One example of an FRP bridge retrofit that may qualify as a more severe environment is the
I-84 Bridge, East of Scranton, Pennsylvania.
During a routine evaluation of the bridge, it was determined that the six, 1.5m diameter
columns supporting this 143 m, four-span, continuous steel plate girder structure had
inadequate confinement steel (Figure 6). Previous retrofits consisted of taking threaded
steel rebar and tightening them with turnbuckles around the base and top of the columns.
This additional reinforcement is then covered with gunite concrete. While this is an
effective seismic retrofit technique, it is also very costly.

Fig. 6. I-84 Bridge East of Scranton, Pennsylvania

Fig. 7. Tyfo SEH-51 System in very good condition the full height of the columns after
18 years
414

When the option of using FRP to retrofit the columns was proposed to the PennDOT
engineers, they quickly approved it due to the cost savings, speed of application, and noncorrosive features of the FRP.
An engineering analysis was performed by HDR Engineering, out of Pittsburgh, which
determined that eight layers of glass FRP were to be applied in the upper and lower shear
zones of the columns with two layers being applied in the flexural zones. The project
started in July of 1993 and took about one week to complete at a total cost of $70,000, or
$8.00/ft2, of FRP material installed.
In November 1998, PennDOT issued a final report on the performance of the glass FRP
seismic upgrade. Their observations show that the FRP material has performed adequately and
that there are no signs of chalking, bleaching, delimitation or stress resulting from the
possibility of accumulated moisture beneath the wrap (as shown in Figure 7). The report goes
on to recommend the use of FRP for further projects as a structural repair material.
4. CONCLUSION
Although the use of Tyfo Systems is relatively new for construction, the early durability
results, both in the laboratory and the field, are encouraging. Appendix E shows additional
photo documentation of corrosion control related projects. Our plan is to continue to monitor
these applications and do some field composite sampling.
5. REFERENCES
1.
2.
3.
4.
5.
6.
7.

Sheikh, S.A., Homam, S.M., Pernica, G., and Mukherjee, P.K. 2000. Durability of
FRP Used in Concrete Structures. University of Toronto Research Report: HS-0100. February.
Steckel, G.L. 1999. Environmental Durability of Fyfe Company SEH51 Tyfo S EGlass Reinforced Epoxy Composite. The Aerospace Corporation Contract No.
SAMPE 0001. October.
Priestley, M.J.N. Seible, F., and Fyfe, E.R.. 1992. Column Seismic Retrofit Using
Fiberglass/Epoxy Jackets. NSF Workshop on Bridge Engineering Research in
Progress. PH. 247-251, November.
Falabella, R., Neuner, J., Isley, Jr., P.P. 1993. Environmental Durability and Resin
Stoichiometry Studies of Tyfo S High Strength Column Wrapping Systems. Report
LSR 932822. Hexcel Corporation, Pleasanton, CA.
Watson, R.J. 1994. Column Retrofit of Short and Medium Span Bridges and
Structures Using High Strength Fiber Composites. Development in Short and
Medium Span Bridges Engineering '94 Proceedings, August, pp. 983.
Watson, R.J. 1998. Case Histories of Structural Rehabilitation Utilizing Composites
in Industrial and Marine Environments. ICCI '98 Proceedings, January, pp. 113-124.
Baron, K.K. and Ziehr, J.J. 1997. Structural Retrofit of Midwest Bridges Using High
Strength Fiber Composites", IBC '97 Proceedings. June.

415

8.
9.
10.
11.

Fyfe, E.R., Watson, R.J., and Watson, S.C. 1996. Long Term Durability of
Composites Based on Field Performance and Laboratory Testing. ICCI '96
Proceedings, January, pp. 982.
Watson, R.J. 2000. Fiber Reinforced Polymers - A Cost Effective Solution for
Extending the Life of Our Aging Infrastructure. ASTM Standardization News,
February.
Sukley, R. and Howrylak, C. 1998. Fiber Column Wrap Seismic Retrofit Report.
Pennsylvania DOT Research Report No. 93-052, November.
Steckel, G.L. Durability Testing by the Aerospace Corporation.

416

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

VEHICLE CRASH TESTING OF A GFRP-REINFORCED PL-3


CONCRETE BRIDGE BARRIER
K. Sennah1, B. Juette2, A. Weber3 and C. Witt4
1

Professor, Civil Engineering Department, Ryerson University, Toronto, Ontario, Canada


Product Manager ComBAR, Division of Glass Fiber Reinforcement, Schck Bauteile GmbH,
Germany
3
Director, Research and Development, Schck Bauteile GmbH, Baden-Baden, Germany
4
President, Schoeck Canada Inc., Kitchener, Ontario, Canada
2

ABSTRACT
Corrosion of steel reinforcement due to environmental effects is a major cause of
deterioration problems in bridge barriers. Glass fibre reinforced polymers (GFRP), not only
addresses this durability problem but also provides exceptionally high tensile strength. The
special ribbed surface profile of the studied GFRP bars and end anchorage heads ensure
optimal bond between concrete and the bar and eliminate the use of custom made bar
bends. A recent design work conducted at Ryerson University on PL-3 bridge barrier
proposed the use of 16 mm and 12 mm diameter GFRP bars as vertical reinforcement in the
barrier front and back faces, respectively, with 12 mm diameter GFRP bars as horizontal
reinforcement in the barrier wall, all at 300 mm spacing. The connection between the deck
slab and the barrier wall utilized the GFRP headed end bars for proper anchorage. This
paper summarizes the procedure and the results of a recent vehicle crash test conducted on
the developed barrier. The crash test was performed in accordance with MASH Test Level
5 (TL-5), which involves the 36000V tractor trailer impacting the barrier at a nominal
speed and angle of 80 km/h and 15 degrees, respectively. Crash test results showed that the
barrier contained and redirected the vehicle. The vehicle did not penetrate, underride or
override the parapet. No detached elements, fragments, or other debris from the barrier
were present to penetrate or show potential for penetrating the occupant compartment, or to
present undue hazard to others in the area. No occupant compartment deformation
occurred. The test vehicle remained upright during and after the collision event.
1. INTRODUCTION
Bridges built prior to the 1970s did not use air-entrained concrete and coated reinforcing
steel bars to protect from the effects of freeze-thaw cycles and the application of winter deicing salt. This leads to corrosion-induced degradation in bridge elements as shown in Fig.
1. Accordingly, exposed bridge elements are all likely candidates for expensive
replacement on the majority of these older bridges. In November 2007, The Residential and
417

Civil Construction Alliance of Ontario, Canada, (RCCAO 2007) released a report on the
state of Ontario bridges, entitled Ontarios Bridges: Bridging the Gap. The report warns
that the integrity of Ontarios municipal bridge infrastructure and public safety are at risk
after years of deferred maintenance, irregular inspections, and lack of government
oversight. Recent media coverage on bridge collapses in Laval, Quebec and Minneapolis,
Minnesota, has highlighted the serious consequences of postponing actions to rehabilitate
or reconstruct deteriorated bridges and the urgent need to take timely responsible action to
safeguard the public from potential infrastructure failure. According to the Provincial
Auditors report in 2004, almost one-third of the approximately 2,800 provincial bridges
under Ministry of Transportation of Ontarios (MTO) jurisdiction are in need of major
rehabilitation or maintenance based on MTOs own figures. However, for the estimated
12,000 municipal bridges in Ontario, the RCCAO report stated that there is a lack of
information on their conditions and a capital investment of at least $2 billion will be
required over the next five years to rehabilitate this aging infrastructure. The RCCAO
report stated some recommendations to be made to promote the publics safety and the
sustainability of Ontarios bridges. One of these recommendations includes promoting
bridge engineering designs that improve the life expectancy and reduce maintenance costs
of bridges. This can be achieved by using fibre reinforced polymer bars.

Fig. 1. Corrosion-induced degradation of steel-reinforced


bridge barrier wall

Fig. 2. View of GFRP bars


with cast heads

2. FRP TECHNOLOGY
Fibre-reinforced polymers (FRPs), as non-corrodible materials, are considered as excellent
alternative to reinforcing steel bars in bridge decks to overcome steel corrosion-related
problems. The GFRP bars used in this study (Schck 2010) have tensile strength of 1188
MPa, compared to 400 MPa yield strength of the currently used reinforcing steel bars. The
special ribbed surface profile of these bars, shown in Fig. 2, ensure optimal bond between
concrete and the bar. Until recently, the installation of GFRP bars was often hampered by
the fact that bent bars have to be produced in the factory since GFRP bars cannot be bent at
the site. Also, bent GFRP bars are much weaker than straight bars, due to the redirection
and associated rearrangement of the fibres in the bend. As a result, number of bent GFRP
bars is increased and even doubled at such locations where bar bents are required. In this
study, GFRP bars with headed ends are used as straight bars at the inside face of the barrier
walls with an end head at the bottom to reduce their development length in the deck slab,
avoiding the use of hooks. This headed end is made of a thermo-setting polymeric concrete

418

with a compressive strength far greater than that of normal grade concrete. It is cast onto
the end of the straight bar and hardened at elevated temperatures. The concrete mix
contains an alkali resistant Vinyl Ester resin, the same material used in the straight bars,
and a mixture of fine aggregates. The maximum outer diameter of the end heads is 2.5
times the diameter of the bar. The head of the 16 mm bar is approximately 100 mm long. It
begins with a wide disk which transfers a large portion of the load from the bar into the
concrete. Beyond this disk, the head tapers in five steps to the outer diameter of the blank
bar. This geometry ensures optimal anchorage forces and minimal transverse splitting
action in the vicinity of the head.
3. BACKGROUND OF THE DEVELOPED GFRP-REINFORCED BARRIER
The design process of bridge barrier walls is specified in the Canadian Highway Bridge
Design Code, CHBC (CSA, 2006) CHBDC Clause 12.4.3.5 specifies that the suitability of
a traffic barrier anchorage to the deck slab shall be based on its performance during crash
testing of the traffic barrier. For an anchorage to be considered acceptable, significant
damage shall not occur in the anchorage or deck during crash testing. It also specifies that if
crash testing results for the anchorage are not available, the anchorage and deck shall be
designed to resist the maximum bending, shear and punching loads that can be transmitted
to them by the barrier wall. As such, the initial design of the proposed PL-3 precast bridge
barrier (Sennah et al., 2010) was carried out to meet the CHBDC design criteria specified
for static loading at the anchorage between the deck slab and the barrier wall. CHBDC
specifies transverse load of 210 kN that can be applied simultaneously over a 2.4 m barrier
length. For the anchorage resistance of the GFRP bars embedded in the deck slab, Pahn
(Pahn 2008) conducted pullout tests on 16 mm diameter GFRP bars provided with headed
ends to determine their pullout capacity when they are embedded in concrete over bond
lengths of 100 mm and 200 mm. The results from this testing formed the basis for the
developed PL-3 barrier-deck joint. As for the design of the vertical and horizontal
reinforcement in the barrier wall, the yield-line analysis was conducted (Sennah et al.,
2010) on the ultimate flexural capacity of the wall per AASHTO-LRFD Bridge Design
Specifications (AASHTO, 2004). Such design work for the PL-3 bridge barrier proposed
the use of 16 mm and 12 mm diameter GFRP bars as vertical reinforcement in the barrier
front and back faces, respectively, with 12 mm diameter GFRP bars as horizontal
reinforcement in case of PL-3 barrier wall, all at 300 mm spacing. The connection between
the deck slab and the barrier wall utilized the GFRP headed end bars for proper anchorage.
Figure 3 shows a schematic diagram of the GFRP reinforcement on the designed barrier
wall. Two full-scale PL-3 barrier models of 1200 mm length were erected and tested tocollapse to determine their ultimate load carrying capacities and failure models (Sennah et
al., 2010). The first barrier was a control one with reinforcing steel bars, while the second
barrier model was reinforced with GFRP bars with headed ends. Based on the experimental
findlings, it was concluded that GFRP bars with headed anchorage can be safely used in
bridge barrier walls to resist the applied vehicle impact load specified in CHBDC at the
barrier-deck slab anchorage. However, CHBDC Clause 12.4.3.4.4 specifies crash testing
for the design of the barrier wall itself (i.e. both vertical and horizontal reinforcement).

419

Fig. 3. PL-3 FRP-reinforced bridge barrier

Fig. 5. Close-up view


of barrier anchorage
to deck slab
4.

Fig. 4. View of the GFRP reinforcement

Fig. 6. View of the


barrier wall before
crash testing

Fig. 7. View of the


test vehicle

Fig. 8. View of the


vehicle before
impact

CRASH TESTING OF THE DEVELOPED GFRP-REINFORCED BARRIER

In November 2010, vehicle crash test was conducted at Texas Transportation Institute in
collaboration with Ryerson University and Schoeck Canada Inc. The crash test was
performed in accordance with Test Level 5 (TL-5) of MASH (MASH, 2009), which
involves the 36000V van-type tractor trailer (cab-behind-engine model of 36,000 kg gross
weight) impacting the barrier at a nominal speed and angle of 80 km/h and 15 degrees,
respectively. Figures 4 and 5 show views of the GFRP reinforcement before making the
timber forms and casting concrete. While Fig. 6 shows view of the built 40-m long barrier
before vehicle impact. Figure 7 shows view of the test vehicle, while Fig. 8 shows the test
vehicle during a mock test before impact.
5.

CRASH TEST RESULTS

On test day, concrete cylinders were tested to-collapse to determine their compressive
strength. The resulting concrete characteristic compressive strength was 32 MPa.
At the time of the test, the tractor trailer was guided into the barrier using a remote control
steering system. The tractor trailer impacted the barrier at 620 mm upstream of the control
joint located at 10.8 m from the barrier downstream end. At 0.100 s, the cab of the test
vehicle began to redirect, and at 0.203 s, the lower right front corner of the van-trailer
contacted near the top of the barrier. At 0.403 s, the cab of the test vehicle was traveling

420

parallel with the barrier at a speed of 79.7 km/h. The van-trailer began traveling parallel
with the barrier at 0.667 s, and was traveling at a speed of 76.3 km/h. At 0.695 s, the lower
right rear corner of the van-trailer contacted near the top of the barrier, and at 0.748 s, the
right rear edge of the van-trailer ruptured. As the test vehicle continued along the barrier, it
righted itself and rode off the end of the barrier wall. The brakes on the test vehicle were
not applied, and the test vehicle subsequently came to rest 35.66 m downstream of the end
of the barrier and 2.7 m toward the field side. Sequential photographs for the crash test are
presented in Figs. 9 and 10 for frontal and side views, respectively.

(a)

(b)

(c)

(d)

(e)

(f)

(g)

(h)

(i)

(a)

(e)

(j)
(k)
(l)
Fig. 9. Sequential photographs for the crash test (frontal views)

(b)

(c)

(d)

(f)
(g)
(h)
Fig. 10. Sequential photographs for the crash test (side views)

421

(i)
(j)
(k)
(l)
Fig. 10. Sequential photographs for the crash test (side views) (continued)
Evaluation criteria for full-scale vehicle crash testing are based on three appraisal areas,
namely: (i) structural adequacy; (ii) occupant risk; and (iii) vehicle trajectory after collision.
Structural adequacy is judged upon the ability of the barrier to contain and redirect the
vehicle, or bring the vehicle to a controlled stop in a predictable manner. The vehicle
should not penetrate or override the barrier although lateral deflection of barrier is
acceptable. Occupant risk criteria evaluate (i) the potential risk of hazard to occupants in
the impacting vehicle and to some extend other traffic, pedestrian, or workers in
construction zones, if applicable; (ii) deformation of, or intrusions into, the occupant
compartment should not exceed preset limits set forth in MASH; and (iii) whether the
vehicle remain upright during and after collision. Post impact vehicle trajectory is assessed
to determine potential for secondary impact with other vehicles or fixed objects, creating
further risk of injury to occupants of the impacting vehicle and/or risk of injury to
occupants in other vehicles. Crash test results showed that the barrier contained and
redirected the 36000V vehicle. The vehicle did not penetrate or override the barrier. No
detached elements, fragments, or other debris from the barrier were present to penetrate or
show potential for penetrating the occupant compartment, or to present undue hazard to
others in the area. No occupant compartment deformation occurred. The vehicle remained
upright during and after the collision event (Buth and Menges, 2011; Sennah, 2011).

Fig. 11. General view of the barrier wall


after vehicle impact

Fig. 12. View of the barrier wall and the test


vehicle at rest

After the crash test, minor cracks in the front and back side of the barrier were observed.
However, no complete damage was observed. Figures 11 and 12 show views of the barrier
reflecting this finding. In practice, these minor cracks may need to be repaired to avoid
possible crack propagation resulting from other possible vehicle impact. It should be noted
that the crack pattern at the location of impact is in the shape of punching shear as depicted
in Figs. 13 through 15. In addition, Fig. 15 did not show any sign of vertical flexural crack

422

similar to that in the yield line failure pattern specified in AASHTO-LRFD Specifications.
In February 2011, Ryerson University research team conducted static load failure tests on
the barrier segments to provide research information that will be used further to evaluate
the applicability of AASHTO-LRFD yield-line equations, developed for reinforcing steel
bars, to the design of GFRP-reinforced barrier under equivalent static vehicle impact
loading. It is interesting to mention that all failure modes resulting from the static load tests
were in the form of punching shear type rather that the flexural yield-line pattern specified
in ASSHTO-LRFD Specifications. More information about this comprehensive research
program can be found elsewhere (Sennah, 2011).

Fig. 13. Views of the crack pattern, at the control joint, on the front (inner) side of the
barrier wall

Fig. 14. View of the punching


shear crack at the location of
impact of the rear end of the
vehicle

Fig. 15. View of the back (exterior) face of the barrier


at the control joint (location of vehicle impact) showing
signs of initiation of punching shear cracks

423

6. CONCLUSIONS
A vehicle crash test was conducted on a newly developed GFRP-reinforced PL-3 bridge
barrier system. Results from tests qualified such innovative barrier system to resist vehicle
impact per MASH crash test requirement. Crash test results showed that the developed
barrier contained and redirected the vehicle. The vehicle did not penetrate or override the
parapet. No detached elements, fragments, or other debris from the barrier were present to
penetrate or show potential for penetrating the occupant compartment, or to present undue
hazard to others in the area. No occupant compartment deformation occurred. The test
vehicle remained upright during and after the collision event. The recorded crack pattern
after crash testing shows signs of punching shear cracks. This is in contrast to the specified
vertical flexural crack at the back face of the barrier wall and the two diagonal cracks at the
front face of the barrier that formed the basis for the AASHTO-LRFD design yield-line
failure equations.
7. ACKNOWLEDGEMENTS
The authors acknowledge the support to this project by Schoeck Canada Inc. of Kitchener,
Ontario, Canada. Also, the authors would like to acknowledge Texas Transportation
Institute for conducting the crash test in collaboration with Ryerson University. Moreover,
the authors would like to thank Mr. Nidal Jaalouk, Lead Technical Officer at Ryerson
University, for instrumenting the barrier wall prior to crash testing.
8. REFERENCES
1. AASHTO. 2004. AASHTO-LRFD Bridge Design Specifications. Third Edition,
American Association of State Highway and Transportation Officials, Washington, DC.
2. Buth, C.E., and Menges, W.L. 2011. MASH Test 5-12 of the Schoeck ComBAR
Parapet. Draft Report submitted to Schoeck Canada Inc., 55 p.
3. Canadian Standard Association (CSA). 2006. Canadian Highway Bridge Design Code.
CAN/CSA-S6-06, Toronto, Ontario, Canada.
4. MASH. 2009. Manual for Assessing Safety Hardware, MASH. American Association
of State Highway and Transportation Officials, Washington. DC.
5. Pahn, M. 2008. Monitoring of Bond Test and Pull-out Tests for ComBar 16 Provided
with Headed Stud Connectors. Technical Report for Schck Bauteile GmbH,
Technische Universitt Kaiserslautern, Germany, pp. 1-17.
6. RCCAO. 2007. Ontarios Bridges Bridging the Gap. Report prepared by MMM Group
for Residential and Construction Alliance of Ontario, RCCAO, 56 p.
7. Schck Canada Inc. 2010. www.Schoeck-canada.com.
8. Sennah, K. 2011. Crashworthiness of GFRP-Reinforced PL-3 Concrete Bridge Barrier.
Technical Report submitted to Schoeck Canada Inc.
9. Sennah, K., Tropynina, E., Goremykin, S., Lam, M., and Lucic, S. 2010. Concrete
Bridge Barriers Reinforced with GFRP Bars with Headed Ends. Proceedings of the 8th
International Conference on Short and Medium Span Bridges, Niagara Falls, Ontario,
pp. 1-10.

424

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

EXPERIMENTAL SETUP AND THERMAL INSULATION


ASSESSMENT OF SANDWICH CONCRETE WALLS WITH GFRP
CONNECTORS FOR SUSTAINABLE CONSTRUCTION
G. Woltman1, M. Hanna2, D.G. Tomlinson3 and A. Fam4
1

Masters Candidate, Dept of Civil Engineering, Queen's University, Kingston, Canada


Research Associate, Dept of Civil Engineering, Queen's University, Kingston, Canada
3
Doctoral Student, Dept of Civil Engineering, Queen's University, Kingston, Canada
4
Professor and Canada Research Chair, Dept of Civil Eng., Queen's University, Kingston, Canada
2

ABSTRACT
Precast insulated concrete sandwich panels have great potential in the residential
construction industry. They can be constructed under controlled conditions, shipped easily
to site, and erected rapidly to close the building envelope. This paper addresses an ongoing experimental program geared towards assessment of thermal response of a sandwich
panel system. The panel consists of two exterior concrete wythes, an interior layer of
concrete studs, and an insulating foam with high thermal resistance. As the concrete studs
are not in contact with the wythes, Glass Fibre Reinforced Polymer (GFRP) connectors are
used to bind the layers together. The experimental evaluation involves thermal testing in a
hot-box apparatus constructed as part of the project. The objectives are to minimize
thermal bridging and study the impact of various GFRP dowel sizes and spacing on the
overall R-values of the sandwich panels. The apparatus was successfully calibrated and
showed a level of uncertainty (13%) consistent with the expected error of an ASTM 1363
test apparatus. Preliminary analysis of selected test data of walls incorporating GFRP
connectors show estimated R-values between 4 and 4.4 m2K/W.
1. INTRODUCTION
While precast concrete sandwich panels have been used reliably in building envelopes for
over a half century, updated building codes which place an increased emphasis on energy
efficiency have made this system increasingly attractive in modern construction.
Traditionally concrete-insulation-concrete sandwich panels consist of two concrete wythes,
separated by a layer of insulation and connected by either regions of solid concrete (ribs) or
metallic ties [1]. These connectors are critical to the panel structure, especially if
composite action between the wythes is desired. However, they create a thermal bridge
across the insulation layer and allow a significant degree of thermal energy transfer through
the panel. McCall (1985) [2] has determined that the thermal bridging created by either

425

concrete or steel connectors can reduce the thermal performance of a sandwich panel by as
much as 40%
Limited efforts have been made to reduce these thermal bridges, including the introduction
of Glass-Fibre Reinforced Polymer (GFRP) connectors, which have significantly lower
thermal conductivity than metallic connectors [3], or lengthening the thermal bridge paths
by using three-wythe panels [4,5]. Experimental evaluation of the thermal properties of
these systems is generally confined to private research, funded by GFRP connector
producers. One possible reason for this is the cost of experimental testing, which in
addition to specimen fabrication requires a significant apparatus and instrumentation.
Increasing the knowledge and prediction of thermal properties of sandwich panels is an
important research focus at present, as the latest building codes contain significant
additional requirements for building energy efficiency. The 2010 edition of the ASHRAE
90.1 standard is targeting 30% energy efficiency savings versus the 2007 edition [6]. In
Canada, many provinces have now codified stringent building envelope insulation
requirements, and the adoption of energy efficiency as the fifth core objective of the
National Building Code of Canada is expected shortly. The thermal properties of sandwich
panels, as well as their many other advantages as building systems, makes them an
attractive option in present and future constructions.
2. EXPERIMENTAL STUDY
2.1 Panel Design
The new sandwich panel system consists of two 50 mm thick concrete wythes, separated by
150 mm of rigid extruded polystyrene (XPS) foam insulation. Within the insulation layer, a
system of isolated concrete studs encapsulate the GFRP connectors for an enhanced
stiffening and reduction of their transverse flexibility, towards improving composite action
of the overall system. However, because the concrete studs are isolated (i.e. are not in
direct contact with the concrete wythes), their thermal bridging effect is reduced. The
overall dimensions of the test panels are 2400 mm height, 1550 mm width, and 250 mm
thickness. As indicated earlier, to bind the system and provide some composite action,
GFRP shear connectors are placed in the studs, projecting through the insulation layer into
each exterior wythe. This structural system can be seen in Figure 1. In addition, one panel
used a conventional galvanized steel connector to compare with the GFRP connectors. The
GFRP and steel connectors are shown in Figure 2.

426

Fig. 1. Sandwich panel cross section view showing various components.

Fig. 2. Panel shear connectors. A: GFRP connector. B: Galvanized Steel Connector.


2.2 Test Parameters
This experimental study involved testing ten full-scale specimens in a hot box apparatus
under steady-state thermal conditions to evaluate their thermal performance. In addition,
the testing was designed to obtain experimental values of surface and internal temperatures
which could be correlated with analytical results. Five parameters were evaluated during
these tests for their effect on thermal conductivity: connector size, connector spacing, total
connector area, concrete stud spacing, and connector material. The ten panel specimens
and their parameter details are shown in Table 1.

427

Table 1. Sandwich panel testing matrix showing various parameters.


Numbe
r
Shear
Number
Connecto
Panel
Connector
of
Connecto
of
r
Numbe
Parameter
Diameter
concret
r
Connector Spacing
r
(mm)
e
Material
s
(mm)
Studs
1
Baseline
4
GFRP
9.5
28
375
2
Loose Spacing
4
GFRP
9.5
16
750
3
Tight Spacing
4
GFRP
9.5
44
225
4
5-Studs
5
GFRP
9.5
35
375
5
3-Studs
3
GFRP
9.5
21
375
Large
6
Connector
4
GFRP
12.7
28
375
Small
7
Connector
4
GFRP
6.4
28
375
8
Steel Connector
4
Steel
4.6
28
375
Large, same
9
Area
4
GFRP
12.7
16
750
Small, same
10
Area
4
GFRP
6.4
64
150
2.3 Test Apparatus
Thermal testing of full scale panel specimens required an elaborate test apparatus. The
apparatus was a hot-box modeled after ASTM 1363. It consisted of two chambers, one was
heated up, while the other was cooled down; referred to as the metering and climate
chambers, respectively. The test specimen was placed between the two chambers. Careful
attention was given to minimizing thermal leakage both around the test specimen and
through the walls and door of the test apparatus. The interior and exterior conditions of the
test apparatus were monitored continuously throughout each test.
Due to the large size and mass of the test specimens, a robust structure was required for the
chambers, to hold the specimens in place. The structure consisted of two converted 1800 x
2400 mm box culverts, with a gap between the culverts for the test specimen to fit in
(Figures 3 and 4). The culverts were placed on a bed of 100 mm thick Dow SM insulation
over a layer of granular material, and their entire surface areas were covered with a 100 mm
thick layer of insulation on all sides, so that each apparatus surface had an insulation value
(R-value) of about 3.52 m2K/W. Gaps between rigid insulation panels were filled with
spray foam as necessary, and all seams were sealed with sheathing tape. As the test
specimens were slightly smaller than the opening between the box culverts, a surround
panel was constructed to prevent heat loss by flanking. This surround panel consisted of
250 mm of Dow SM insulation, with an R-value of 8.81 m2K/W.
To calibrate the apparatus, a test was performed using a 100 mm thick Dow insulation layer
specimen, with a reported manufacturers R-value of 3.52 m2K/W.

428

Cold Chamber

Cold
Side
door

Hot chamber

2000

2400

40
22

Hot
Side
Door

40
27

1400
1800

Fig. 3. Hot-box testing apparatus concrete box culvert dimensions.

Fig. 4. Cross section of thermal test showing equipment and materials required apparatus.
2.4 Instrumentation and Test Procedure
During testing, the air in the hot chamber of the apparatus was maintained at 24o C using an
800 W heating coil, with the current supplied to the heating coil monitored using an F.W.
Bell PC-50 current sensor. The cold side of the test apparatus was maintained at -5o C
using a walk-in freezer compressor and evaporator coil. Current supplied to the freezer was
also monitored to keep track of defrost cycles and ensure that testing was conducted under
steady-state conditions. The location of these can be seen in Figure 4.
429

Monitoring and recording of apparatus and test specimen data was achieved using a 24
channel OMEGA-320 data logger. A total of 22 thermocouples were used to monitor
temperatures on the surface of the panel, within the panel thickness, and on the surfaces of
the apparatus walls. In addition, air temperature, humidity, and velocity were monitored
inside both apparatus chambers and outside the apparatus structure.
3. TEST RESULTS AND DISCUSSION
3.1 Preliminary Calibration Results
In order to calculate the heat losses through the apparatus and determine the uncertainty of
the results, the apparatus was calibrated using a 100 mm thick specimen of XPS insulation,
with the same rectangular dimensions as the test specimens. The manufacturers provided
R-value was taken as the thermal resistance of the calibration specimen. The losses through
each wall of the hot chamber were determined using the surface temperatures of the walls,
and an estimated R-value for the concrete and insulation of each wall. The total heat
supplied to the chamber was calculated using the current supplied to the heating coil and
the measured resistance of the heater. These values were then compared to the theoretical
values calculated using heat transfer equations. A sample of these preliminary results is
shown in Table 2. The error in the test apparatus was found to be 13.6 %, which is
consistent with the expected error of an ASTM 1363 test apparatus. The R-value of the
tested foam is higher than the documented one by the 13.5%. This error is excepted due to
many factors including the conditions used by manufacturer to establish their reported Rvalue, which is likely slightly different from the test cell conditions in this study. In
addition, the complicity of the shape of the test cell used in this study creates some minor
uncertainty in the amount of the heat loss from the hot chamber.
Component
of chamber
1
2
3
4
5
6
7

Table 2. Thermal losses with calibration panel.


Rtotal
A
Tin
Description
2
2
m K/W
m
deg
North
3.7070
6.2554 24.4000
South
3.7070
6.2554 24.4000
Door
4.0700
4.2400 24.4000
Floor
3.6775
4.6114 24.4000
Ceiling
3.7070
4.6110 24.4000
Surrounded(foam)
8.8619
2.3981 24.4000
corner loss
N/A
N/A
24.4000

Tout
deg
15.0000
15.0000
15.0000
18.0000
15.0000
-5.5000
15.0000

Qloss
W
15.8621
15.8621
9.7926
8.0253
11.6923
8.0912
22.2321
91.5578

3.2 Preliminary Test Results


Testing of the ten panel specimens was very recently completed, and preliminary results
show that the experimental thermal resistance of the panels is consistent with analytical
calculations. The system was able to reach steady-state conditions for temperature and
current, and consistent internal and surface temperatures were obtained for each panel. Of

430

Temperature
(oC)

particular interest are the values obtained for internal temperatures, as these can be
correlated with an analytical modeling, currently being developed, to check the models
validity. Internal temperatures for each panel were determined along three section paths:
through the GFRP connector, through a concrete stud, and through a solid insulation region.
Thermocouples were located at the panel surfaces, at the interior of the concrete wythes, at
the stud surfaces, and in the middle of the solid insulation. Internal temperatures for Panel
1 are shown in Figure 5.
30
25
20
15
10
5
0
-5
-10

Through
Connector
Through
Stud
Through
Solid
Insulation

Hot Wythe Surface


Hot Insulation Surface
Hot Stud Surface Middle Cold Stud Surface
Cold InsulationCold
Surface
Wythe Surface

Position

Fig. 5. Temperature distribution through Panel 1 (9.5mm GFRP connectors @ 375mm).


While initial results suggest that the various GFRP diameters had only a small impact upon
the overall thermal conductivity, the use of metallic connectors in Panel 8 significantly
increased the thermal bridging effect, as expected due to the high thermal conductivity of
steel. Preliminary results of three analyzed panels with GFRP connectors show that their
R-values are between 4.04 and 4.43 m2K/W. These values are slightly less than those
calculated from a one-dimensional heat transfer model, 5 m2K/W. The differences for these
three experimental results to the analytical values are between 12 and 20%.
4. CONCLUSIONS
This project aimed at developing and validating thermal response of new sandwich panels
employing GFRP materials, towards achieving a sustainable and energy efficient
construction of buildings. The following preliminary conclusions can be drawn from the
experimental program:
1. An elaborate large scale hot-box was successfully constructed to accommodate a
new full scale (1550x2400x250 mm) concrete sandwich panels with GFRP shear
connectors used to minimize thermal bridging. The test apparatus was successfully
calibrated and showed a level of uncertainty (13%) consistent with the expected
error of an ASTM 1363 test apparatus.
2. Preliminary analysis of test data of three walls with GFRP connectors show R-value
of the new wall between 4.04 and 4.43 m2K/W. These values are slightly less than
those calculated from a one-dimensional heat transfer theoretical model, 5 m2K/W.
Work is in progress to continue interpretation of the thermal test data, including
confirmation of the heat balance of the test cell. Also, a two dimensional finite difference
431

model has been developed to simulate thermally the new panel in more details. The Rvalue and the temperature distribution inside the panel are in progress too.
5. ACKNOWLEDGEMENT
The Authors wish to acknowledge the financial and in-kind support of the Ontario Centres
of Excellence (OCE) and Anchor Concrete Products Ltd.
6. REFERENCES
1.
2.
3.
4.
5.
6.

PCI Committee on Precast Sandwich Wall Panels. 1997. State-of-the-Art of


Precast/Prestressed Sandwich Wall Panels. PCI Journal, 42(2): 92-134.
McCall, W.C. 1985. Thermal Properties of Sandwich Panels. Concrete International,
7 (1): 35-41.
Einea, A., Tadros, M.K., Salmon, D.C., and Culp, T.D. 1994. A New Structurally and
Thermally Efficient Sandwich Panel System. PCI Journal, 39(4): 90-101.
Lee, B.J. and Pessiki, S. 2004. Analytical Investigation of Thermal Performance of
Three-Wythe Sandwich Wall Panels. PCI Journal, 49(4): 88-101.
Lee, B.J. and Pessiki, S. 2008. Revised Zone Method R-value Calculation for Precast
Concrete Sandwich Panels Containing Metal Wythe Connectors. PCI Journal, 53(5):
86-100.
Hunn, B.D. 2010. 35 Years of Standard 90.1. ASHRAE Journal, 52(3): 36-46.

432

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

FRP COMPOSITES FOR STRENGTHENING BRIDGES IN


VICTORIA
Y.C. Koay
VicRoads, Australia

ABSTRACT
FRP composites are commonly used in the aerospace, marine and defence industries. Due
to their inherent advantages, FRP composites are increasingly making their way into civil
engineering applications. In the field of bridge engineering, FRP composites are used
mainly for strengthening and rehabilitation purposes. To date the number of primary
structural applications of polymer composite materials in Victoria bridge construction
remains relatively low. However, FRP composites are thinner, lighter, and non-corrosive,
exhibit much higher tensile strength and are much easier and quicker to install without the
need for expensive and disruptive temporary support systems. FRP are much more
appropriate for curved applications and can be applied over longer lengths without
requiring overlapping or connections. This paper presents the different forms of FRP and
discusses the specification requirements and acceptance criteria for the strengthening of
concrete structures using FRP strengthening systems.
1. INTRODUCTION
Approximately 70% of the bridges in Victoria were built before 1975. They were designed
to carry much lower loads than todays bridges and this capacity is being further reduced by
deterioration in their condition. A high proportion of these bridges require strengthening
and rehabilitation and, ideally, replacement in order to meet the ever increasing traffic
volumes and heavier commercial vehicle loads.
Conventional strengthening methods include the addition of external reinforcement using
steel plates or additional prestressing. FRP composites systems are showing potential as
alternative ways of strengthening and maintaining bridges. In the last 20 years, some
overseas concrete and steel bridges have been strengthened or have had major remedial
repairs carried out using FRP composites. Also, a number of new bridge superstructures
using FRP composites have also been built.
Australia however there have been few applications of FRP composites in bridge works to
date and the main use of composites has been in building projects. Although Victoria has

433

already made use of FRP composites in some key projects such as the strengthening of
West Gate Bridge and has taken some steps towards the use of FRP composites, the use of
this class of material must be increased significantly in order to gain the maximum benefits.
In order to make this step-change, the relevant State Road Authority must increase its
knowledge of the materials and the design and other specification issues that are necessary
for full scale implementation.
At this point in time, design guidelines and codes to which design engineers and contractors
can refer are not available in Australia. Much progress is however being made in countries
like Canada, USA, Japan and Europe. Because of limited experience in the use of FRP
composites in the civil infrastructure area meaningful comparisons of performance and cost
with conventional systems and materials are difficult. There is a need to research how FRP
composites may perform in Australian conditions and to develop design guidelines and
codes of practice.
2. MATERIALS FOR FIBRE REINFORCED POLYMER COMPOSITE
STRENGTHENING
The important fibres used in strengthening materials for external repair and strengthening
are carbon, aramid and glass. Their selection for particular applications depends on many
factors including their individual properties, the type of component to be strengthened,
loading history, temperature and moisture conditions and other environmental effects. In
selecting the type of fibre to be used, consideration should be given to such factors as
chemical resistance, compressive strength, elastic modulus, and other factors such as
impact resistance, resistance to ultraviolet light, fire resistance, and electrical conductivity
etc.
There are numerous FRP systems commercially available from various manufacturers. The
properties of the FRP materials offered also vary from one manufacture to another. It is
important that designers and users have the most recent technical data of the FRP system
from the manufacturers. The FRP material properties reported by the manufacturers are
minimum values achievable and these values should be verified by independent testing
whenever necessary. The properties of several typical commercially available proprietary
FRP systems are summarised in Table 1.
The performance of FRP systems can be greatly influenced by the volume, type and
orientation of fibre used, quality control during manufacture, the type of resin used and
dimensional effects. At this stage no standard design code is available for FRP composite
systems in Australia. FRP composites are available in a number of forms as follows:
2.1 Poltruded FRP Plates
Poltruded FRP plates are generally made with continuous carbon fibres taken through a
resin impregnation bath and a heated die shaping process, which consolidates the
combination and cures the resin. Plates formed by poltrusion are produced with high fibre
content, have a uniform cross-sectional shape and are of a consistent quality. They are
generally 1mm to 2mm thick, supplied in widths generally in the range of 50mm to 100mm
and can easily be cut to length on site.
434

Table 1. Properties of commercial FRP systems


Tensile
Tensile
Weight Thickness
FRP System
Strength
Modulus
(g/m)
(mm)
(MPa)
(GPa)
MBrace
MBrace EG900
915
0.353
1,730
88
MBrace CF130
300
0.165
3,480
227
MBrace CF530
300
0.165
2,940
372
Replark (Mitsubishi)
Grade 20
1.8
0.11
3,400
230
Grade 30
1.8
0.17
3,400
230
Grade MM
1.8
0.17
2,900
390
Grade HM
2.1
0.14
1,900
640
Sika
Sika Wrap Hex
920
1.0
600
26.1
100G
Sike Wrap Hex
610
1.0
960
70
103C
Sika CarboDur S
2,240
1.2
2,800
165
Sika CarboDur M
2,240
1.4
2,400
210
Sika CarboDur H
2,240
1.4
1,300
300

Elongation at
failure (%)
2.0
1.5
0.8
1.2
1.2
0.7
0.3
2.2
1.3
1.7
1.2
0.45

2.2 Poltruded FRP Strips or Sheet Material


Poltruded FRP strips or sheet material are made of either carbon or aramid fibres in a resin
matrix. These are in the form of either parallel unidirectional fibres or woven rovings for
two directional properties on a removable backing sheet. The strips can be as low as 0.1mm
thick and up to 500mm wide and can be provided in the form of a coil, which can be easily
cut to length.
2.3 Reinforcing Fibres (Sheet or Roll Form)
Reinforcing fibres (either carbon or aramid) in sheet or roll form are pre-impregnated with
a resin (prepreg process) and stored for use and are heat cured once in place. These have
relatively high fibre content in the longitudinal direction and may incorporate some glass
fibres diagonally to improve the handling of the material. The rolls can be up to 12m long
and thicknesses up to 30mm can be achieved.
2.4

Fabrics or Flexible Sheets

Fabrics or flexible sheets (dry or with a small amount of factory applied resin) can be used
in wet lay-up FRP systems to form thick laminate applications. On site the dry fabric/sheets
are mechanically or manually saturated with resin and placed by hand to form a composite
laminate. This is a very adaptable technique and can be used for strengthening of columns
and completely surrounding the sides and soffits of beams. However, it is very labour
intensive and requires experienced personnel to undertake the work. Proper orientation of
435

the fabric/sheet, adequate saturation with resin, removal of air pockets and initial
pretensioning of the material are critical to the in-situ performance of the system.
3. FRP COMPOSITE SYSTEM MATERIALS AND GENERAL PROPERTY
REQUIREMENTS
The requirements for the supply and quality of materials, surface preparation, installation,
relevant inspection and testing and acceptance criteria for the strengthening of concrete
structures using FRP composites must always be specified in a very clear and succinct way
to avoid any misinterpretations which may affect the long term performance of the product
in place. The application of any antigraffiti or decorative/ anticarbonation coatings as part
of the FRP composite strengthening system installation work must also be specified as
necessary.
It should be a specification requirement to ensure that the FRP composites strengthening
system is supported with information including the proposed resins, primers, patties,
saturants, adhesives and reinforcing fibres, substrate preparation, method of application,
equipment and proposed experienced operators, prior to the commencement of the FRP
composite strengthening works. It is good practice to thoroughly review and approve this
vital information including proposed work method statements or procedures prior to
commencement. Other supporting documentation should include relevant test certificates
which demonstrate conformance of the various materials to acceptable national or
international standard test methods, and if possible, evidence of previous performance. In
addition to any specified requirements, FRP composite materials and systems should
always address the material manufacturers recommendations and material safety data
sheets.
Strengthening works should be designed to upgrade the structural capacity and performance
of particular concrete components to recognised design standards as specified in AS 5100
(2004) i.e. to T44 or SM1600 etc. The FRP composite materials used for strengthening
works such as for bridge works should offer the following characteristics:
3.1 Carbon Fibre Laminate (Carbon Fibre Reinforced with Epoxy Matrix)
The carbon fibre laminate material should generally be a pre-fabricated, poltruded section,
specifically designed for adding tensile strength as part of a compatible, load transferring,
bonded system. The carbon fibre laminate should be available in a range of moduli and
strength grades, widths and thicknesses and possess minimum elastic modulus, tensile
strength, elongation at break, minimum volumetric fibre fraction and a good resistance to
very high temperatures.
3.2

Carbon Fibre Fabric (High Strength Carbon Fibre)

The fibre fabric materials should be pre-woven into sheets, specifically designed for adding
strength as part of a compatible, load transferring, bonded system. The fibre fabric should
possess minimum elastic modulus, tensile strength, elongation at break, minimum thickness
for static design and a minimum fibre density.
436

3.3

Adhesive for Carbon Fibre Laminate

The adhesives must be a thixotropic paste used to bond procured FRP composite laminate
systems to the concrete substrate and provide the required shear load path between the
concrete substrate and the FRP composite reinforcing laminate. Adhesives should be used
to bond together multiple layers of FRP composite laminates where required. The adhesive
for the carbon fibre laminate should possess minimum adhesive strength or result in
concrete failure, shear strength, compressive strength, tensile strength, a minimum static EModulus and coefficient of thermal expansion, and have the ability to be built up to 20mm
thickness without additional filler.
3.4

Saturating Resin

The saturating resin should be used to impregnate the reinforcing fibre fabric to fix it in
place and must be capable of providing a shear load path to effectively transfer the load
between fibres. The saturating resin should also be able to serve as the adhesive for wet layup systems, and be capable of providing a shear load path between the previously primed
concrete substrate and the FRP composite system. The saturating resin should possess
minimum adhesive strength, tensile strength, compressive strength, flexural strength, a
minimum flexural E-Modulus and relatively low viscosity. The application temperature
should be between 5C and 35C (ambient and substrate).
3.5

General Resin Requirements

Resins used as part of FRP composite strengthening systems including primers, putty
fillers, saturants and adhesives should offer, good compatibility with and adhesion to the
concrete substrate; good compatibility with and adhesion to the FRP composite system;
good resistance to in-service environmental effects, including but not be limited to
moisture, salt water, temperature extremes and chemicals normally associated with exposed
concrete; good filling ability and workability; compatibility with and adhesion to the
reinforcing fibre; development of appropriate mechanical properties for the FRP composite;
and a resin pot life consistent with the application. Any primers used should be a very low
viscosity resin used to penetrate the concrete surface and provide an improved adhesive
bond for the saturating resin or adhesive. The leveling putty filler should be a thixotropic
paste used to fill small voids, including bug holes, in the concrete substrate, to provide a
smooth surface to which the FRP composite system bonds and also prevent bubbles from
forming during curing of the saturating resin. The leveling putty filler and primers should
have the similar properties to the adhesive for carbon fibre laminate and saturating resins
respectively.
3.6

Durability of FRP Composites

The technical literature suggests 30 years as the design life of fibre composite materials.
However, the overall long-term performance of FRP systems may be affected by a number
of environmental factors, including temperature, humidity, moisture effects and chemical
exposure. Such factors would have an adverse effect on both the FRP fibres and the
437

adhesives depending on the resin type and formulation. Due to the lack of long-term
experience of the performance of fibre composite strengthening systems, regular inspection,
monitoring and maintenance regimes should be put in place.
4. DESIGN GUIDELINES
The use of externally bonded FRP to strengthen and rehabilitate existing concrete structures
is a relatively new challenge. There is a lack of design code and guidelines for their use.
ACI Committee 440 (2002), ISIS Canada (2008) and fib (2001) each have developed and
issued separate design manuals on the use of FRP for strengthening concrete structures.
Design codes have also been proposed by Teng et al. (2002).
The selection of suitable FRP system depends on numerous factors including
environmental, loading, durability and protective coating considerations. Environmental
considerations concern effects such as alkalinity, acidity, temperature and electrical
conductivity. For loading considerations, impact tolerance, creep rapture, fatigue and FRP
strength and stiffness are of concern. Durability concerns hot-wet cycling, alkaline
immersion, freeze-thaw cycling and ultraviolet exposure. Protective coatings are relied
upon to impede the degradation of the mechanical properties of the FRP systems. In
bridgeworks, strengthening are generally in the areas of flexure, shear or axial or a
combination of these.
4.1 Flexural strengthening
The design concept for flexural strengthening with FRP is essentially an extension of
existing flexural strength theory with appropriate limit checks to account for possible FRP
induced failure modes. In ACI Committee 440 (2002) and ISIS (2001), the following
assumptions are made in calculating the flexural resistance of the strengthened section:
plane section remains plane after loading
no slip or perfect bond between the FRP and concrete
shear deformation in the adhesive layer is neglected
maximum concrete compressive strain of 0.003 in ACI Committee 440 (2002) and
0.0035 for structures and 0.003 for bridges in ISIS Canada (2008)
tensile strength of concrete is neglected and
FRP reinforcement has a linear elastic stress strain behaviour up to failure.
An additional assumption that may be adopted is disregarding the initial strains in the
concrete and reinforcing steels. In most cases, the effect of these strains on the ultimate
flexural strength of the FRP strengthened beam is negligible. Nevertheless, these strains
should be considered and can be determined from an elastic analysis of the existing
member. In general, flexural strengthening involves bonding the FRP reinforcement to the
tension side of the structural member, which is the soffit of the beam shown in Figure 1(a).
A variation to this arrangement is to bond the FRP reinforcement on the web along the
length of the beam as illustrated in Figure 1(b) (Oehler 1990).

438

Fig. 1. Variation of flexural strengthening


The failure modes that can occur for an FRP strengthened section is flexure as reported by
ACI Committee 440 (2002) are given below:
crushing of the concrete in compression prior to yielding of the flexural reinforcing
steel;
yielding of the flexural reinforcement followed by FRP rupture;
yielding of the flexural reinforcement followed by concrete crushing;
shear / tension delamination of the concrete cover (cover delamination); and
FRP debonding from the concrete substrate
The strengthened section should be check against each case above if the failure mode
governing the design is difficult to determine. For obvious reasons, it should be noted that
the design of failure by FRP rupture before yielding of the flexural reinforcement is not
considered. The last two failure modes can be classified generally as debonding failures.
There are four types of debonding failure as illustrate in Figure 2. Failure modes (a) and (b)
are commonly referred to as plate-end debonding failure, while failure modes (c) and (d)
are commonly referred to as intermediate crack-induced interfacial debonding failure (Teng
et al., 2002). It is generally believed that plate end debonding failures are initiated by high
interfacial shear and normal stresses near the end of the plate. Intermediate crack-induced
interfacial debonding failures are initiated by widening and differential displacement of
crack surfaces causing high local interfaces stresses between the FRP and concrete near the
crack.

439

Fig. 3. Typical shear strengthening


arrangements

Fig. 2. Debonding failure modes of FRP


strengthened RC beam sections
4.2 Shear Strengthening

The existing shear capacity of a reinforced concrete section can be enhanced by bonding
FRP materials to the web, typically with the dominant fibre direction perpendicular to the
length of the member. There are a variety of FRP systems and arrangements reported in the
literature. The type of arrangement used strongly influences the contribution of the
externally bonded FRP to the shear load capacity. The terms used in this section are based
ont hose used in the study carried out by Al-Sulaimani et al. (1994), which provide accurate
description of the schemes used. In the literature, the terms plates, sheets and strips have
been used loosely to describe various forms of FRP. In this investigation, plates are taken to
mean pre-cured FRP systems typically with two or more layers of fibres. The pre-cured
FRP are bonded to the concrete surface with an adhesive specified by the manufacturer.
Sheets are taken. Sheets are taken to mean FRP fashioned using either the dry or wet lay-up
technique. Single or multiple layers of dry or uncured fabric or sheet are bonded to the
concrete surface to form the FRP. Strips refer to narrow pieces of either the plates or sheets
having a length to width ratio of more than two. The strips are placed at regular intervals
along the beams length. Plates are typically in strip form whereas sheets can be cut to
obtain individual strips. The various arrangements that may be used to shear strengthened
an existing member are illustrated in Figure 3.
5. CONCLUSIONS
Although FRP composite strengthening systems in Victoria are a relatively new technology
they can provide durability and long term performance, provided they are correctly chosen,
440

designed, specified and installed. The success of the FRP system installation is highly
dependent on the quality of the on site practices adopted and therefore work should be
undertaken by trained personnel who are competent in the installation procedures
developed by the manufacturer of the particular FRP system. Adequate inspection and
testing should be undertaken on a frequent basis as part of an overall quality assurance and
quality control program.
6. ACKNOWLEDGEMENTS
The author wishes to thank VicRoads for permission to publish this paper. The views of
this paper are those of the author and do not necessarily represent those of VicRoads.
7. REFERENCES
1. ACI Committee 440. 2002. Guide for the design and construction of externally bonded
FRP systems for strengthening concrete structures. American Concrete Institute (ACI),
Farmington Hill, MI.
2. Al-Sulaimani, G.J., Sharif, A.M., Basunbul, I.A., Baluch, M.H. and Ghaleb, B.N. 1994.
Shear repair for reinforced concrete by fibreglass plate bonding. ACI Structural Journal,
91(3): 458-464.
3. AS 5100. 2004. Bridge design, Australia Standards.
4. fib. 2001. Externally bonded FRP reinforcement for RC structures. The International
Federation for Structural Concrete (fib), Technical Report, Task Group 9.3, Bulletin No
14, Lausanne, Switzerland.
5. ISIS Canada. 2008. Strengthening reinforced concrete structures with externally bonded
fibre reinforced polymers. Design Manual No. 4, The Canadian Network of Centres of
Excellence on Intelligent Sensing for Innovative Structures, ISIS Canada, Canada.
6. Oehlers, D.J. and Moran, J.P. 1990. Premature failure of externally plated reinforced
concrete beams, Journal of Structural Engineering, ASCE, 116(4): 987-1003.
7. Teng, J.G., Chen, J.F., Smith, S.T. and Lam L. 2002. FRP Strengthened RC Structures,
John Wiley & Sons, UK.

441

442

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

AGEING EFFECTS ON THE DEBONDING BEHAVIOUR OF FRP


REINFORCEMENTS: EXPERIMENTS AND DAMAGE MODELING
K. Benzarti1, F. Freddi2, G. Ruocci1, M. Quiertant1 and S. Chataigner3
1

Universit Paris-Est / IFSTTAR, Paris, France


Universita di Parma, Dept. of Civil-Environmental Engineering, Parma, Italy
3
CETE de Lyon / DLA, Autun, France
2

ABSTRACT
This study investigates the influence of hydrothermal ageing on the debonding behaviour of
FRP-concrete interfaces subjected to a single lap shear test. In a first part, an experimental
campaign reveals a major change in the debonding process due to ageing at 40C and 95%
R.H., as the failure progressively evolves from a concrete failure towards a failure within
the adhesive joint. This phenomenon has been correlated to a substantial modification of
the load transfer mechanism along the joint, which itself originates from the large decrease
in the mechanical properties of the epoxy adhesive during ageing (plasticization). In a
second part, a damage model is used to simulate the debonding process and its evolution
during hydrothermal ageing. This model has been developed within the framework of the
principle of virtual power, and considers both concrete, FRP and adhesive as damageable
materials. Preliminary simulations show that such a model captures well the debonding
behaviour and is able to account for ageing-induced effects, provided physically relevant
evolutions of the damage parameters are introduced.
1. INTRODUCTION
Externally bonded fibre reinforced polymer (FRP) composites offer an effective and
versatile solution for strengthening concrete infrastructures in the framework of
rehabilitation or retrofit programs. Although this technique has become very popular
worldwide, the adhesive bond between concrete and FRP systems is known to be sensitive
to various environmental factors which may affect its long term performances. Among
others, humidity is suspected to be a major cause of degradation of the adhesive bond
properties and its effects on the fracture of FRP-concrete interfaces subjected to shear or
tensile loading have been pointed out in several studies 1-2.
To ensure the safety and reliability of FRP strengthened concrete members, design
procedures and predictive models that properly consider debonding problems are definitely
needed. Recently, an original model was proposed by Freddi et al. 3 to describe the
damage behaviour of bonded assemblies, which takes into account damages of both the
443

adherents and the adhesive interface, and interactions between the volume and surface
damages as well. This model is based upon the adaptation of the principle of virtual power
previously introduced by Nedjar and Frmond 4, assuming that damage results from
microscopic motions whose macroscopic effects are included in the power of the internal
forces. Besides, Freddi et al. 5 developed a simplified version of the model, whose
validity is restricted to the case of rate independent problems (quasi static fracture tests for
instance) and which involves a reduced set of theoretical parameters. This simplified model
was integrated into a finite element (FE) code (Deal II open source package 6), and
authors showed that it provides an accurate description of FRP-concrete debonding during a
single lap shear test conducted at low displacement rate (6m/s), and that it is able to
predict the fracture lines (in 2D) associated to specific geometries of the test body 5.
In this paper, it is proposed to implement the simplified damage model in order to predict
the influence of hydrothermal ageing on the fracture mode of FRP-concrete interfaces. In a
first part, experimental evidences of ageing effects are recalled, based on the main results of
a previous test campaign 2. Then, the basic equations of the model are briefly recalled,
and preliminary simulations of the debonding behaviour are reported for FRP-plated
specimens tested in shear before and after hydrothermal ageing. Numerical results are
finally discussed in the light of the experimental data.
2. EXPERIMENTAL INVESTIGATIONS
2.1. Test specimens and characterization methods
2.1.1. Test specimens
FRP strengthened specimens designed for single lap shear tests were prepared according to
the following stages:
concrete blocks of dimensions of 210 210 410 mm3 were cast from a standard
formulation based on CEM II 32.5 cement and silico-calcareous aggregates. After
28 days cure, the hardened material exhibited a compressive strength of 32 4
MPa, and a tensile strength of 3.6 0.1 MPa (as measured by tensile splitting tests)
a sanding treatment was then performed on the top face of the blocks, and a
commercial unidirectional carbon FRP (CFRP) plate of width 100mm and thickness
1.5mm was bonded to the concrete substrate using a bi-component epoxy adhesive;
the bonded length was 200mm and the adhesive layer started 50mm away from the
front side of the concrete blocks in order to prevent edge effects (Figure 1.a). Let us
underline that surface treatments of concrete and installation of FRP systems were
performed by experienced professionals from FRP suppliers.
Besides, dumbbell samples were also molded from the pure epoxy adhesive. They were
intended for tensile characterizations of the bulk polymer before and after ageing.
Due to a confidentiality agreement, the commercial denominations of FRP and adhesive
systems can not be released, but further details regarding the general properties and

444

chemical compositions of epoxy constituents are available in a previous paper (see Tables
2-3 and section 4 in reference [2]).
strain gauges
180
135

90

grips

50
15

50

20

210

70

410

(a)
(b)
Fig. 1. (a) Geometry of the FRP strengthened specimens (dimensions in mm), and details of
the gauge instrumentation and clamping conditions. (b) Picture of the shear test set-up.
2.1.2. Ageing conditions
Together, all the testing bodies (FRP strengthened blocks and bulk epoxy samples) were
exposed to ageing conditions at 40C and 95% relative humidity in a climatic chamber. The
temperature of 40C was chosen to accelerate the water sorption process, while remaining
below the glass transition temperature of the epoxy adhesive (51C). Samples were
periodically removed from the chamber in order to be mechanically tested.
2.1.3. Mechanical characterizations
Shear tests were performed using a home made testing machine (Figure 1.b). The test
consists in applying a tension to the bonded FRP plate by means of a hydraulic jack of
capacity 100 kN operating at a constant displacement rate of 6 m.s-1. In order to prevent
any rotation of the test specimens, adequate clamping conditions are chosen, as depicted in
Figure 1.a. Applied load and displacement of the grips are recorded continuously during the
test, and additional strain gauges bonded to the surface of the FRP material allow to build
the stain profile along the lap joint (positions of the gauges are reported in Figure 1.a). The
overall test procedure has been previously described in 2, 7-8].
2.2. Results of the test campaign
2.2.1. Evolution of the tensile properties of the bulk epoxy adhesive during ageing
As regards the property evolutions of the individual materials (concrete substrate, FRP
plate, epoxy adhesive) in the ageing environment, it was shown that only the bulk adhesive
exhibited remarkable changes in terms of tensile strength, strain at failure, and elastic
modulus 2]. No evolution was found for the FRP material and a slight increase in the
compressive strength of concrete was observed due to the ongoing hydration process.

445

Hydrothermal ageing leads to a large decrease in the tensile strength and an enhanced
ductility of the bulk epoxy (Table 1). Tensile strength and Youngs modulus are
respectively decreased by factors 3.8 and 4, and the value of the tensile strength becomes
even close to that of the concrete material (3.6 MPa). Such evolutions have been attributed
to an extensive plasticization of the polymer network resulting from water sorption 2].
Table 1. Results of the tensile tests for initial and aged samples of the bulk epoxy adhesive
Ageing time
0
3 months
8 months
Strength (MPa)
9.0 2
6.6 0.5
25 1
Strain at failure (%)
1.5 0.4
1.4 0.3
0.6 0.1
Elastic modulus (GPa)
4.9 0.2
1.3 0.2
1.2 0.2
2.2.2. Ageing effects on the debonding behaviour of CFRP plated specimens
Shear tests were performed on the CFRP strengthened specimens after various periods of
ageing, up to 20 months. No clear variation was found for the maximum shear capacities
during ageing, with values distributed in the range 40-55 kN 2. However large changes
were observed in both the shape of the load-displacement curves and the fracture mode of
the test bodies, as shown in Figures 2 and 3.
60.00
7 days ageing
9 months ageing

Applied load (kN)

50.00

40.00

30.00

20.00

10.00

0.00
0.00

0.20

0.40

0.60

0.80

1.00

1.20

1.40

Displacement (mm)

Fig. 2. Load versus displacement curves for CFRP plated specimens tested in the single lap
shear configuration, after ageing periods of 7 days and 9 months.

a)

b)

c)

Fig. 3. Evolution of the fracture patterns for CFRP plated specimens tested after various
ageing periods: 7 days (a), 9 months (b) and 12 months (c).
446

The debonded plate and the fractured substrate are respectively shown on the right and left
hand sides of each picture.

2500

2000

2000
Axial strain (x10 -6)

2500

-6

Axial strain (x10 )

As regards the load-displacement curves, unaged test bodies and specimens exposed for
short period of time (< 1 months) exhibit similar of behaviours, with an initial linear elastic
domain followed by a pseudo-plateau associated to the propagation of the concrete failure
along the interface. Specimens exposed for longer periods show a slight decrease in
stiffness and exhibit a kind of elasto-plastic behaviour; these effects are consistent with
evolutions previously reported for the bulk epoxy adhesive. Moreover, the length of the
pseudo-plateau becomes shorter and seems to vanish progressively. These features are in
agreement with the apparent evolution of the fracture mode from a cohesive substrate
failure towards a failure within the glue layer (Fig. 3).
Besides, the gauge instrumentation allowed to assess the strain profiles at the surface of the
FRP plates, as a function of the abscissa along the joint. Fig. 4 shows that ageing induces a
broad change in the load transfer mechanism along the lap joint, due to the drop in the
adhesive properties. In particular, an increase in the transfer length is observed.

1500
1000

1500
1000
500

500

a)

50

100

150

200

b)

Abscissa (in mm)

50

100

150

200

Abscissa (in mm)

Fig. 4. Strain profiles on the CFRP plate versus abscissa along the joint, for initial (a) and 9
months aged (b) samples. Each color corresponds to a given load level during the shear test.
3. MODELLING APPROACH
3.1. Brief description of the damage model
In this section will be recalled the main definitions and equations of the damage model
which will be used to simulate the debonding behaviour at the FRP-concrete interface. An
overall description of the mechanical and mathematical developments can be found in [3-5].
Let us consider a bonded assembly made of two domains i, i = 1,2, subjected to mixed
boundary conditions and connected by a cohesive interface s = 1 2. In the present
case, the two domains are the concrete specimen and the FRP reinforcement, and the
cohesive interface is defined as the adhesive layer.
For each domain i, the state quantities are the macroscopic damage quantity i, its
gradient gradi and the strain tensor i, all of them depending on position and time:

447

i may be understood as the volume fraction of active links or of undamaged


material. The values of i vary between 0 and 1, where 1 represents the undamaged

state and 0 corresponds to cracked zones.


the gradient gradi accounts for local interactions of the damage at a specific point
on damage of the surrounding area,
the deformation i accounts for the local interaction of the displacement at a specific
point on displacement of its neighbourhood.

The velocities of these states quantities describe the damage evolution in each domain.
As regards the contact surface s, state quantities are i) variables describing the evolution of
the surface, such as the surface damage quantity s and its gradient grads, the latter
accounting for the local damage interaction on the surface, and ii) quantities which
describes macroscopic and microscopic interactions between the domains and the contact
surface. The macroscopic interactions are described by the gap between the domains (u2u1), where ui represents the displacement field of each domain i. Microscopic interactions
are governed by the traces of the damage quantities.
The principle of virtual power and a proper use of the constitutive laws lead to partial
differential equations describing the evolution of damage in the domains and at the
interface [3, 5]. If considering a rate independent problem [5], the equations governing the
numerical problem can be simplified as follows for the domains (1-2) and the interface (3):
divi + fi = 0

(1)

1
ki i wi i i :Ci : i
2
k
k s s s ws s s (u2 u1 ) 2 k s , c ( s c ) k s , p ( s p )
2

(2)
(3)

With i=1,2 equivalent to i=c,p where c and p represent concrete and FRP plate,
respectively; s refers to the bonded interface (glue). and s are the Laplace operators.

i , fi and Ci are respectively the stress tensor, the volume exterior forces and the classical
rank four elasticity tensor. Key damage parameters are: ki and ks, which are the local
interaction coefficients in the domains and at the interface, ks,c and ks,p which respectively
quantifies the interactions between the surface damage and damages within concrete and
within the FRP plate, and ks which represents the rigidity of the adhesive bond.
The initial conditions are: i ( x , 0) i0 ( x ) in i and s ( x , 0) s0 ( x ) on s .
The boundary conditions are: i ni g i on i s , and ki

i
0 on i s ,
ni

a n d c .nc ( x) p .np ( x) s ks (u p uc ) , kc c ks, c (s c ) , kp p ks, p (s p ) o n s


np
nc
where gi represents the surface exterior forces and ni is the outward normal to the domain.
The model can now be used to simulate the damage behaviour of FRP plated specimens.
The main challenge is to evaluate the values of the damage parameters and to consider
448

physically relevant evolutions due to ageing. In this preliminary study, the objective is at
least to propose trends for the damage parameters providing numerical solutions in
agreement with experimental observations.
3.2. Damage simulation and effect of ageing on the fracture line.
FE calculations were carried-out using Deal II software package 6 and considering the
mesh and clamping conditions depicted in Figure 5.

Fig. 5. Mesh used in the finite element calculations.


3.2.1. Determination of the model parameters before and after ageing
All the characteristics and coefficients used in the model are reported in Table 2.
Table 2. Characteristics and theoretical coefficients associated to the domains (concrete,
FRP laminate) or to the interface (glue), and values of the interaction parameters.

ks (N.mm)
ws (MPa.mm)
ks (MPa.mm-1)

Non aged
32000
0.21
0.1
9.76510- 5 MPa
165000 MPa
0.4
0.005
600
4500
0.4
0.01
0.6
2140

Aged
uc (unchanged)
uc
uc
uc
uc
uc
uc
uc
1000
uc
0.01
0.001
476

ks (MPa.mm-1)
ks,c (MPa.mm)
ks,p (MPa.mm)

6000
0.2
0

1333
0.01
uc

Ec (MPa)

Concrete

FRP

kc (N)
wc (MPa)
Ep (MPa)

kp (N)
wp (MPa)
Es (MPa)

s
Interface
(glue)

Interaction
parameters

Initial mechanical characteristics, i.e., Youngs modulus Ei and Poissons coefficient i, of


the constitutive materials (concrete, FRP, and epoxy adhesive) were either determined from
standard mechanical tests or provided by the manufacturers. It was considered that
449

characteristics of concrete and FRP laminates remained globally unchanged after ageing, as
shown in [2]. Differently, the actual property evolutions of the glue (large drop in modulus,
as reported in section 2.2.1.) were taken into account in the modelling of aged specimens.
For the concrete material and the FRP plate, values of the damage threshold (respectively
wc and wp) and values of the local interaction parameters (respectively kc and kp) were taken
from previous studies [3, 4] and kept unchanged after ageing.
As regards the damage parameters of the adhesive bond:
- the value of the damage threshold ws was initially set at 0.6 MPa.mm [5], and was
much lowered after ageing to account for an early damage of the tested joint,
- surface rigidities along the transversal and longitudinal directions, ks and ks , were

calculated by dividing the shear and Youngs moduli by the actual thickness of the
adhesive layer, 0.75 mm. Hence the significant decrease after ageing,
the local interaction parameter ks was initially set at 0.01 [5] and kept unchanged,

The initial value of the interaction parameter between the damage of the glue and that of
concrete ks,c was numerically estimated using the procedure described in [3], and was
lowered after ageing to account for a degradation of interfacial linkages. Interactions
between the damage of the glue and that of the FRP were neglected (ks,p = 0). To
summarize, it is stated that ageing mainly affects the damage threshold and the rigidity of
the adhesive bond (ws, ks and ks ), as well as the interaction parameter ks,c.
3.2.2. Numerical results - Effect of ageing
Figure 6 shows the progression of damage within concrete during shear tests performed on
unaged (a) and aged specimens (b), as computed using parameters listed in Table 2. It is
obvious that concrete damage occurs closer to the interface after ageing. In addition, totally
damaged areas are obtained in the adhesive bond of the aged specimen (Fig. 6.c), whereas
no bond degradation is found in the initial state (s =1 along the lap joint).

t = 130 s

t = 130 s

t = 180 s

t = 180 s

t = 238 s

t = 228 s

a)

b)

t = 228 s

c)
during shear tests performed on an unaged
Fig. 6. Progression of damage c within concrete
specimen (a) and a long term exposed specimen (b). Damage on the adhesive bond (surface
damage s) of the aged specimen is also displayed (c).

450

Figure 7 illustrates the horizontal displacements obtained from the simulations after
complete debonding of the initial and aged specimens, and gives a better visualization of
the fracture line in both cases. It is obvious that the quantity of concrete material that
remains attached to the FRP plate is very limited in the case of the aged specimen, which is
in perfect agreement with experimental evidences (Cf. Fig. 3).

a)

b)

Fig. 7. Horizontal displacements (mm) after debonding of initial (a) and aged (b)
specimens.
4. CONCLUSIONS
This paper investigates the debonding behaviour of CFRP strengthened concrete blocks by
means of experimental and theoretical approaches. It was found that the proposed damage
model is able to depict the actual effect of hydrothermal ageing on the fracture line,
provided adequate values of the damage parameters are considered. Although the variations
of these parameters after ageing are chosen as relevant as possible, absolute values remain
more or less arbitrary at this stage. In a further step, it seems necessary to develop an
identification procedure that allows a proper evaluation of the damage parameters.
5. REFERENCES
1.
2.
3.
4.
5.
6.
7.

Grace, N.F. and Singh, S.B. 2005. Durability evaluation of carbon-fibre reinforced
strengthened concrete beams: Experimental study and design. ACI Structural Journal,
102(1): 40-53.
Benzarti, K., Chataigner, S., Quiertant, M., Marty, C., and Aubagnac, C. 2011.
Accelerated ageing behaviour of the adhesive bond between concrete specimens and
FRP overlays. Construction and Building Materials, 25(2): 523-538.
Freddi, F. and Frmond, M. 2006. Damage in domains and interfaces: a coupled
predictive theory. Journal of Mechanics of Materials and Structures, 1(7): 1205-1233.
Nedjar, B. and Frmond M. Damage, gradient of damage and principle of virtual
power. International Journal of Solids and Structures, 33(8), 1996, 1083-1103.
Benzarti, K., Freddi, F., and Frmond, M. 2011. A damage model to predict the
durability of bonded assemblies Part I: debonding behaviour of FRP strengthened
concrete structures, Construction and Building Materials, 25(2): 547-555.
Bangerth, W., Hartmann, R., and Kanschat, G. 2009. Deal II differential equations
analysis library, technical reference ( http://www.dealii.org).
Chataigner, S., Caron, J.F., Benzarti, K., Quiertant, M., and Aubagnac, C. 2009.
Characterization of composite to concrete bonded interface: Description of the single
lap shear test. European Journal of Environmental and Civil Engineering, 13(9):10731082.

451

8.

Chataigner, S., Caron, J.F., Benzarti, K., Quiertant, M., and Aubagnac, C. 2011. Use of
a single lap shear test to characterize composite to concrete or steel bonded interfaces.
Construction and Building Materials, 25(2): 468-477.

452

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

NUMERICAL SIMULATION OF THERMAL DEFORMATIONS IN


FRP BARS-REINFORCED CONCRETE MEMBERS IN HOT REGION
A. Zaidi1, R. Masmoudi2 and M. Bouhicha1
1
2

Professors, Laboratory of Civil Engineering, University of Laghouat, Algeria


Professor, Department of Civil Engineering, University of Sherbrooke, Canada

ABSTRACT
The coefficient of thermal expansion of fibre reinforced polymer (FRP) bars in the
transverse direction is 3 to 6 times greater than that of the hardened concrete. This thermal
incompatibility between concrete and FRP bars may cause radial cracks within concrete
and ultimately deterioration of the bond between FRP bars and concrete. This Paper
presents a non-linear finite element analysis using ADINA Software to investigate
transverse thermal deformations through concrete cover surrounding FRP bars under high
temperatures up to +60 C varying the ratio of concrete cover thickness to FRP bar
diameter (c/db). This non-linear numerical analysis permits to evaluate the ratio of c/db that
reduces the radial cracks in concrete. Comparisons between numerical, analytical, and
experimental results in terms of transverse thermal deformations are presented.
1. INTRODUCTION
Numerous experimental and theoretical studies were carried out on thermal effects on
concrete structures reinforced with fibre reinforced polymers (FRP) bars (Rahman et al.
1995 Gentry and Hussein 1999, Masmoudi et al. 2005). However, the non-linear finite
element analysis of the thermal behaviour of concrete structures reinforced with FRP bars
under temperatures need more investigation, particularly, distributions of thermal strainsstresses and radial or circumferential cracks propagation through concrete cover from FRP
bar/concrete interface. Unidirectional FRP reinforcement composites used in concrete
structures present an anisotropic thermal and mechanical behaviour. As a consequence, the
transverse coefficient of thermal expansion of FRP bars is 3 to 6 times greater than that of
hardened concrete (Gentry and Hussein 1999, Masmoudi et al. 2005). As a result, a rise in
temperature produces circumferential tensile stresses, within the concrete surrounding FRP
bars, that may cause splitting cracks and eventually degradation of the bond between
concrete and reinforcing bars, reducing the durability and the serviceability of concrete
structures. This paper presents the results of a non-linear finite element analysis utilizing
ADINA software to investigate the distribution of transverse thermal deformations in
concrete cover and glass FRP (GFRP) bars under high temperatures up to +60 C. Also,
this study allows to predict the ratio of concrete cover thickness to FRP bar diameter (c/db)
453

reducing radial splitting cracks in concrete structures submitted to high temperatures.


Theoretical results in terms of transverse thermal deformations predicted from numerical
and analytical models are compared with experimental results.
2. NON-LINEAR NUMERICAL SIMULATION OF THERMAL DEFORMATIONS
2.1 Numerical model
A non-linear finite element study was carried out utilizing ADINA Software to analyse the
transverse thermal deformations in concrete cover and FRP bar for concrete cylinders
reinforced with GFRP bars submitted to temperature increase (T) up to +60 C, as shown
in Figure 1. The cross-section of GFRP bar-reinforced concrete cylinders has been modeled
by means of two dimensional plane stress elements since the axial deformations are
constant. As the symmetric of the cross-section, this study was performed only for the
quarter of the cross-section of reinforced concrete cylinders, as shown in Figure 1b. The
temperature increase, T, applied statically was assumed constant in both concrete and
GFRP bar. This temperature variation, T, was increased from 0 to +60 C with an
increment of +5 C. The meshing of the cross-section of GFRP bar-reinforced concrete
cylinders was carried out utilizing triangular elements with 6 nodes, as shown in Figure 1c.
The selected values of the ratio of concrete cover thickness to FRP bar diameter (c/db)
varied from 0.8 to 3.6, are those used by Masmoudi et al. (2005) in theirs experimental
tests. Mechanical and physical properties of both concrete and GFRP bars presented in
Table 1 and 2, respectively, have been evaluated by Masmoudi et al. (2005). The concrete
has been considered to have a non-linear behaviour, whereas the GFRP bar was considered
to have a linear elastic behaviour.
z

a = db/2
b = db/2 + c

Concrete cover

GFRP bar

y
db/2

(a)

c
(b)

(c)

Fig. 1. FRP bar-reinforced concrete cylinder modeled: (a) Concrete cylinder reinforced
with GFRP bar, (b) Quarter of the cross-section, (c) Meshing of GFRP bar and concrete
Table 1. Mechanical and physical properties of concrete (Masmoudi et al. 2005)
Poissons
Coefficient of
Compressive
Tensile
Modulus of
ratio
thermal expansion
strength
elasticity
strength
fc28 , MPa
ft28, MPa
Ec , GPa
c (x10-6/C)
c
0.17
11.6 2.1
40 3
4.1 0.1
28 2

454

Table 2. Mechanical and physical properties of the GFRP bars (Masmoudi et al. 2005)
Transverse Longitudinal
Transverse coefficient coefficient
Ultimate Longitudinal Longitudinal
Transverse
Bar
Poissons
Poissons of thermal of thermal
tensile
modulus of
modulus of
diameter
expansion expansion
ratio*
ratio
strength* elasticity*
elasticity
db , mm
ffu , MPa
El , GPa
lt
tt
t
l
Et, GPa
(x10-6/C) (x10-6/C)
9.5
0.28 0.02
627 22
42 1
12.7
0.28 0.02
617 16
42 1
7.1
0.38
33 4
8.9 1.8
15.9
0.28 0.02
535 9
42 1
19.1
0.28 0.02
600 15
40 1
25.4
N/a
N/a
0.28 0.02

* Based on 5 identical specimens


2.2 Results of non-linear numerical thermal deformations
Figure 2 presents transverse thermal strains predicted from non-linear numerical model at
FRP bar/concrete interface for concrete cylinders reinforced with GFRP bars under high
temperatures up to +60C having ratios of concrete cover thickness to FRP bar diameter
(c/db) varied from 0.8 to 3.6. This figure proves that the transverse strain curves are linear
and similar until temperature variations (T) +35, +45, and +50 C corresponding to ratios
c/db of 0.8, 1 to 1.5, and 1.9 to 3.6, respectively. From these temperature variations, the
thermal strain curves increase abruptly as a consequence of radial cracks produced within
concrete at the interface which are great and profound, as shown in Figure 3. It is clear in
Figure 3 that the depth of radial cracks decreases with the increase in the ratio of c/db
despite the increase in temperature variations. Figure 4 shows that the ratio c/db variation
has insignificant effect on transverse thermal strains at FRP bar/concrete interface for a
temperature increase less than or equal to +30 C. However, for T > +30 C the transverse
thermal strains were affected with ratios of c/db<1.9. Figure 5 illustrates transverse concrete
strain curves versus temperature variations on the external surface of concrete cover for a
ratio of c/db varied from 0.8 to 3.6. This Figure shows that transverse strain curves are
linear and similar up to +35C and 55C corresponding to ratios of 0.8 and 1 to 1.5,
respectively. From these thermal loads the transverse concrete strains increase suddenly
because of the appearance of radial cracks on the external surface of concrete cover.
4500

Transverse thermal strain (m/m)

4000

C.#10.34
c/db=3.6
C.#13.32
c/db=2.5

3500

c/db=1.9
C.#16.30

3000

C.#25.38
c/db=1.5

2500

C.#19.22
c/db=1.2

2000

c/db=1
C.#25.25
c/db=0.8
C.#25.19

1500
1000
500
0
0

10

20

30

40

50

60

70

Temperature variation (C)

Fig. 2. Numerical transverse thermal strains at FRP bar/concrete interface for reinforced
concrete cylinders having different ratios of c/db
455

lcr = Depth of radial

(c) For c/db = 1.9 ; T = +50 C


lcr = 5 mm

(b) For c/db = 1.5 ; T = 45C


lcr = 6.4 mm

(a) For c/db = 0.8 ;T = +35 C


lcr = 10.6 mm

Fig. 3. Radial cracks within concrete cover at FRP bar/concrete interface versus of c/db.

Transverse thermal strain (mm/m)

2500

10 C

2000

20 C
30 C
40 C

1500

50 C
50 C

1000

40 C
30 C

500

20 C
10 C
0
0

0.5

1.5

2.5

3.5

4.5

c/db

Fig. 4. Numerical transverse thermal strains versus c/db ratio and temperature variation T
at FRP bar/concrete interface
7000

Transverse concrete strain (m/m)

6000
5000

C.#10.34
c/db=3.6
C.#13.32
c/db=2.5
c/db=1.9
C.#16.30

4000

c/db=1.5
C.#25.38
c/db=1.2
C.#19.22
c/db=0.8
C.#25.19

3000
2000
1000
0
0

10

20

30

40

50

60

70

Temperature variation (C)

Fig. 5. Numerical transverse thermal strains versus temperature variation at external surface
of concrete cylinders having different ratios c/db

456

3. LINEAR ANALYTICAL SIMULATION OF THERMAL DEFORMATIONS


3.1 Linear analytical models
Analytical models based on the linear-elastic theory were established by Rahman et al.
(1995) and Masmoudi et al. (2005) for a concrete cylinder concentrically reinforced with
FRP bar and subjected to temperature variations (T). At the interface of FRP bar/concrete,
the transverse thermal strains in FRP bar (ft) and in concrete ( ct) due to the radial
pressure P and the temperature variation T, are given by:
ft a t T

1 tt P
Et

[1]
ct a

P r 2 1
2
c c T
Ec r 1

[2]
Where P is the radial pressure exerted by FRP bar on the concrete, obtained from the
compatibility equation of transverse deformations (Eq. 1 = Eq. 2):
P

t c )T
1 r 1
1
c 1 tt

Ec r 1
Et

[3]
where Et is the modulus of elasticity of the FRP bar in the transverse direction; tt is the
Poissons ratio of FRP bar in the transverse direction; t is the transverse coefficient of
thermal expansion of FRP bar; r = b/a is the ratio of concrete cylinder radius to FRP bar
radius (see Figure 1b); Ec is the modulus of elasticity of concrete; c is the Poissons ratio
of concrete; and c is the coefficient of thermal expansion of concrete.
Nevertheless, at the external surface of concrete cover, the transverse thermal strain of
concrete ct(b) due to the radial pressure P and the temperature variation T, is as follows:
ct b

2P
c T
Ec r 2 1

[4]
3.2 Results of linear analytical thermal deformations
Figure 6 shows analytical transverse thermal strains in concrete cover as a function of the
distance () measured from the centroid of concrete cylinder reinforced with GFRP at a
temperature increase T = +30 C for a ratio c/db varied from 0.8 to 3.6. From this figure it
can be observed that the transverse thermal strains are maximums at the interface of FRP
bar/concrete (a) and decrease through concrete cover until theirs minimum values at the

457

external surface of concrete cover (b). Also, for ratios of c/db < 2, the transverse
thermal strains increase with the decrease in the ratio of c/db. However, for c/db 2, the
transverse thermal strains are almost similar. It can be concluded that the ratio of c/db 2
has no big effect on transverse thermal strains in concrete under high temperatures. Figures
7 and 8 illustrate analytical curves of transverse thermal strains of concrete versus high
temperature variations at the external surface of concrete cover of concrete cylinders
reinforced with GFRP bars for ratios of c/db varied from 0.8 to 3.6. From these figures, it is
obvious that the transverse thermal strain increases with the temperature increase. Also, it is
shown that transverse thermal strains are linear and almost similar for ratios of c/db 1.9
(Figure 7). While for ratios of c/db < 1.9, transverse thermal strains are relatively higher
particularly for T +30 C. It can be concluded that the variation of c/db ratio has
insignificant effect on transverse thermal strains at the external of concrete cover of GFRP
bar-reinforced concrete cylinders for c/db 1.9.
800

600

c/db= 0.8

T = +30 C

bar/concrete interface

Transverse thermal strain (m/m)

c/db= 1.5
c/db= 1.2

c/db= 2

400

200

c/db=3.6
C.#10.34
c/db=3.6
c/db= 2
C.#10.20
c/db= 1.5
C.#13.19
c/db= 1.2
C.#19.22
c/db= 0.8
C.#25-19

0
0

db/2

10

15

20

25

30

35

40

45

Distance (mm)

Fig. 6. Analytical transverse thermal strains in concrete versus the distance measured from
the centroid of concrete cylinders at T = + 30C for different ratios of c/db.
800
C.#10.34
c/db=3.6

Transverse concrete strain (m/m)

700

c/db=2.5
C.#13.32
600
c/db=1.9
C.#16.30
500
400
300
200
100
0
0

10

20

30

40

50

60

70

Temperature variation (C)

Fig. 7. Analytical transverse thermal strains versus temperature variations at the external
surface of concrete cylinders reinforced with GFRP bars having ratios c/db 1.9

458

900
c/db=1.2
C.#25.19

Transverse concrete strain (m/m)

800

c/d
C.#19.22
b=0.8

700

c/db=1.5
C.#25.38

600
500
400
300
200
100
0
0

10

20

30

40

50

60

70

Temperature variation (C)

Fig. 8. Analytical transverse thermal strains versus temperature variations at the external
surface of concrete cylinders reinforced with GFRP bars having ratios c/db <1.9
4. COMPARISON BETWEEN THEORETICAL AND EXPERIMENTAL RESULTS
Figures 9 and 10 present a typical comparison between numerical, analytical, and
experimental results in terms of the transverse thermal strains at the interface of FRP
bar/concrete under high temperatures up to +60 C for c/db =1.9 and 0.8, respectively.
These figures show that transverse thermal strains predicted from numerical and analytical
models are in good agreement with experimental results until a temperature increase around
+30C from which numerical and experimental results are relatively higher because of the
presence of radial cracks which is not considered in the linear analytical model. It is noted
that experimental results used in this comparison were obtained from experimental tests
carried out by Masmoudi et al. (2005) on concrete cylinders reinforced with GFRP bars
submitted to temperatures increased from -30 to +80C (T = -50 to +60C).
3000

Transverse thermal strain (m/m)

c/db=1.9
2500

numerical curve
analytical curve
experimental curve

2000

1500

1000

500

0
0

10

20

30

40

50

60

70

80

Temperature variation (C)

Fig. 9. Comparison between numerical, analytical and experimental results of transverse


strains at FRP bar/concrete interface for c/db = 1.9
459

3000

Transverse thermal strain (m/m)

c/db = 0.8
2500

numerical curve
analytical curve

2000

experimental curve

1500

1000

500

0
0

10

20

30

40

50

60

70

80

Temperature variation (C)

Fig. 10. Comparison between numerical, analytical and experimental results of transverse
strains at FRP bar/concrete interface for c/db = 0.8
5. CONCLUSIONS
- The numerical transverse strains curves of concrete versus high temperatures up to +60C,
at FRP bar/concrete interface for a ratio of c/db varied from 0.8 to 3.6, are linear and similar
until thermal loads ranged between +30 and +50C (depending of c/db) from which the
transverse strain curves increase suddenly because of radial cracks occurred in concrete.
-

At high temperature up to +60 C, the ratio of concrete cover thickness to FRP bar
diameter (c/db) has no big influence on the transverse thermal strains at FRP
bar/concrete interface and also at external surface of concrete cover for a ratio of c/db
1.9.
A ratio of c/db 1.9 with a thermal load T +30 C seems to be sufficient to reduce
extremely the radial crack depths in concrete cover for materials used in this study
At the interface of FRP bar/concrete, the transverse thermal strains predicted from
numerical and analytical models are in good agreement with experimental results until
T around +30 5 C from which numerical and experimental results are relatively
important because of the presence of radial cracks within concrete at the interface
which is not considered in the linear analytical model.

6. REFERENCES
1.
2.
3.

Rahman, H.A., Kingsley, C.Y., and Taylor, D.A. 1995. Thermal stress in FRP
reinforced concrete. Proceedings, Annual Conference of the Canadian Society for
Civil Engineering, CSCE, Ottawa, pp. 605-614.
Gentry, T.R. and Husain, M. 1999. Thermal compatibility of concrete and composite
reinforcements. Journal of Composites for Construction, ASCE, 3 (2): 82-86.
Masmoudi, R., Zaidi, A., and Grard, P. 2005. Transverse thermal expansion of FRP
bars embedded in Concrete. Journal of Composites for Construction, ASCE, 9(5):
377-387.

460

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

EFFECTS OF THE INTERPHASE ON THE PROPERTIES OF GLASS


FIBRE REINFORCED POLYMER (GFRP) COMPOSITES
A.S.M. Kamal1, M. Boulfiza2 and S. Panigrahi3
1

Ph.D. Candidate, Division of Environmental Engineering, U of S, Saskatoon, Canada


Professor, Dept of Civil and Geological Engineering, U of S, Saskatoon, Canada
3
Professor, Dept of Agriculture and Bio-resource Engineering, U of S, Saskatoon, Canada
2

ABSTRACT
Although typically present in minor amounts in Fibre Reinforced Polymer (FRP)
composites, sizing, which is primarily used to enhance the fiber matrix adhesion, plays a
major role in the composites strength. However, some recent theoretical studies suggest
that the fiber-matrix interphase, where sizing is located, is also the most likely source for
damage initiation and subsequent performance deterioration in high performance
composites such as VE-Eglass systems under excessively harsh accelerated tests. In order
to study and better understand the effects of sizing on the properties of GFRP rebars,
custom plane GFRP sheets were produced by a Vacuum Infusion Process (VIP) and
evaluated under various exposure conditions. Use of the custom sheets instead of actual
rebars allows one to study, understand, and ultimately model the behaviour of GFRP
composites at an intermediate scale between a microscopic level, as represented by a single
fibre surrounded by a matrix (too simplistic), and the macroscopic level, as represented by
GFRP rebars (too complicated).
1. INTRODUCTION
The benefits of GFRP rebars in highly corrosive environments over steel rods have been
shadowed partly due to a lack of reliable tools for assessing their long-term performance in
the field. Accelerated aging tests have typically been used in practice to estimate the service
life of GFRP rebars without a clear understanding of their limitations. This practice has led
to a number of seemingly conflicting results in the published literature.
Major constituents in GFRP rebars, such as glass fibers and vinyl ester (VE) matrix have
been observed intact even after 12 months of exposure in harsh environments (Mufti et al.
2007). They have found no diffusions of pore solution species, such as Na, K and Ca into
the GFPR rebars. However, the strength loss of GFRP rebars in accelerated aging
experiments has been reported by several researchers (Kim et al. 2008, Benmokrane et al.
2002). Some recent studies suggest that this apparent discrepancy can be explained by the

461

performance of sizing at fiber-matrix interphase. Therefore, this study focuses the sizing
effects on the properties of GFRP composites.
In order to study the sizing effects on GFRP rebars, GFRP custom plane sheets were
produced due to the following reasons. The proposed sample configuration should allow
one to study, the behavior of GFRP composites at an intermediate level between a
microscopic level, as represented by a single fiber surrounded by a matrix (too simplistic),
and the macroscopic level, as represented by GFRP rods (too complicated). Researchers
have encountered erroneous results in detecting the interfacial shear strength by pulling out
a single fiber embedded into a polymer matrix, as fiber breaks before successful pull-out
occurs. Finally, a single fiber embedded in a polymer matrix does not represent a GFRP
rebar that contains large volume fraction of fibers. Therefore, an intermediate composite
called the GFRP custom plane sheet was produced with both sized and desized glass fibers
by a vacuum infusion method. The sized and de-sized sheets were manufactured in order to
identify the effects of sizing with respect to tensile strength and mass gain.
2. MATERIALS AND METHODS
2.1 Materials
The GFRP custom plane sheets are composed a fiberglass mat (E-glass), a polymer matrix
and few minor ingredients. The fiberglass mat had a thickness of 0.32 mm and a mass of
319.55 g/m2. The polymer matrix applied was Derakane 411-350 (VE resin) containing
45% styrene with the density of 1.046 g/ml at 25C. As minor ingredients, Methyl Ethyl
Ketone Peroxide (MEKP) and Acetylacetone were promoted in Derakane during the sheet
preparation.
Unsized woven-rovin glass fibers could not be ordered. Hence, regular sized fiber mats had
to be desized in the lab by boiling fiberglass mats in pure water. Only sizing from the mat is
removed as the other ingredients (sand, kaolin, limestone and colemanite) of glass fibers
require much higher temperatures to melt. The weight loss of fiberglass mat as a function of
boiling time is shown in Figure 1. It shows that eight hours of boiling is enough to get a
constant mass loss of fiberglass mats.

462

Mass loss (%)

1.8
1.6

Constant mass loss

1.4

after boiling 8 hrs

1.2
1
0.8
0.6
0.4
0.2
0
0

10

11

12

Boiling time (hrs)

Fig. 1. Mass loss of fiberglass mats as a function of boiling time.

(a)
(b)
Fig. 2. Configuration of custom plane sheet (a) schematic representation, and (b) a real
sample (Figures are not one scale).
2.2 Samples Preparation
The custom plane sheets, as shown in Figure 2, were prepared with the final dimensions of
125 mm x 20 mm x 1.25 mm using two runs of a vacuum infusion technique. The Y and
Z dimensions of the specimens have been chosen so that capillary uptake from the sides
of the sample is excluded. The thickness of the sample, X has been chosen such that only
diffusion and possible chemical reactions in the X direction can take place. Under such

463

conditions, the time lag for the Y and Z direction relative to the glass fibers will be at least
a 100 times larger. This simple configuration of a GFRP composite allows one to study,
understand, and ultimately model the behavior of GFRP composites at an intermediate level
between a microscopic level, as represented by a single fiber surrounded by a matrix (too
simplistic), and the macroscopic level, as represented by GFRP rods (too complicated).
2.3 Methods
The test specimens were initially dried at 50C for about one month to obtain the baseline
mass (Wb). Then they were immersed in deionized water at 4C, 23C and 50C until
saturated. After that, the saturated samples were wiped out and weighed (Wi). None of the
specimens was kept outside more than five minutes during weighing as suggested by
ASTM D 5229. The mass gain and moisture diffusivity coefficient were calculated using
equations 1 and 2, respectively.
W Wb
x 100
Mass gain, % = i
(1)
Wb
Where, Wi is the current specimen mass, and Wb is baseline specimen mass.

h
Moisture diffusivity: Dz

4M m

M M1
2
t t
1
2

(2)

Where, Dz is the diffusion coefficient or diffusivity mm2/s, h is the average specimen


thickness, mm; Mm is the effective moisture equilibrium content, %; and M 2 M1 /
( t2 t1 ) is

the slope of moisture absorption plot in the initial linear portion of the curve,

seconds . An Instron 5500R materials testing machine aided with Bluehill 2 software
was used to determine the tensile strength of test specimens before and after exposures.

3. RESULTS AND DISCUSSION


3.1 Effects of Sizing on the Tensile Strength of GFRP Custom Plane Sheets
As can be seen in Figure 3, the strength of GFRP custom plane sheets varied with the
exposure environments and the specimen types. The higher the exposure temperature, the
higher the strength loss. Moreover, it is worth noticing that the sized specimens have lost
about 10% of their original strength at 23 C, and 28% at 50C with deionized water,
whereas the desized specimens lost about 40% of their strength in both exposure
environments. It seems that sized specimens lose their strength gradually with increasing
temperature, while the desized specimens immediately loss their strength immediately in
the harsh environments. This phenomenon can be attributed to the increased solubility of
sizing as temperature increases. As expected, sized specimens acted more effectively over
desized specimens in retaining strength, at any particular exposure environment considered
here. Sized specimens, for example were 36% stronger at as produced conditions, 96%
stronger at wet condition with 23C and 59% stronger at wet condition with 50C than the
desized ones.

464

Fig. 3: Effect of sizing on the tensile strength GFRP plane sheet specimens at various
exposure environments.
3.2 Effect of Sizing on the Mass Gain of GFRP Custom Plane Sheets
Following their exposure to deionized water, the GFRP sheets continued gaining mass until
they became saturated with deionized water. The sheets showed a typical Fickian behaviour
in gaining mass. All curves showed similar characteristics of linear, concave and plateau
regions, but they took a shorter time to achieve those plateaus at higher temperatures
(Figure 4).

Fig. 4. Mass gain of sized and desized GFRP plane sheets as a function of immersed times.

465

Table 1. Mass Gain Characteristics of GFRP Custom Plane Sheets


GFRP Sheets with sized
GFRP Sheets with desized
Temperature
fiberglass mat
fiberglass mat
of deionized Maximum Coefficient Duration
Maximum Coefficient Duration
water (C)
moisture of diffusion of Experi
moisture of diffusion of Experi
-7
gain
(x10 )
-ment
gain
(x10-7)
-ment
2
mm /s
mm2/s
(wt%)
(days)
(wt%)
(days)
4
0.734
0.86
123
0.720
0.80
123
23
0.828
3.56
39
0.836
3.81
39
50
1.350
50.6
39
1.327
50.5
39
The mass gain-time curve reveals that temperatures has a direct impact on the amount of
water absorbed (mass gain) as well as how fast this process takes place (moisture
diffusivity) for both sized and desized specimens. For instance, the maximum mass gain by
the sized specimens are 0.73%, 0.83% and 1.45% at 4C, 23C and 50C, respectively.
Moreover, the moisture diffusivities have also increased by 0.86 x10-7 mm2/s, 3.56 x10-7
mm2/s, and 50.6 x10-7 mm2/s at 4C, 23C and 50C, respectively (Table 1). Due to the
greater moisture diffusivity at the higher temperatures, the samples at higher temperatures
have saturated faster than samples at lower temperatures. However, the sized and desized
specimens have showed similar behaviours of mass gain and moisture diffusivity at any
particular exposure temperature. For instance, the maximum mass gain at 23C for the
sized specimen is 0.83% and for the desized specimen is 0.84%. Similarly, the moisture
diffusivity at 23C for the sized specimen is 3.56 x10-7 mm2/s and for the desized specimen
is 3.81 x10-7 mm2/s. The overall mass uptake results with respect to maximum moisture
gain, diffusivity coefficients, and the duration of experiment are listed in Table 1.
Though both sized and desized specimens of GFRP custom plane sheets have shown
similar behaviours of mass gain and moisture diffusivity, they have shown different
behaviour in tensile strength retention. The strength loss difference is significant after
exposing sized and desized composites in deionized water at 23C. For instance, the sized
specimen in this study is 96% stronger than the desized one. Thus, the silane coupling agent
an active ingredient in sizing in composites seems to play a significant role in strength
retention even under adverse exposure environments.
Similar to strength retention, strength loss in a composite can be explained with the
deterioration of its components. GFRP rebars, neither glass fibers nor polymer matrix in
GFRP rebars was deteriorated easily even after 12 months of exposure in high alkali
solutions (Kamal and Boulfiza, 2008). Therefore, the strength loss, which has been
observed by several researchers, can be explained with the performance of sizing at the
fiber-matrix interphase.
4. SUMMARY AND CONCLUSIONS
The effects of sizing in GFRP custom plane sheets have been discussed in this study to
understand the contribution of sizing in GFRP rebars. The main reason for manufacturing

466

and studying the GFRP custom plane sheets instead of GFRP rebars has to do with the fact
that the custom plane sheets provide an intermediate scale between the overly simplified
case of a single fibre imbedded into a polymer matrix and the very complicated case of a
typical GFRP rebar with its strongly heterogeneous nature which is still relatively easy to
analyse and interpret. Strength loss mechanisms and mass gain characteristics of the sheet
specimens have been discussed with respect to sizing performance.
Sizing in GFPR custom plane sheets played a major role to retain the tensile strength due to
the strong interfacial bonds between fibers and a matrix. However, the interphase bonding
is shown to be affected by the severity of exposure environments. Although the weak Van
der Waals bonds seem to have been affected by the exposure environments, the strong
covalent bonds still managed to contribute in retaining the tensile strength of sized
specimens. Conversely, desized specimens exhibited lower strength due to the weaker
fiber-matrix bond.
The GFRP custom plane sheets showed a Fickian behaviour of moisture gain until they
become saturated. Both mass gain and moisture diffusivities increased with temperature.
Though both sized and desized specimens have shown similar characteristics of moisture
gain and moisture diffusivity at any particular temperature, they showed distinct behaviours
in terms of tensile strength at adverse exposure environments.
The experimental results of GFPR custom plane sheets is a step forward towards
understanding the effects of sizing on GFRP composites. However, to extrapolate this result
into GFRP composites, fiber volume fraction, manufacturing and curing process have to be
addressed with a greater attention.
5. ACKNOWLEDGEMENTS
The authors gratefully acknowledge the financial support from the Intelligent Sensing for
Innovative Structures (ISIS), Canada; and Saskatchewan Ministry of Agriculture Research
Chair Program in Engineering and NAFGEN-ABIP program.
6. REFERENCES
1.
2.
3.
4.

Benmokrane, B., Wang, P., Ton-That, T.M., Rahman, H. and Robert, J.F. 2002.
Durability of glass fiber-reinforced polymer reinforcing bars in concrete environment.
Journal of Composites for Construction, ASCE, 6(3): 143-153.
Kamal, A.S.M. and Boulfiza, M. 2008. On the interactions of fiber reinforced
polymer composites with highly alkaline concrete pore solutions. Proceedings of the
International Symposium of Theortcl. Chemstry, ISTC, Algiers, pp. 212-218.
Kim, H.Y., Park, Y., H., You, Y.J., and Moon, C.K. 2008. Short-term durability test
for GFRP rods under various environmental conditions. Composite Structures, 83:
37-47.
Mufti, A., Onofrei, M. Benmokrane, B., Banthia, N., Boulfiza, M., Newhook, J.P.,
Bakht, B., Tadros, G., and Brett, P. 2007. Field Study of Glass-Fibre-Reinforced
Polymer Durability in Concrete. Canadian Journal of Civil Engineering, 34(3): 355366.
467

468

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

SUSTAINABLE STRUCTURES USING HYBRID-COMPOSITE


BEAMS
J. Hillman
President and Founder, HC Bridge Company, LLC, Wilmette, IL, USA

ABSTRACT
The "Hybrid-Composite Beam (HCB), is a structural member developed for use in bridges
and other structures. The HCB is comprised of three main sub-components that are a shell,
compression reinforcement and tension reinforcement. The shell comprises a fiber
reinforced plastic (FRP) box beam. The compression reinforcement consists of concrete
pumped into a profiled conduit (generally an arch) within the beam shell. The tension
reinforcement consists of carbon, glass or steel fibers anchored at the ends of the
compression reinforcement. The HCB combines the strength and stiffness of conventional
concrete and steel with the lightweight and corrosion advantages of advanced composite
materials. What results is a cost effective alternative for major infrastructure projects using
state-of-the-art sustainable structures with the following benefits:

LIGHTWEIGHT - 1/10th the weight of concrete.


SAFER capacities exceeding code requirements.
REDUCED CARBON FOOTPRINT 80 % less cement than concrete beams.
CONGESTION RELIEF through Accelerated Bridge Construction.
SUSTAINABLE 100+ Year Service Life.

The benefits listed above result in a significant reduction in greenhouse gases when using
the HCB in place of conventional bridge framing technology. Four HCB bridges are
currently in service. Numerous deployments for 2011 are under design representing a
diverse range of applications.
1. PURPOSE AND NEED
Throughout the history of civilization, engineers have sought to improve on the longevity
of structures in the built environment. Regardless, conventional materials have limited
service lives and despite good intentions, the environment eventually takes its toll on these
structures. Given a worldwide trend towards environmental consciousness, more and more
emphasis is being directed towards sustainability in new construction. This includes not

469

only in material selection, but also in construction techniques that minimize adverse effects
on the environment.
There are many different definitions for sustainability. One of the most popular definitions
can be traced back to a 1987 UN conference. It defined sustainable developments as those
that meet present needs without compromising the ability of future generation to meet
their needs (WCED 1987). Another definition that lends itself well to HCB technology is
Sustainability means using methods, systems and materials that wont deplete resources or
harm natural cycles (Rosenbaum 1993). The development of the The Hybrid-Composite
Beam, or HCB, was fostered by a desire to create a beam that optimizes the structural
performance of each material used and exploits the lightweight and corrosion resistant
properties of fiber reinforced polymer (FRP) to facilitate a structural system that satisfies
these definitions of sustainability.
In its simplest form, the HCB is comprised of three main sub-components that are a fiber
reinforced polymer (FRP) shell, compression reinforcement and tension reinforcement.
The compression reinforcement consists of self-consolidated concrete (SCC) which is
pumped into a profiled conduit within the interior volume of the beam shell. The tension
reinforcement consists of a high-strength, galvanized prestressing strands that run along the
bottom flange of the beam and that are infused with the same vinyl ester resin during
fabrication of the beam shell. These components are identified in Figure 1.

Fig. 1. Fragmentary Perspective of HCB


The beams are manufactured using a closed-mold, Vacuum Assisted Resin Transfer
Method, or VARTM process. In this process, the preforms that include; quad-weave glass
fabrics, untensioned prestressing steel and a low density (polyisocyanurate) foam core are
all placed inside of a simple, inexpensive, reusable composite mold. Once the preforms are
in place, the mold is sealed and vacuum pressure applied to the evacuate all of the air.
While maintaining vacuum pressure at 1 atmosphere, a vinyl ester resin is pulled into the
mold to wet-out the preforms and consolidate the glass, foam and resin into a monolithic
unit. The entire process for lay-up and infusion and demolding of a beam can be performed
in a one day cycle.
In its most simplistic embodiment, the beams are simply supported and the profile of the
compression reinforcement follows a parabolic curve, emulating the funicular shape of the
470

applied dead load and live load moment envelope. The vertical component of the thrust in
the compression arch also results in a dramatic reduction in the amount of shear that has to
be carried by the webs of the FRP shell. In this simplest embodiment, the beam essentially
functions like a tied-arch in a glass box. Although this is the simplest form of the beam, the
same technology can be employed to manufacture beams with variable depths or widths.
The beams can also be made continuous over several supports with simplistic splicing
technology using the same materials and manufacturing processes as will be demonstrated
later.
In the final embodiment of a bridge, the HCB can be erected in the same manner as
prestressed concrete or steel beams, but with smaller equipment due to the lighter weight.
Prior to injecting the compression reinforcing, the self weight of an HCB is approximately
1/3 that of a steel beam, and 1/10 that of a prestressed concrete beam having the same
strength and stiffness. This results in a net savings in shipping and erection costs and
reduced greenhouse gases due to the smaller and reduced equipment demands. Once filled
with concrete, the HCB has essentially the same mass as an equivalent steel beam, which is
desirable from the standpoint of dynamic behavior. The combination of an efficient use of
steel, concrete and FRP, coupled with the lightweight characteristics of the HCB results in a
cost effective application of FRP in a structural member.
The HCB also provides flexibility in construction in that there are basically three methods
of deployment which vary depending on project requirements related to weight and speed
of installation. These variations include:

Installing as an emply shell and subsequently filling with concrete wth the HCB shell
serving as a stay-in-place form.
Prefill the arch concrete in the HCB and erect the beams as true hybrid-composite that
supports the deck casting.
Where speed of erection is the most important criteria, such as in Accelerated Bridge
Construction (ABC), the arch concrete is precast in advance. Preferrably this would
be done in a precast yard or staging area that would allow for HCBs to be set up to
the geometry consistent with the final erected cross-slopes and superelevation
transitions. At this point the concrete deck is checkerboard cast with either
longitudinal matchcast joints or small sections of omitted concrete that allow for castin-place closures after the units are erected.

This final erection sequence can result in the ability to install and open a bridge
superstructure in as little as one day. The remaining portions of this paper will expand
upon the design aspects of the HCB as well demonstrate the various methods of
deployment that exemplify the applicability to ABC.
2. HISTORY OF THE HCB
The author originally conceived the concept of the HCB in 1996. The initial validation of
HCB technology included not only developing design limit states, but also devising the
manufacturing processes to construct a beam of this size with such an unusual embodiment

471

of materials. Through years of experimentation in collaboration with the University of


Delaware Center for Composite Materials (UD-CCM), a cost-effective and reliable
method of manufacturing was developed. Over the course of this initial research several
full size prototype beams were fabricated, subjected to fatigue testing in the laboratory and
finally loaded to failure to validate the design limit states. The HSR-IDEA research
culminated in the deployment and testing of the worlds first composite railroad bridge on
November 7, 2007. This first bridge with a span of 9.14m has since been subjected to over
237 million gross tonnes (MGT) of heavy axle freight train loading on the Heavy Axle
Load (HAL) loop at the Transportation Technology Center, Inc. (TTCI) in Pueblo, CO.
Testing was discontinued in June 2010, but throughout the course of testing there was no
evident change in the performance of the bridge.
3. DESIGN ASPECTS OF THE HCB
Although the HCB contains materials that are generally new to most practicing structural
engineers, with a basic understanding of the mechanics of Bernoulli-Euler beam theory and
a working knowledge of standard bridge design codes, it is not difficult to assess the load
carrying capacity of the HCB. In fact most design codes, including AASHTO, AREMA
and the CHBDC are compartmentalized and allow the engineer a fair amount of flexibility
in assessing how forces are resisted by a structure. Further, the applied loads as well as the
load and resistance factors can easily be rationalized for assessing the response and
structural capacity of the HCB.
A thorough explanation of the structural behaviour of the HCB warrants a completely
separate paper. As noted above, there are already many design tools in the existing codes
and textbooks that allow the engineer to make same reasonable assumptions in designing an
HCB. For example, the ultimate capacity in bending can be calculated using strain
compatibility and force equilibrium, just as in a reinforced concrete beam. Even neglecting
the contributions from the FRP box, one can calculate a bending capacity that is safe and
very close to those validated in the laboratory.
For the most part, the controlling limit state for the HCB typically tends to be live load
deflections. Although the FRP shell laminates have reasonably high strengths, on the order
of 414 MPa, they also have a relatively low modulus of elasticity at 22 GPa. The laminate
thicknesses are also relatively thin at 5mm. Subsequently most of the stiffness of the HCB
is derived from the concrete arch and the tension steel in the bottom flange of the beam.
The HCB Bridge also derives a great deal of stiffness from the composite action between
the HCB and the supported concrete deck. For HCB bridges constructed to date, the spanto-depth ratios have ranged between span/16 to span/25. For railroad bridges these ratios
will be more on the order of span/12. Ultimately, the most effective way to increase the
flexural rigidity to satisfy a live load deflection criterion of span/800 is simply to add
additional steel tension reinforcement.
Shear behaviour of the HCB is equally as intriguing and perhaps a bit more complex. In
the HCB, shear resistance is facilitated by three mechanisms acting in concert. To start
with, the primary reason for the arch shape of the compression reinforcing in the HCB is to
carry a significant portion of the shear as direct compression forces down to the thrust line
472

and into the bearings. Another significant component of the shear mechanism is that the
quad-weave fabrics in the HCB shell provide tremendous shear capacity due to the +/- 45degree plies in the laminate. Test results have demonstrated that the diagonal plies in the
FRP shell help facilitate tension field action at the center part of the HCB, where the angle
of the arch is very slight. Finally, the third mechanism involves a thin concrete web
(typically on the order of 75 mm) that extends vertically along the longitudinal centerline of
the beam between the top of the arch and the bottom of the supported deck. This concrete
web is actually in fluid communication with the arch concrete when cast. Further, the
diagonal shear connectors extending from the arch to the deck provide for the most efficient
reinforcing of this concrete web. Through this complex interaction in shear, laboratory
results have typically indicated shear capacities well beyond the code requirements.
As validated by laboratory testing, it should be noted that there are inherent benefits to
public safety resulting from the structural behaviour of the HCB. Since design is usually
governed by satisfying live load deflections the HCB consistently exemplifies significantly
increased capacity for strength. In fact there is enough redundancy in the HCB that if the
bridge deck were to completely deteriorate, the HCB would still have sufficient capacity to
carry all of the factored design loads applied to the bridge. This has been further validated
on subsequent project testing where ultimate capacity tests have demonstrated load rating
factors on the order of 2.68 for Operating (equivalent to HS-54) and 3.47 for Inventory
(equivalent to HS-69).
4. FABRICATION AND CONSTRUCTION
As noted earlier, the HCB is fabricated using a closed mould, vacuum-assisted resin
transfer method (VARTM) to manufacture the composite shells. In this process, all of the
preforms, i.e. glass, foam and steel are placed in a mould. The mould is completely
evacuated of air at which point the resin is pulled into the mould, wetting out all of the
constituent components into a monolithic shell. By filtering out any hazardous materials
through the vacuum system, the VARTM process becomes is an environmentally friendly,
zero VOC emission manufacturing process.
One of the biggest advantages of HCB is the extremely lightweight for shipping an
erection. Each of the 17.8m long beams for the High Road Bridge weighed only 2 tonnes.
As a result, all six HCBs for the bridge could be shipped on a single truck instead of six
trucks as would have been required for precast concrete beams. Further, the contractor was
able to erect all six beams using a 30 tonne utility crane instead of mobilizing a 200 tonne
crane. As the contractor already had a 30 tonne crane on sight, this resulted in a significant
cost savings for installation. Examples of these benefits will be demonstrated in the
following pilot projects.
4.1 High Road Bridge
The first permanent highway bridge to utilize the HCB is the High Road Bridge. The
framing plan for the High Road Bridge emulates a very conventional bridge system. The
bridge itself is a 17.4m single span bridge that carries two lanes of traffic over Long Run
Creek. The superstructure is comprised of six 1067mm deep by 508mm wide HCBs
473

supporting a conventional 200mm reinforced concrete deck with an out-to-out dimension of


13.15m and a curb-to-curb width of 12.2m. The HCBs are spaced at 2.24m centers. The
shipping and erection advantages of the HCB can be seen in Figures 2 and 3.

Fig. 3. Erecting HCB with 30 tonne

Fig. 2. Six HCBs on Single Truck

It should be noted that the framing configuration was intentionally designed to be


interchangeable with precast beams in that compatibility with conventional bridge design
and construction can help expedite the acceptance of a new technology in the bridge
industry. This particular deployment exploits the lightweight advantages of the HCB in
that the arch concrete and deck are both cast-in-place.
The High Road Bridge was originally scheduled to be open at the end of November, 2008.
Despite a month of delays from weather and utility relocations as well as the prototype
testing and learning curve associated with the first deployment of this new technology, the
bridge was opened three months ahead of schedule on August 25, 2008. As a result of
careful monitoring of every aspect of construction, this project was completed ahead of
schedule as well as under budget.
When all was said and done, the cost of the HCBs for the High Road Bridge was no more
than it would have been for a conventional bridge. The only premium was related to the
prototype testing and research aspects of the project that were covered at 100% by the
discretionary IBRD funds. In actuality the owner saved the 20% cost share of the girder
framing costs as these were covered under the IBRD funds as well. Despite learning curves
associated with the implementation of new technology, there was no impact to the project
schedule as a result of the use of the HCB. With improvements to the overall system
resulting from lessons learned on the High Road Bridge Project, further refinements to
materials and manufacturing processes make this technology competitive with concrete and
steel on a first cost basis.
4.2 Knickerbocker Bridge
The next level of development of HCB Bridge technology is exhibited in the
Knickerbocker Bridge, recently completed in Boothbay, Maine. This structure is
comprised of an eight span continuous HCB superstructure with 18.3m end spans and
474

21.3m interior spans. The framing system comprises eight 838mm deep HCBs at 1.24m
centers supporting a 178mm reinforced concrete deck. Additional negative moment
reinforcing steel is provided in the deck, over the piers to make the superstructure
continuous for live load. This bridge in particular sets a new benchmarks for the
deployment of composite bridge technology with an overall length of 164.6m.
This particular configuration of an HCB bridge is designed to be a direct substitution for
precast box beam bridges. Again, the actual beam unit is only 610mm wide at the base, but
still has a 1244mm top flange. Once erected, the tips of the flange overhangs are side by
side, eliminating the need for any deck forming. Again, these types of girders can be
installed empty, with concrete arches or with the deck already cast. Even with the top slab
precast in a yard, these units would still be lighter than precast box beams and would not
require an additional overlay. At this time the Knickerbocker Bridge is scheduled to be
open to traffic in spring of 2011.
4.3 Railroad Test Bridge, Pueblo, CO.
The first HCB Bridge deployed under live loads was the railroad bridge constructed under
the HSR-IDEA project. Several years of research went into the validation of the
technology, including development of the manufacturing processes to make the HCB
economically viable. This long journey of development culminated with the live load test
of a 9.32m span on the Association of American Railroads (AAR) test track at the
Transportation Technology Center, Inc. (TTCI) in Pueblo, CO as seen in Figures 4 and 5.

Fig. 4. Erection of HCB with 50 ton Crane

Fig. 5. Live Load Test - Railroad

The railroad bridge exemplifies the optimization of rapid installation as it relates to


Accelerated Bridge Construction of ABC. This is not unusual for railroad construction; in
fact for the most part it is standard operating procedure. Many railroad bridges are installed
on active, revenue service tracks where traffic interruptions can often be limited to only 8
hours for a complete span replacement. As a result, for an HCB installation under these
conditions, it is necessary to precast the arches as well as the deck and ballast curbs.
Although the bridge installed at TTCI was placed on a test track without the limitations
imposed by revenue service, there was still a desire to demonstrate the rapid installation.
475

Likewise, the commercial success of this new technology is predicated on standardization


and compliance with conventional bridge technology. For this reason, the HCB test bridge
was designed and fabricated to facilitate a direct substitution with the existing precast
concrete box beam bridge currently on the Facility for Accelerated System Testing (FAST)
at TTCI.
The first live load train test on an HCB Bridge took place on November 7, 2007. Since this
time, the HCB Railroad Bridge has been subjected to over 1,000,000 cycles of fatigue of
heavy axle loading. It should also be noted that the measured deflection was on the order
of 9mm, which is below the allowable live load deflection of 14mm or span/640.
5. CONCLUSIONS
As a result of the collapse of the I-35W Bridge in Minnesota on August 1, 2007, it has
become common knowledge that no less than 26% of the over 600,000 bridges in the
United States, National Highway System (NHS) bridge inventory are either structurally
deficient or functionally obsolete. In 2005 it was estimated by the American Society of
Civil Engineers that the cost of repairing and/or replacing these 150,000 plus bridges would
require a capital expenditure of more than $9.4 billion dollars a year for 20 years. These
types of events and statistics clearly identify the need to develop advanced bridge
technologies that provide longer lasting and safer structures.
The introduction of the HCB results in an innovative, cost effective and sustainable
advancement in engineering with far reaching socioeconomic impacts. The HCB provides
a revolutionary bridge technology that demonstrates a commitment by the civil engineering
community to not only rectify the state of our decaying infrastructure, but also to provide a
solution to this problem that will reduce the burden of decaying infrastructure for future
generations.
In addition to railroad and highway bridge structures, the HCB provides an attractive
alternative to other types of structures as well, including marine structures like docks, piers
and berths. The technology also lends itself well to green roof technology. There are many
hurdles to introducing a new technology to a conservative industry like such as exists in the
bridge world, not the least of which is our own reluctance as practicing engineers to
embrace change. Regardless, opportunities still exist to advance our profession, albeit
slowly and methodically. The HCB challenges the engineer to embrace change.
6. REFERENCES
1.
2.
3.

United Nations World Commission on Environment and Development. 1987.


Rosenbaum. 1993 (www.arch.wsu.edu/09%20publications/sustain/defnsust.htm).
Hillman, J.R. 2000. Plasticon-Optimized Composite Beam System U.S. Patent
6,145,270. United States Patent and Trade Office, November 14.

476

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

CONSIDERATIONS FOR DESIGN OF FRP RETROFITTED


REINFORCED CONCRETE WALL PANELS SUBJECTED TO BLAST
LOADING
E. Jacques1 and M. Saatcioglu2
1
2

Graduate Student, Dept. of Civil Engineering, Univ. of Ottawa, Ottawa, Ontario, Canada
Professor, Dept. of Civil Engineering, Univ. of Ottawa, Ottawa, Ontario, Canada

ABSTRACT
Externally bonded fibre reinforced polymer (FRP) retrofits for reinforced concrete elements
provide an efficient, economical and accepted method of improving structural performance.
A large volume of research has been conducted on FRP retrofitted structures under static
and quasi-static loads. Previous research has resulted in the publication of numerous retrofit
design methodologies, including those outlined in ACI 440 (2008) and CSA S806-02
(2007) documents. However, only a limited number of publications address the issue of
designing FRP retrofit systems to resist blast induced impulsive loads.
A comprehensive research program has been underway at the University of Ottawa to
develop FRP retrofit methodologies for structural and non-structural elements. Simulated
blast testing of as-built and FRP retrofitted reinforced concrete wall panels were performed
using the University of Ottawas Shock Tube. The test program resulted in a large volume
of research data for the development of blast-resistant design guidelines. This paper is
aimed at highlighting some of the challenges associated with the use of FRPs as a blast
retrofit material.
Test results indicate that FRP retrofitted reinforced concrete wall panels may survive initial
inbound displacements, only to fail by moment reversals during the negative displacement
phase, unless retrofitted on both sides. They also indicate that the increase in flexural
stiffness due to the application of FRP, relative to that associated with shear distress, may
result in shear failures. A discussion on the increased likelihood of blast-induced shear
failures in FRP retrofitted elements will be highlighted. The paper will provide guidance for
improved design of FRPs in protecting reinforced concrete walls against blast-induced
shock waves.

477

1. INTRODUCTION
Critical infrastructure may require an upgrade, or retrofit, to improve structural
performance when subjected to impulsive blast loads. One potential retrofit technique is to
increase strength and stiffness of a deficient structure through the use of externally bonded
fibre reinforced polymer (FRP) composites. This type of retrofit results in structural
elements with a greater ultimate resistance and stiffness than similar as-built structural
elements. However, this benefit comes at the expense of ductility and energy dissipation
characteristics. The advantages of externally bonded FRP retrofits, including adaptability,
high strength-to-weight ratio and chemical inertness, have made them ideally suited as an
efficient, economical and accepted method of improving structural performance (ACI 440,
2008). To-date they have been used extensively to retrofit structures suffering from
inadequate lateral load resistance, excessive deflections and corrosion related problems.
However, the application of these retrofits to reduce the effects of blast-induced impulsive
loads remains largely unexplored, with many outstanding issues preventing their
widespread application to blast resistant design.
Previous research has shown that externally bonded FRP composites can increase the
flexural strength and stiffness of reinforced concrete slabs, beams and walls resulting in
increased blast resistance (Ross, Purcell and Jerome, 1997; Buchan and Chen, 2007; Silva
and Lu, 2007). However, despite the clear improvement in blast resistance reported in
literature, a number of issues limiting the effectiveness of these retrofit systems have been
identified. Under blast loading, the most commonly reported issue is debonding of FRP
from the concrete substrate prior to achieving the complete effectiveness of the material
itself (Buchan and Chen, 2007). Another challenge with FRP retrofits is the increased
likelihood of shear or a combined flexure-shear failure (Razaqpur et al, 2007; Silva and Lu,
2007). It is believed that the change from flexure to shear failure for stiff members may be
attributed to the activation of higher modes of vibration during impulsive loading
(Magnusson, 2007). Another commonly reported issue is the increased likelihood of
premature failure as a result of reversal of internal moments caused by rebound
displacements (Ross, Purcell and Jerome, 1997; Silva and Lu, 2007). Addressing these
issues is critical in ensuring the continued effectiveness of blast-resistant externally bonded
FRP retrofits.
This paper describes the experimental program and preliminary results of simulated blast
testing of reinforced concrete wall strips retrofitted with externally bonded carbon fibre
reinforced polymer (CFRP) sheets. The experimental results will be discussed in such a
way that highlights the challenges associated with ensuring an adequate level of
performance through the use of this blast-resistant retrofit methodology.
2. EXPERIMENTAL METHODOLOGY
2.1 Specimen Construction and Retrofit Details
Five reinforced concrete wall strips, 2440 mm long by 440 mm wide, were constructed.
Three of these wall strips, with 80 mm thickness, were collectively part of companion set 1
(CS1). The remaining two were 120 mm thick, and part of companion set 2 (CS2). Each
478

specimen was doubly reinforced with four 6.3 mm diameter non-deformed steel
reinforcement on the top and bottom faces. Clear cover to reinforcement was 6 mm. The
specimens were subjected to simply-supported bending with a clear span of 2232 mm
between supports.
After the wall strips had cured, two specimens from companion set CS1 (known as CS1-R1
and CS1-R2, respectively) and one specimen from companion set CS2 (known as CS2-R)
were retrofitted with one layer of externally bonded CFRP/epoxy laminate. This retrofit
covered the entire clear span of the unloaded face terminating just prior to the supports. The
two unretrofitted as-built specimens, CS1-C and CS2-C, served as a control for comparison
of results against the CFRP retrofitted members. Figure 1 (a) shows a typical as-built wall
strip. The manufacturer reported that the CFRP/epoxy laminate had an ultimate tensile
strength of 876 MPa, elastic modulus of 72.4 GPa and rupture strain of 1.2%, based on a
gross composite thickness of 1 mm (Fyfe Co. LLC, 2009).
2.2 Shock Tube Testing Facility
The University of Ottawas Shock Tube Testing Facility was used to simulate the shock
waves produced by the detonation of high explosives. The pneumatically driven shock tube,
shown in Figure 1 (b), is comprised of three main components: a variable length driver
section, a spool section and an expansion section with test frame. Shock wave energy and
parameters are generated within the driver section. A differential pressure firing mechanism
contained within the spool section is responsible for triggering the shock tube. The shape of
the expansion section ensures planar shock wave formation. Specimen testing occurs at the
test frame located at the end of the expansion section. In terms of the shock wave
parameters at the test frame, the shock tube can generate a reflected pressure up to
approximately 100 kPa with a positive phase duration of between 5 and 70 ms, and
reflected impulse between 250 and 2700 kPa-ms. More information regarding the operation
and capabilities of the shock tube may be found in Lloyd (2010).

(a) Typical wall strip with load transfer

(b) Driver and spool sections

Fig. 1. University of Ottawa Shock Tube Testing Facility.

479

2.3 Load Transfer Device and Instrumentation


Since the wall strips did not cover the entire opening of the shock tube test area, a load
transfer device was used to transfer shock wave pressures to the face of the specimens. The
load transfer device consisted of a light gauge sheet metal skin attached to 15 rigid out-ofplane steel beams, equally distributed over the clear span of the wall strips. Shock wave
pressures were applied to the 2.432 m wide by 2.032 m tall load transfer device and
transferred to the loaded face of the wall strips as a series of point loads uniformly
distributed along the length. The load transfer device and the specimen were not connected
and the dynamic response of the two systems is coupled only up to the time of maximum
displacement inertia of the load transfer keeps it in contact with the specimen. The use of
a load transfer device adds 394 kg to the total mass of the system. Furthermore, the load
transfer device is flexible and does not affect the resistance of the dynamic system. Figure 1
(a) shows a typical wall strip with associated load transfer device prior to testing.
Experimental data was recorded using a high-speed data acquisition unit recording at 100
kilo-samples per second. Shock wave reflected pressures, as well as FRP and steel strains,
and specimen displacements were recorded. Each wall strip was instrumented with three
LVDTs to record displaced shape. One LVDT was located at mid-span, while the
remaining two LVDTs were positioned 450 mm and 900 mm away from mid-span.
2.4 Testing Regime
To facilitate direct comparison of experimental results, specimens within each companion
set were subjected to the same reflected pressure-impulse combinations. The first simulated
explosion performed on each set was selected to generate elastic behaviour in all
specimens, while subsequent tests were performed at increased pressure-impulse
combinations.
3. EXPERIMENTAL RESULTS
3.1 Advantages of Blast-Resistant FRP Retrofits
A typical displacement-time history plot generated during testing is shown in Figure 2. This
particular plot compares the experimental mid-span displacements of as-built specimen
CS1-C and the retrofitted specimens CS1-R1 and CS1-R2. For this particular test, each
specimen was subjected to a shock wave with a reflected pressure of 46.2 kPa, reflected
impulse of 317 kPa-ms and positive phase duration of 16.2 ms. At this pressure-impulse
combination, both the control and one of the retrofitted specimens failed: control specimen
CS1-C failed by excessive inelastic deformations with a maximum displacement of 167.8
mm (support rotation 8.6) at a time of 90 ms; retrofitted specimen CS1-R1 failed by
debonding caused by poor workmanship (insufficient saturation of the CFRP sheet with
epoxy) with a maximum displacement of 146.9 mm (support rotation 7.5) and at a time of
85 ms. Conversely, the other retrofitted specimen, CS1-R2, experienced little damage with
a maximum displacement of 42.5 mm (support rotation 2.2) at a time of 29.7 ms.
Furthermore, CS1-R2 remained elastic with no permanent deformations after testing, while

480

CS1-C and CS1-R1 both had significant permanent deformations. Note that these results
are also typical of the observed behaviour of the thicker wall strips of companion set CS2.

180
150
120
90
60
30
0
-30

90
75
60
45
30
15
0
-15

CS1-C - Control
CS1-R1 - With FRP
CS1-R2 - With FRP
Reflected Pressure

-30

30

60

90 120 150 180 210 240 800

Reflected Pressure, P r (kPa)

Midspan displacement, (mm)

A significant reduction in maximum displacement and time-to-maximum displacement for


all CFRP retrofitted specimens was observed over corresponding as-built members for both
companion set CS1 and CS2. However, this comparison does not include specimen CS1-R2
which failed due to poor workmanship. At failure, the maximum support rotation for asbuilt one-way wall panel strips was between 7 and 10. The corresponding CFRP
retrofitted wall strips were able to survive the same simulated explosions and sustain elastic
support rotations of approximately 2 - a reduction in maximum response by a factor of 3.5
to 5 times the as-built response. Prior to failure, the CFRP retrofitted wall strips were able
to sustain elastic support rotations of between 2 and 4. Retrofitted panels remained elastic
prior to failure. Furthermore, the retrofitted specimens required much larger pressureimpulse combinations to fail the retrofit. These tests show that the use of an externally
bonded FRP retrofit to increase flexural strength and stiffness can improve the blast
resistance of reinforce concrete walls.

900

Time, t (ms)
Fig. 2. Displacement-time history comparison of Companion Set 1.
A bar chart comparing maximum displacements for companion set 1 for progressively
increasing reflected pressure-impulse combinations is shown in Figure 3. The trend
observed during testing is that the maximum displacement of all specimens increases as the
pressure-impulse combination of the shock wave increases. However, as wall strips
retrofitted with externally bonded FRP have significantly higher strength and stiffness
characteristics, a much larger shock wave is required to fail the retrofit. This, of course,
assumes that premature failure of the retrofit system does not occur. In this case, retrofitted
specimen CS1-R2 failed at a reflected pressure-impulse combination of 81.1 kPa and 599
kPa-ms, respectively. This corresponds to an approximately 200% increase in the pressureimpulse combination required to fail the retrofitted specimen relative to the control. This
shows that externally bonded FRP retrofits can also increase the size of the explosion
required to generate failure the retrofitted structure.

481

200

Excessive inelastic
deformations

Debonding (Poor

Plate-end interfacial

workmanship)

debonding

50

CS1-R2

100

CS1-R1

150
CS1-C

Maximum
Displacement (mm)

250

0
Shot 1

Shot 2

12.5 kPa
93 kPa-ms

46.2 kPa
317 kPa-ms

Shot 3

Shot 4

Shot 5

73.8 kPa
81.1 kPa
66.0 kPa
452 kPa-ms 511 kPa-ms 599 kPa-ms

Fig. 3. Comparison of maximum mid-span displacements for companion set 1 subjected to


progressively increasing pressure-impulse combinations.
3.2 Moment Reversal
Externally bonded CFRP sheets add considerable elastic flexural strength and stiffness to
retrofitted members. This results in an overall reduction in displacement and in increased
pressure-impulse combination required to fail the retrofit. During the inbound displacement
phase, kinetic energy of retrofitted specimens is stored as elastic strain energy in the
externally bonded CFRP. However, as the traditional mechanisms of energy dissipation
have been eliminated or reduced through retrofitting (i.e. no crushing of concrete and
reduced yielding of steel), elastic strain energy is violently released as kinetic energy during
the rebound displacement phase. These retrofitted panels interacted violently with the
transverse beams of the load transfer device during rebound. Although the load transfer
device prevented the formation of negative displacements, rebound of the retrofitted
specimens bent many of the transverse beams. The displacement-time history plot for
specimen CS1-R2, shown in Figure 3, is an excellent example of the load transfer device
effectively preventing rebound displacements. Had the load transfer device not prevented
rebound, it is likely that the specimens retrofitted on one face only would have survived the
first inbound displacement, only to suffer extensive damage during the rebound
displacement phase as a result of moment reversal. Based on these experimental
observations, it is recommended that FRP retrofits be applied to both faces of the member
to reduce the likelihood of failure due to moment reversal during rebound.
3.3 Shear Failure
Applying externally bonded FRP sheets to the tensile face of a reinforced concrete wall or
slab will increase its flexural capacity but will not affect its shear strength significantly.
This increases the risk of diagonal shear failure as a result of increased applied loads
exceeding shear strength. When excited by dynamic blast-induced response, members with
increased flexural stiffness exhibit a greater tendency to fail in shear due to the activation of
higher modes of vibration (Magnusson, 2007). Increasing the stiffness of a member with
flexural FRP reinforcement will also increase its natural frequency. In many cases, due to
the rapid nature of blast loading, an increase in the natural frequency will result in higher

482

dynamic load factors and necessitate larger shear capacities. Determining the shear
response of structures subjected to impulsive loading is further complicated as the dynamic
shear force is a function of time-varying member resistance, inertia and applied loading
(Biggs, 1964). Accordingly, the effect of member resistance and inertia must be accounted
for in design to ensure brittle shear failures are avoided and the full capacity of the flexural
FRP retrofit is utilized.
3.4 Debonding
Two types of FRP debonding failures were observed during testing. The 80 mm thick
retrofitted wall strips of companion set CS1 failed by plate-end interfacial debonding
(shown in Figure 4 (b)). The 120 mm thick retrofitted wall strips of companion set CS2
failed by critical diagonal crack debonding (shown in Figure 4 (c)). Plate-end debonding
failures occur at the location of the ends of the FRP sheet. High normal and interfacial
shear stresses at this location exceed the tensile capacity of the concrete. Critical diagonal
crack debonding failures occur when large flexure-shear cracks form in regions of
intermediate moment and shear. This results in high local interfacial shear stresses between
the concrete substrate and the FRP sheet (Teng and Chen, 2009). The flexural stiffness of
the thick specimen, CS2-R, was increased to such an extent that flexure-shear interaction
became the cause of retrofit failure. This indicates that shear capacity limits the
effectiveness of blast resistant externally bonded FRP retrofits by increasing the likelihood
of intermediate flexure-shear crack induced debonding.

(c) Critical diagonal crack


deboning of specimen CS2R
Fig. 4. Typical failure modes of one-way wall strips observed during testing.

(a) Typical as-built


specimen after testing

(b) Plate-end debonding of


specimen CS1-R2

5. CONCLUSIONS
This paper presented a brief review of the experimental program and preliminary results of
simulated blast testing of reinforced concrete wall strips retrofitted with externally bonded

483

carbon fibre reinforced polymer sheets. Based on the observed effectiveness of externally
bonded FRP sheet to improve blast performance, the following conclusions can be drawn:
1. FRP retrofitted wall strips resulted in a significant reduction in maximum displacement
by a factor of 3.5 to 5 relative to as-built specimens at failure.
2. FRP retrofitted wall strips were able to survive much larger pressure-impulse
combinations relative to their as-built specimens prior to failure.
3. FRP retrofitted wall strips remained elastic prior to failure.
4. Several challenges limit the effectiveness of FRP retrofits to improve blast resistance.
These include:
a. FRP retrofitted structures may fail by moment reversal. Retrofitting both inbound
and rebound faces may mitigate this failure mode.
b. Increased flexural strength and stiffness resulting from FRP retrofitting may lead
to either diagonal shear failure or flexure-shear crack induced debonding.
c. Debonding of FRP from the concrete substrate will reduce the effectiveness of the
retrofit and should be avoided at all costs.
6. REFERENCES
1.

ACI Committee 440. 2008. Guide for the design and construction of externally bonded
FRP systems for strengthening concrete structures, ACI 440.2R-08, American
Concrete Institute (ACI), Farmington Hills, MI, 76 p.
2. Canadian Standards Association (CSA). 2002. Design and construction of building
components with fibre-reinforced polymers. CAN/CSA S806-02, Mississauga,
Ontario, 177 p.
3. Ross, C.A., Purcell, M.R., and Jerome, E.L. 1997. Blast response of concrete beams
and slabs externally reinforced with fiber reinforced plastics (FRP). Structures
Congress Proceedings: Building to Last, 1: 673-677.
4. Buchan, P.A., and Chen, J.F. 2007. Blast resistance of FRP composites and polymer
strengthened concrete and masonry structures A state-of-the-art review. Composites:
Part B, 38(5-6): 509-522.
5. Silva, P.F., and Lu, B. 2007. Improving the blast resistance capacity of RC slabs with
innovative composite materials. Composites: Part B, 38(5-6): 523-534.
6. Razaqpur, G.A., Tolba, A., and Contestabile, E. 2007. Blast loading response of
reinforced concrete panels reinforced with externally bonded GFRP laminates.
Composites Part B: Engineering, 38(5-6): 535-546.
7. Magnusson, J. 2007. Structural concrete elements subjected to blast loading. Licentiate
thesis, KTH, Department of Civil and Architectural Engineering, Stockholm, Sweden.
8. Fyfe CO. LLC. 2009. Tyfo SCH-41S-1 Composite Product Data. Manufacturer
material data sheet.
9. Lloyd, A. 2010. Performance of reinforced concrete columns under shock tube induced
shock wave loading. Masters thesis, University of Ottawa, Ottawa, Ontario.
10. Teng, J.G., and Chen, J.F. 2009. Mechanics of debonding in FRP-plated RC beams.
Proceedings of the Institution of Civil Engineers, Structures and Buildings, 162(SB5):
335-345.
11. Biggs, J.M. 1964. Introduction to Structural Dynamics. McGraw-Hill, NY, NY, 341 p.

484

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

MECHANICAL PROPERTIES AND DURABILITY OF


SUSTAINABLE FRP COMPOSITES
P.V. Vijay1, H.V.S. GangaRao2 and M.P.K. Reddy3
1

Research Assistant Professor, CFC, Dept. of Civil & Env. Engrg., West Virginia University, USA
Professor, CFC, Dept. of Civil & Env. Engrg., West Virginia University, USA
3
Graduate Research Assistant, CFC, Dept. of Civil & Env. Engrg., West Virginia University, USA
2

ABSTRACT
In the recent years, automobile, construction and other infrastructure industries are
considering natural fiber reinforced (FRP) and recycled composites as potential candidates
for structural and non-structural applications such as dashboard panels and house decking
planks. Use of natural and recycled FRP materials provide environmental benefits by
reducing: energy consumption, green house and hazardous gas emissions, water
consumption, contamination of soil, water, air, and solid wastes, thus contributing to
sustainable solutions. This paper describes the manufacturing process, durability, and
mechanical property evaluation of natural (flax) fiber reinforced and recycled polymer
composites with the potential for infrastructure applications.
1. INTRODUCTION
Natural fiber reinforced polymer composites can be used as alternatives to organic glass
fiber reinforced composite materials in infrastructure and automotive industries because of
their excellent mechanical properties. Natural cellulose fibers can be used as reinforcements
for composites due to many advantages such as annual renewability, sustainability, low
cost, high specific modulus, light weight, biodegradability, and biocompatibility [01-4].
Based on the literature, it is evident that success of natural fiber composites for structural
applications depends on several factors including fiber grade, processing parameters, sizing
treatment and resin compatibility.
Similar to the consideration of natural fiber composites for infrastructure applications,
recycled polymers are drawing equal attention for developing light-duty structural and nonstructural products. In the US, use of recycled plastics started gaining momentum since
1980s. Such momentum is attributed primarily to growing environmental concerns and
high-volume use of landfills (~35%) by discarded plastics contributing to health hazards
and decreasing landfill capacities.

485

2. PROPERTIES OF NATURAL AND E-GLASS FIBERS


Natural Fibers used in composites are mostly derived from plant fibers such as flax, kenaf,
jute, hemp, abaca rattan and sisal with good to excellent specific properties. Among the
natural fibers, high grade flax fibers mechanical properties are nearly on par with organic
fibers such as glass used as composite reinforcement [5-9]. E-glass fibers are the most
economical among organic fibers that include carbon and aramid. They are available in
forms such as mat, woven, fabric, chopped strands etc. [10]. Table 3 lists some of the
properties of E-glass and natural fibers.
Table 3. Comparison of mechanical properties of fibers [10, 11]
Density
Tensile strength
Youngs
Elongation
Fiber type
(g/cm3)
(MPa)
modulus (GPa)
(%)
E-Glass
2.5
2000-3500
70
2.5
Bamboo
0.6-1.1
140-230
11-17
Flax
1.5
345-1035
27.6
2.7-3.2
Hemp
1.48
690
70
1.6
Jute
1.3
393-773
26.5
1.5-1.8
Kenaf
930
53
1.6
In this study, natural flax obtained in the fiber and fabric form was used for manufacturing
coupon specimens. These natural fiber/fabrics are available in varying grades of strengths
and stiffnesses. Natural fibers require further processing with appropriate sizing for better
compatibility with resin and improvement in shear strength. Cellulose fibers (natural fibers)
have excellent bonding properties with phenolic resins. Hence, phenolic resins are used as
matrix materials in this investigation. Properties of natural fiber composites are also
compared with those of recycled glass fiber composites. Optimization of natural
fiber/fabrics for their mechanical properties is beyond the scope of this work.
3. MANUFACTURING OF NATURAL (FLAX) FRP COMPOSITE
Natural fiber reinforced composite laminates (Fig. 1) were manufactured using compression
molding with flax fabric as reinforcement and phenolic resin as matrix. Composite
laminates measuring 0.254 m x 0.381 m (10 in x 15 in) were manufactured with different
fiber volume fractions. Before manufacturing, the flax fabric was cut to required
dimensions of the mold unit and oven dried at 48.890C (1200F) for about 6 hours to remove
any moisture content. Phenolic (thermoset) resin was mixed thoroughly with paraformaldehyde (compatible curing agent) in the ratio of 5:1 by weight for wetting the natural
fabrics. Top and bottom plates of the mold were covered with aluminum foil and coated
with demolding agent to facilitate easy release of the molded product.
Each layer of flax fabric was coated with phenolic resin and stacked sequentially on top of
each other in the mold. After positioning each layer of resin coated fabric in the mold, they
were pressed with a manual roller squeeze excess resin and minimize the voids. These
stacked layers of fabrics were placed against top and bottom plates of the mold were
compression molded under a load of five tons at 48.89-65.560C (120-1500F) for 2 hours.
486

The heat was turned off for next 14 hours without releasing the applied pressure to
eliminate laminate warping effects.

Fig. 1. Flax fabric and flax-phenolic fiber reinforced composite laminate


4. TESTING: SPECIMENS AND PROCEDURES
The composite laminates were manufactured with two types of fabrics (Flax1 and Flax2)
with different fiber volume fractions (Table 4) in order to compare their mechanical
properties. Flax1 and flax2 were obtained from different manufacturers. Flax1 fabric was
bleached and Flax2 with a better finish was non-bleached. The laminates were cut to
required dimensions for performing tension, bending, impact and moisture absorption tests.
Moisture absorption specimens were coated with resin along the cut edges to prevent easy
water ingress at those locations.
Table 4. Details on configuration of test composites
Types of fabric

Resin

No. of layers of fabric


5
10
15
20
10
10
10

Flax1
Cascophen phenolic
Flax2
E-glass
E-glass

Derakane vinyl ester

Fiber volume
fraction
0.368
0.355
0.304
0.321
0.461
0.402
0.589

4.1 Fiber and Composite Testing and Instrumentation


Individual flax fiber strands separated from the natural fabrics with diameters of (383-390
m) were embedded in end-grips and tested for their tensile strength. A fiber can be defined
as an elongated material having a more or less uniform diameter or thickness less than 250
m and an aspect ratio greater than 100 [12]. Results of fiber testing are shown in Table 5.
Results indicate that the strands have lower strength (134-196 MPa, Table 5) than a typical
low grade flax fiber (345-1035 MPa, Table 3). Strands are known to achieve lower strength
than fibers due to additional fiber processing, surface modifications (flaws) and orientation
(e.g., change in alignment due to weaving, twisting etc.).

487

Table 5. Results for tension test on flax fibers


Type of fiber
Diameter of the fiber(m)
Stress (MPa)
Flax1
389.88
134
Flax2
383.44
196
Tension coupon specimens measuring 0.247 m x 0.0127 m (9.75 in x 0.5 in) were tested in
an Instron Universal Testing Machine (Fig. 2) as per ASTM D3039/D3039M-08. Bending
coupon specimens measuring 0.127 m x 0.0127 m (5 in x 0.5 in) were tested in an Instron
Universal Testing Machine (Fig. 3) as per ASTM D790-10. Impact coupon specimens
measuring 0.0635 m x 0.0127 m (2.5 in x 0.5 in) were tested in an Instron Universal
Testing Machine (Fig. 4) as per ASTM D256-10. Results of the tension, bending, and
impact strengths are shown in Table 6.

Fig.2. Tension test on natural fiber reinforced composite coupon

Fig.3. Flexural test on natural FRP composites

Fig.4. Impact test machine and coupons after impact test.

488

5. MOISTURE ABSORPTION AND DURABILITY


Coupons measuring 0.152 m x 0.0127 m (6 in x 0.5 in) were immersed in water for 360
hours in order to perform moisture absorption test. Weight gains were periodically
monitored. Moisture absorption specimens are shown in Fig. 5 and the results are plotted in
Fig. 6.

Fig.5. Moisture absorption test


6. RESULTS AND DISCUSSION
6.1 Tensile Test on Natural Fiber Reinforced Polymer Composite
Tensile and bending strengths of natural composites with Flax1 and Flax2 type fabrics were
found to be varying from 56-75 MPa and 51-93 MPa, respectively. Similarly, impact
strengths for Flax1 and Flax2 natural composites were found to be 0.441-1.86 Joules and
1.173 Joules, respectively. Fiber volume fraction of Flax2 composite was higher than that
of Flax1 (Table 6).
Table 6. Results for tension, bending and impact test coupons
Details of
fabric

Flax1

Flax2
E-glass
E-glass

Resin

Cascophen
phenolic

Cascophen
phenolic
Cascophen
phenolic
Derakane
vinyl ester

No. of
layers
of fabric

Fiber
volume
fraction

5 layers

0.368

10
layers
15
layers
20
layers
10
layers
10
layers
10
layers

0.355
0.304
0.321
0.461
0.402
0.589

489

Tensile
stress
MPa
(ksi)
74.95
(10.87)
72.26
(10.48)
72.95
(10.58)
72.33
(10.49)
56.26
(8.16)
306.95
(44.52)
376.32
(54.58)

Flexural
stress
MPa
(ksi)
93.49*
(13.56)*
78.05
(11.32)
73.43
(10.65)
87.84
(12.74)
51.71
(7.50)
-

Impact
strength
Joules
(ft-lbs)
0.441
(0.325)
0.908
(0.67)
1.586
(1.17)
1.86
(1.37)
1.173
(0.865)
2.481
(1.83)
-

*Test incomplete due to specimen slide-out.


6.2 Moisture Absorption Test on Natural and GFRP Composites

Percentage of
moisture
absorption

Moisture absorption tests indicated that the initial values of moisture gains could exceed
10% in some of the natural composite specimens within 24 hours. These high values
indicate the effect of cutting on the specimens and subsequent creation of moisture
transport channels (fiber/fabric interface and exposed fibers). High amount of moisture
absorption indicates that the compression molding process may require further refinement
or use of secondary operation such as surface coating to control moisture ingress. The
results in Fig. 6 show the moisture absorption values after 24 hours. Moisture absorption up
to 5.74% was noted between 24-360 hours. Glass-phenolic composites showed about half
the moisture absorption of flax-phenolic composites between 24-360 hours.
8
6
4
2
0

5Layers(Flax1)

200

400

15Layers
(Flax1)

Numberofhoursofmoistureabsorption
Fig.6. Moisture absorption in different FRP specimens
6.3 Mechanical Properties of Recycled ABS and Glass FRP
As demonstrated by CFC-WVU [10], thermoplastics from post consumer, post industrial,
and electronic shredder residue (ESR) can be recycled to manufacture consumer products
and structural components. Specifically, ABS polymers were reinforced with 4%, 7% and
12% fiber volume fractions of bi-directional glass fabric (fabric density: 407 gm/m2 or 12
oz per square yard) and were evaluated.
Table 5. Tensile stress and modulus of recycled ABS with GFRP fabrics
Fabric Avg.
Per layer
Increase in
Tensile Per layer
Increase in
layers stress
avg. stress
stress (per
mod.
mod.
mod. (per layer
(fvf)
increase
layer increase)
increase mod. increase)
MPa
MPa
GPa
GPa
0
28.3
-------2.1
------1
34.5
6.2
22%(22%)
3.2
1.1
49% (49%)
3
59.0
10.2
108%(36%)
4.7
0.8
117%(39%)
5
72.2
8.8
155%(31%)
5.3
0.6
149%(30%)
Note: fvf (fiber volume fraction). Tensile strength results were in agreement with rule of
mixtures (ROM) commonly used in composites.

490

Tensile stress of 72.2 MPa in GFRP reinforced ABS with only 0.04 fvf was found to be
better than natural reinforced FRP having 74.95 MPa with 0.368 fvf. These values can be
further improved with the use of high-grade natural fibers and recycled materials, along
with improved adhesion (sizing) between fibers and resin.
7. CONCLUSIONS
The mechanical properties with low-grade natural fiber reinforced composites are about
half of the organic glass fiber reinforced composites with phenolic resins (by normalizing
to equal fiber volume fractions). GFRP composite with vinyl ester resins carried 24% more
stress than GFRP with phenolic resins. Further enhancements in mechanical properties can
be achieved using high grade natural fabric, compatible sizing and minimizing voids and
better manufacturing processes e.g., pultrusion with better curing and process control as
opposed to compression molding for better curing and void control. Natural FRP
composites exhibited good impact because of their high damping properties. Moisture
absorption was noted to be higher in natural composites than glass-phenolic composites.
Natural and recycled composites have the potential to be used as alternate materials for
many non-structural or low-end structural applications. Gradually, their use can be
expanded to demanding structural applications.
8. REFERENCES
1.
2.
3.
4.
5.
6.
7.
8.
9.

Mohanty, A.K., Mishra, M., and Hinrichsen, G. 2000. Biofibers, Biodegradable


Polymers and Biocomposites: An Overview. Macromal. Mter. Eng. 276-277: 1-24.
Dahlke, B., Larbig, H., Scherzer, H.D., Poltrock, R. 1998. Natural Fiber Reinforced
Foams Based on Renewable Resources for Automotive Interior Applications. J. Cell.
Plast. 34: 361-379.
Bledzki, A.K. and Gassan, J. 1999. Composites Reinforced with Cellulose Based
Fibers. Prog. Polym. Sci. 24: 221-274.
Wambua, P., Ivens, J., and Verpoest, I. 2003. Natural Fibers: Can They Replace Glass
in Fiber Reinforced Plastics? Compos. Sci. Technol., 63: 1259-1264.
Kim, J.T. and Anil N.N. 2010. Mechanical, Thermal, and Interfacial Properties of
Green Composites with Ramie Fiber and Soy Resins. J. Agric. Food Chem., 58: 54005407.
Bhyrav, M., Aktas, C., Joe, M., Bilec, M., and Hota, G. 2010. Natural Fiber Reinforced
Pultruded Composites. Composites 2010. American Composites Manufacturers
Association. 9-11 February.
Drazl, L.T., Mohanty, A.K., Burgueno, R., and Mishra, M. Biobased Structural
Composite Materials for Housing and Infrastructure Applications: Opportunities and
Challenges. (www.pathnet.org/si.asp?id=1076)
Mueller, D.H. and Krobjolowski, A. 2003. New Discovery in the Properties of
Composites Reinforced with Natural Fibers. J. Industrial textiles, 33(2): 111-130.
Gaceva, G.B., Avella, M., Malinconico, M., Buzarovska, A., Groxdano, A., Gentile,
G., and Ericco, M.E. 2007. Natural Fiber Eco-Composites. Poly. Composites, pp. 98107.

491

10. Joshi, S.V., Drzal, L.T., Mohanty, A.K., and Arora, S. 2004. Are Natural Fiber
Composites Environmentally Superior to Glass Fiber Reinforced Composites?
Composites Part-A, 35: 371-376.
11. Vijay, P.V., Hota, H.V.S., and Siddalingesh, K. 2011. Recycled Polymers for
Sustainable Solution in Infrastructure. SPE, Antec-2011, May.
12. John, M.J. and Rajesh, D.A. 2008. Recent Developments in Chemical Modification
and characterization of Natural Fiber-Reinforced Composites. Polymer Composites,
pp. 187-207.
13. Chawla, K.K. 1988. Fibrous Materials. Cambridge University Press, 9 p.

492

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

DURABILITY STUDY OF BASALT FIBER EPOXY WET LAYUPS


USED IN REHABILITATION
G. Xian, C. Wang, B. Xiao, H. Zhang and H. Li
School of Civil Engineering, Harbin Institute of Technology (HIT), Harbin 150090, China

ABSTRACT
In the present study, unidirectional basalt fabric reinforced epoxy plate was fabricated with
wet layup technique. Its durability performances were investigated regarding to the harsh
environments of water/alkaline solution immersion and elevated temperatures. After one
year immersion, the tensile properties of basalt fiber reinforced polymer (BFRP) deteriorate
dramatically, especially for alkaline solution immersion. Water uptake curves and scanning
electron micrograph (SEM) analysis indicates that the resin, fiber and bonding between
fibre and epoxy were weakened by water ingression, responsible for the tensile property
deterioration. BFRP wet layup strips, exposed to 100, 150 and 200oC for up to 72 hours,
shows a negligible change in tensile modulus, while the strength degraded obviously.
Finally, BFRP wet layups were tested in tension properties directly at elevated
temperatures. As found, when the testing temperature excesses the glass transition
temperature (~80oC), the tensile strength is reduced by 45% of its initial values, while the
modulus decreases linearly with testing temperatures up to 300oC.
1. INTRODUCTION
As a novel reinforcement, basalt fibers are produced directly from basalt rock through a
melting process with high mechanical properties and chemical resistance [1]. Since BFRP
has been applied in civil engineering only for several years, the comprehensive knowledge
of BFRP from the basic physicochemical properties to the long term durability in various
civil environments is far from completely understood yet, which is prerequisite for its
acceptance for the acceptability by civil engineering. In the past several years, some works
have been conducted on the performance and application of BFRP composites in structural
strengthening, rehabilitation etc. [1-5]. Some contradictory results were reported, which
may due to the fact that the properties of basalt fiber varied from each other seriously,
dependent on the basalt rock mines and producing process. Without consistent and reliable
property data of BFRP, as believed, the real civil engineering practice with BFRP will be
hazarded, especially for some key infrastructures.
In view of this, the present work is focused on a thorough performance investigation of
commercial BFRP composites. The effects of water immersion, alkaline immersion as well
493

as elevated temperatures on the performance degradation of BFRP wet layups are explored.
The study is aimed to evaluate the possible advantage and withdraw of the BFRP when
applied in civil engineering.
2. EXPERIMENTAL
2.1 Materials
An epoxy resin system used in the present study can be cured at room temperature, and
reaches a high deflection temperature (>80oC). Unidirectional basalt fiber fabrics are
commercially available in China. As indicated, the fiber diameter is 13 micron; the area
density of the fabric is 360 g/m2; tensile strength is 2.1GPa and modulus is 105GPa. BFRP
samples prepared with a hand wet layup process. Two layers of fabrics were used for a
sample with the dimensions of 300 x 300 mm2, or 600 x 300 mm2 for elevated temperature
testing. The nominal thickness is 0.8 mm.
2.2 Test procedures
Tensile test was performed for BFRP wet layups, according to ASTM D3039. The tensile
speed is adopted as 5 mm/min. Testing results are reported with 5 replicates for each
condition. Immersion ageing of BFRP wet layup strips were performed with water bath at
23, 40, 60 and 80oC. The water uptake testing is described in the previous paper [6]. The
elevated temperature performance of BFRP specimens were divided two sets. One set was
tested at elevated temperatures, while the other set was tested after elevated temperature
treatment for various pre-determined periods. The detailed testing process can be seen in
[7].
3. RESULTS AND DISCUSSION
3.1 BFRP Wet Layups Immersed in Water and Alkaline Solution
In immersion conditions, the water uptake curves are plotted as a function of a square root
of immersion time (Fig. 1) for both water and alkaline solution immersion. It is seen that
the water uptake behaviour clearly deviates from the classic Fickian diffusion mode, at low
temperatures, e.g., less than 80oC. The water uptake continuously increases rather than
levels off as the Fickian diffusion mode [8] with the square root of the immersion time after
the first linearly water uptake stage. All specimens did not reach water uptake saturation in
12 months (Fig. 1 a&b). Those characteristics are frequently reported for the water or
moisture diffusion in polymers or polymer composites. As proposed, the first stage of water
uptake is impelled by the water concentration gradient (Fickian diffusion mechanisms) and
the second stage is related to the relaxation of polymer molecular structures due to water
ingress [8, 9].

494

a)
b)
Fig. 1. Water uptake as a function of square root of immersion time for BFRP wet layups
immersed in distilled water (a) and alkaline solution (b).
It is worth noting that the system shows abnormal water uptake trends at 80oC, which is
believed coming from the hydrolysis of the polymer matrix as well as the degradation of
the interface between the fiber and the resin matrix. The SEM photograph of fractured
samples clearly indicates the degradation of fiber, resin and the interface between fiber and
resins (Fig. 2). At higher temperature, e.g., 80oC, even for a short immersion period (i.e., 1
month), the very smooth fiber seen from the tensile fracture surface is indicative of the
significant deterioration of the interface between fiber and resin. Two types of degradation
of fibers are found. One is the etching of fiber surfaces for both water and alkaline solution
immersion. One another degradation mode of the fiber is found for water immersion cases,
where the surface layer in micron size of the fiber were delaminated from the core part. The
degradation of the basalt fiber subjected to alkaline solution is also reported by Sim [1].

a)
b)
c)
Fig. 2. SEM pictures of tensile fracture surface of various BFRP samples, subjected to
water or alkaline immersion: a) Water immersion at 80oC for 1 month; b) Water immersion
at 80oC for 1 year; cAlkaline immersion at 80oC for 1 year.
According to Bao et al [9], for two stage features, the water uptake (Mt) at time t can be
approximately written as
M t M (1 k t )(1 exp[7.3( Dt / h 2 ) 0.75 ])
(1)
where k is the time dependent coefficient, characterizing the rate of polymer relaxation due
to water ingress; M, the equilibrium water content; D is the diffusivity at Fickian stage; h
the specimen thickness; t the immersion time. In the present study, the parameters related to
495

water absorption are determined by curve fitting using Equation 1. A nearly perfect fit was
achieved for all water uptake curves. The determined parameters are tabulated in Table 1.
All presented diffusivities of the composites were not edge-corrected.
Table 1. Pseudo equilibrium water content (M), maximum water content (Mmax),
diffusivity as well as k corresponding to best fit of the two stage diffusion model.
Temp. (oC) Medium
M [%]
Mmax [%]
D [10-7 mm2/s]
k [x 10-5 s-1]
Water
0.73
1.31
0.43
14.1
23
Alkaline
0.88
1.73
0.40
17.7
Water
1.10
1.85
1.76
13.1
40
Alkaline
1.18
2.53
1.51
21.5
Water
1.29
2.33
6.85
32.32
60
Alkaline
1.18
3.49
8.30
37.94
Water
2.46
2.80
5.01
4.46
80
Alkaline
1.58
4.73
27.6
30.68
The equilibrium water uptake (M) represents the saturation water content if there is not the
second water absorption stage. Except for the immersion case in water at 80oC, the M is
closed for all specimens. As known, the saturation water content for a resin is related to the
polar groups as well as the free content. For moderate polar polymer, such as epoxy, the
effect of the temperature has a very litter effect on M [10].
For immersion in both water and alkaline solution, the higher the temperature, the higher D
is realized. This is consistent with previous works [11, 12]. The activation energy of the
diffusion process calculated according to Arrhenius relationship equation is 61.3 kJ/mol for
water immersion and 65.6 kJ/mol for alkaline immersion. Those values are supported by
our previous study with a value in the order of 60~70 kJ/mol for a epoxy system in water
immersion [6].
3.2 Mechanical Properties as a Function of Immersion Time
Tensile properties of BFPR in fiber direction were tested at various immersion periods, and
the results are shown in Fig. 3 and 4 for water immersion and alkaline immersion,
respectively. The tensile strength was dramatically deteriorated for both cases, while the
tensile modulus was less affected adversely.
Note such degradation level is much higher than the literature values for GFRP or CFRP
[13-16]. The discrepancy is coming from the thickness of the specimens. As shown in the
experimental part, the current studied specimen is only about 0.8 mm, much less than the
samples in the literature, generally in the range of 3 to 6 mm. The significant deterioration
of basalt fiber at high temperature (see in Fig. 2) is believed to be the main reason.

496

a)
b)
Fig. 3. Tensile strength (a) and modulus (b) as a function of immersion time in water for
BFRP wet layups.

a)
b)
Fig. 4. Tensile strength (a) and modulus (b) as a function of immersion time in alkaline
solution for BFRP wet layups.
3.3 Influence High Temperature Treatment
May due to the effect of postcuring, after a short term of exposure to elevated temperatures,
the tensile strength of BFRP strips increased (Fig. 5). At 200oC, despite of fluctuation, the
trend of the tensile strength with the exposure time shows a clear decrease. After 720 hours
exposure, the tensile strength decreased by more than 17%. BFRP did not show change in
tensile strength at 100oC, and a reduction less than 5% at 150oC. The above testing results
are consistent with that reported by Foster on CFRP and GFRP [17].

a)
b)
Fig. 5. Variation of tensile strength (a) and modulus (b) of BFRP wet layups after exposure
to elevated temperatures.

497

On the contrary, the tensile modulus of the BFRP is less affected by exposure temperature
and time (Fig. 5b). As shown, after 720 hours exposure to elevated temperatures, the
decrease of the tensile modulus is less than 7%. It is worth noting that the high temperature
exposure can lead to postcuring of BFRP and thermal degradation of polymer matrix. With
both effects, the modulus does not change with the time monotonic.
3.4 High Temperature Performance of BFRP
Fig. 6a presents the tensile strength as a function of testing temperatures. As shown, when
the temperature excesses Tg (79.9oC), the tensile strength reduced by 53%, followed by
leveling off with further increase of temperatures. When the temperature ranging from Tg to
Td, the polymer is in the high elastic plateau, and show less changes with temperatures in
the mechanical properties. Therefore, the tensile strength of BFRP shows relatively stable
in this temperature stage (Fig. 6a). Fig. 6b gives the relative modulus of the BFRP strips
with testing temperature. As shown, with the temperature increasing up to 300oC, the
relative modulus shows decrease monotonically. This is also can be attributed to the
softening of the resin at elevated temperatures.

a)
b)
Fig. 6. Variation of tensile strength (a) and relative modulus (b) of BFRP wet layups as a
function of exposure temperature
4. CONCLUSIONS
According to the current study on BFRP wet layup composites, the following conclusions
can be drawn.
Two stage diffusion model describe the water uptake of the BFRP wet layups
immersed in both water and alkaline solution well, except for the case of 80oC water
immersion due to remarkable weight loss.
Water and alkaline immersion leads to significant deterioration of the resin, fiber as
well as the interfaces. As a result, the tensile strength was seriously reduced after one
year immersion. Alkaline solution shows more corrosive effect on the BFRP wet
layups.
After exposure to 100 and 150oC for up to 720 hours, the tensile strength and modulus
of BFRPs change slightly. However, as the temperature exceeding 200oC, about 17%
reduction in the tensile strength is resulted.
When the testing temperature excesses Tg of BFRP, the tensile strength was reduced by
45% of its room temperature value. With further increase of the testing temperature by
498

300oC, the tensile strength shows less dependence on the temperatures. The tensile
modulus decrease with testing temperatures monotonically. The maximum reduction is
about 35% at 300oC.
5. REFERENCES
1.
2.
3.
4.
5.
6.
7.

8.
9.
10.
11.
12.
13.
14.

Sim, J., Park, C. and Moon, D.Y. 2005. Characteristics of basalt fiber as a
strengthening material for concrete structures. Composites Part B-Engineering, 36(67): 504-512.
Cerny, M. Glogar, P., Golias, V., Hruska, J., Jakes, P., Sucharda, Z., and Vavrova, I.
2007. Comparison of mechanical properties and structural changes of continuous
basalt and glass fibres at elevated temperatures. Ceramics-Silikaty, 51(2): 82-88.
Deak, T. and Czigany, T. 2008. Investigation of basalt fiber reinforced polyamide
composites. Materials Science, Testing And Informatics IV, 589: 7-12.
Mingchao, W., Zuoguang, Z., Yubin, L., Min, L., and Zhijie, S. 2008. Chemical
durability and mechanical properties of alkali-proof basalt fiber and its reinforced
epoxy composites. Journal Of Reinforced Plastics And Composites, 27(4): 393-407.
Tang, Y., Wu, Z., Yang, C., Shen, S., Wu, G., and Hong, W. 2009. Development of
self-sensing BFRP bars with distributed optic fiber sensors. Proceedings of the SPIE The International Society for Optical Engineering, Proceedings Vol. 7293, 10 p.
Xian, G.J. and Karbhari, V.M. 2007.DMTA based investigation of hygrothermal
ageing of an epoxy system used in rehabilitation. Journal Of Applied Polymer
Science, 104(2): 1084-1094.
Wu, J., Li, H., and Xian, G. 2010. Influence of elevated temperature on the
mechanical and thermal performance of BFRP rebar, Proceedings of the 5th
International Conference on FRP Composites in Civil Engineering (CICE 2010),
Lieping Ye, Peng Feng and Qingrui Yue, Editors, pp. 69-72.
Bond, D.A. 2005. Moisture diffusion in a fiber-reinforced composite: Part I - NonFickian transport and the effect of fiber spatial distribution. Journal Of Composite
Materials, 39(23): 2113-2141.
Bao, L.R., Yee, A.F., and Lee, C.Y.C. 2001. Moisture absorption and hygrothermal
aging in a bismaleimide resin. Polymer, 42(17): 7327-7333.
Pascault, J.-P., Sautereau, H., Verdu, J., and Williams, R.J.J. 2002. Thermosetting
Polymers., Marcel Dekker, New York, 477 p.
Zhou, J.M. and Lucas, J.P. 1999. Hygrothermal effects of epoxy resin. Part I: the
nature of water in epoxy. Polymer, 40(20): 5505-5512.
Chin, J.W., Nguyen, T., and Aouadi, K. 1999. Sorption and diffusion of water, salt
water, and concrete pore solution in composite matrices. Journal Of Applied Polymer
Science, 71(3): 483-492.
Mourad, A., Abdel-Magid, B., El-Maaddawy, T., and Grami, M. 2010. Effect of
Seawater and Warm Environment on Glass/Epoxy and Glass/Polyurethane
Composites. Applied Composite Materials, 17(5): 557-573.
Liao, K., Schultheisz, C.R., and Hunston, D.L. 1999. Effects of environmental aging
on the properties of pultruded GFRP. Composites Part B-Engineering, 30(5): 485493.

499

15.
16.
17.

Guan, H., Karbhari, V.M. and Sikorsky, C.S. 2007. Long-term structural health
monitoring system for a FRP composite highway bridge structure. Journal Of
Intelligent Material Systems And Structures, 18(8): 809-823.
Abanilla, M.A., Li, Y., and Karbhari, V.M. 2006. Durability characterization of wet
layup graphite/epoxy composites used in external strengthening. Composites Part BEngineering, 37(2-3): 200-212.
Foster, S.K. and Bisby, L.A. 2008. Fire survivability of externally bonded FRP
strengthening systems. Journal of Composites for Construction, ASCE, 12(5): 553561.

500

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

COMPARATIVE DURABILITY ANALYSIS OF CFRP


STRENGTHENED RC HIGHWAY BRIDGES
O. Ali1, D. Bigaud2 and E. Ferrier3
1

Ph.D student, LASQUO Laboratory, Angers University, Angers, France.


Professor, LASQUO Laboratory, Angers University, Angers, France.
3
Professor, LGCIE Site Bohr Laboratory, Lyon 1 University, Lyon, France.
2

ABSTRACT
The paper presents parametric analysis of durability factors of RC highway bridges
strengthened with CFRP laminates. Durability factors considered are concrete cover and
CFRP laminates thickness. Three deterioration factors were considered. First, growth of
live load with time. Second, resistance losses due to steel corrosion. Pitting and uniform
corrosion were studied.Third, deterioration due to CFRP aging. Monte-Carlo simulation
was used to develop time dependent Statistical models for steel area and live load extreme
effect. Reliability is estimated in term of reliability index using FORM method. For
illustrative purpose, reliability of RC bridge girder is evaluated under various traffic and
corrosion environments. Bridge design options follow AASHTO specifications. The
present work extends to calibrate CFRP resistance safety factor which corresponds to two
target reliability levels, T=3.5 and 4.2. The results show that, corrosion has the most
significant effect on bridge life time especially pitting type. Initial reliability index is traffic
dependent. AASHTO design equation (load factors corresponds T =3.5) seems to be
overestimated for strengthening purpose. Strengthening with (T =4.2) provide better
reliability than T proposed by AASHTO with no significant differences in CFRP amounts
required.
1. INTODUCTION
Reinforced concrete bridges safety is strongly affected by many deterioration factors. The
most important factor is reinforcement corrosion. Considerable research effort had been
done to evaluate corrosion effects. It has been achieved that, corrosion causes reduction of
steel properties [1], losses of bond between concrete and steel rebars, cracking, and spalling
of concrete cover [2]. Another factor of great importance is live load growth with time.
Additional resistance through a certain strengthening technique is required to compensate
losses in section resistance. Recently, carbon fibre reinforced plastic CFRP materials were
used in strengthening purpose; Laminates are externally bonded to extreme concrete tensile
surface to produce additional bending strength. In the last two decades, flexural behaviour

501

of CFRP strengthened RC beams had been studied in significant number of studies.


Different failure modes were observed, which can be classified in two types according
location. First, failure modes occur at position of maximum moment. Such modes are
concrete crushing, CFRP mid span debonding, and CFRP rupture. Second, failure modes
occur at plate end. Herein, failure modes of the second type were neglected as they can be
omitted using anchorage system. In an early study proposed by Plevris et al [3], the authors
suggested a specific reduction factor (CFRP = 0.8) for the CFRP contribution to
strengthened element resistance. However, the study was limited to CFRP rupture failure
mode. Recently, Pham and Al-Mahaidi [4] discusse the reliability of strengthened beams
considering all failure modes of first type. Atadero and Karbhari [5] studied the reliability
of CFRP strengthened RC beams considering steel corrosion and CFRP durability. The
authors focused on the importance of CFRP statistic properties on CFRP safety factor. The
present work aims to compare the effect of durability factors - concrete cover and CFRP
laminates thickness - on girders service life.
2. CORROSION OF REINFORCEMENT BARS
Corrosion mechanism can be divided into three stages; initiation, propagation before
cover cracking, and after cover cracking. This classification corresponds to the variation
in corrosion current at each stage. Corrosion current was evaluated according to
corrosion model detailed in [6]. Table 1 presents probabilistic environmental,
geometrical, and material variables used in the study.
Table 1. Probabilistic parameters of random variables.
Variable.
fc
ct
fy
ds
As,intial
c
Cs
Cth
Dcl,
lpit
icurrent
d
Ef
d
ff,u
Dead load
(concrete)
Asphalt
thickness
Truck load
w
Impact
factor (IL)
Model
error m

distribution
Normal
Normal
Normal
Normal
Normal
Lognormal
Lognormal
Uniform
GEV
Uniform
Normal
Lognormal
Weibull
Normal

Units
nominal
MPa
34.5
--0.3
MPa
414
mm
700
mm2
490.87
mm
25-45
kg/m3
3.0
kg/m3
0.9
mm2/yrs 1
mm
60
--GPa
51.7
MPa
620.5
depend on analysis

Biasa(mean)
1
1
(460)
(700-4.7)
0.97
1
1
1
54.65
1
1
1
1
1.05

covb(stdc)
0.15
0.15
(46.2)
(12.7)
0.024
0.2-0.1
0.5
0.19
0.35
0.385
0.2
0.2
0.15
0.1

Description [source]
Concrete compressive strength [3]
concrete tensile strength randomness [7]
Steel yield strength [9]
Steel rebars depth [3]
Initial steel rebars area[5]
concrete cover [6]
Surface chloride concentration [9]
chloride concentration threshold value [6]
chloride diffusion coefficient [9]
Pit length[7]
Uncertainty in corrosion current [9]
CFRP modulus [5]
CFRP strength [5]
[7]

Normal

mm

90

0.25

[7]

Normal

kN

250

0.4

[7]

Normal

---

0.1

0.8

[7]

--1
See Fig. 3
[3]
Concrete crushing: Lognormal, CFRP debonding: Lognormal, and CFRP: rupture normal.

a
Bias; mean/nominal, bcov; coefficient of variation, cstd standard deviation. d property decorrelated from CFRP
thickness

502

Reduction in the diameter of the corroded bar D in (mm) was evaluated according to the
following equation [7]:
t

D 0.0232 icorr (t )dt

(1)

t ini

Eq. (1) is valid for losses due to uniform corrosion as losses are uniform along bar length.
Pitting corrosion losses concern particular small areas on bar surface. Val et al [7]
distinguish two types of pitting corrosion along bar length based on in site measurements of
repaired bridges: coarse pits which are characterized by length lp ranges between 20-100
mm and fine pits (lp: 2-4 mm). The authors assumed that, reinforcing rebar pit area is
constant over its length, thus losses in rebar diameter equivalent to,
(2)

D [ Do2 D ( 2 Do D )lm / l p ] / 4

where lp is the length of the pit, and lm (=105mm) is the bar length of site measurement of
corrosion current (see Fig. 1a). The choice of lm value corresponds to maximum pit length
observed in the survey [7]. It is worth mentioning that corrosion did not affect steel area
only, but its actions also change yield strength with time. It can be assumed that, yield
strength is linearly proportional to the reduced steel area As(t) such that [8].
f y (t ) (1 y As (t ) / Aso ) f yo

(3)

where, fyo and Aso are steel initial yield stress and area respectively, and y is an empirical
coefficient which taken equal to 0.005 according to recommendation given Stewart and AlHarthy [8], Monte-Carlo random number generation algorithm was used to provide data set
(5x105 cases) based on statistical corrosion variables in Table 1. Each case was simulated
with the corrosion model to provide time dependent steel area As(t) data set. The model is
served by monthly temperature and relative humidity profiles in the ranges (2040C) and
(0.60.75) respectively. Various corrosion environments were simulated (c=25, 30,35
65mm). Fig. 1b and 1c shows results for concrete cover: c=45mm. An extreme value of
chloride surface concentration is considered: Cs=3. As shown in Fig. 1 steel area density
function changes from normal distribution with two parameters N(,) at initial age - to
bimodal distribution which can be fitted with five parameters (1,1,2,2,). Fitting
algorithm was focused on a range of As(t) data corresponds to a range [-2.5, 0.5] in the
standard normalized space. The choice of this range depends on the contribution of steel
bar area uAs in reliability index . Reliability analysis gives: uAs=-2.10 of , which satisfy
the considered range of fitting algorithm. Evaluation of steel damage using the proposed
simulation obviates many differential calculations required for FORM algorithm.
0.04
0.03

pdf

(a) Corroded bar configuration


lp

Do
lm
p ittin g

u n ifo rm

0.02

0.04

(b) uniform corrosion c=45mm


f As(x,t)= (t) (x, 1(t), 1(t))(1- (t)) (x, 2(t), 2(t))

: normal density function

0.02
0.01

0.01
0
350

400

450

Bar area mm2

500

gnerated
data

0
300

t=0 year
t=50 year
t=100 year

350

400

450

Bar area mm2

Fig. 1. Corroded rebar configuration and rebar area As(t) density.

503

fitted curve

(c) pitting corrosion c=45mm

0.03

500

3. LIVE LOAD MODEL


Based on analysis of data survey done by Vu and Stewart [9], truck (heavy) weight follows
normal distribution; mean w=250kN and standard deviation: w=100kN. Truck weight
distribution between axels and axel configuration were adopted to standard HS20 truck
proposed in AASHTO [10]. Impact was taken as fraction of truck weight [11]. Highway
Traffic survey studies confirm that, traffic load volume and weight increase with time. Vu
and Stewart [9] suggested time dependent mean truck weight: w(t)=w(1+m)t, standard
deviation: w(t)=w(1+m)t, and average daily truck traffic: ADTT(t)=ADTTinitial(1+ v)t.
Where v is the traffic volume annual increase: v1-3% taken 2.3%, m is the truck weight
annual increase: m=0.005, and ADTTinitial is the initial average daily truck traffic which can
be assumed as a fraction of the total traffic [10]. Extreme event (i.e., bending moment) is
affected by four variables [12]. First is truck model. Second is impact factor. Third is
variation in transverse traffic position dlane in each lane (for standard 3.6m wide lane, dlane
follows lognormal distribution with coefficient of variation 0.33 and mean value 0.9m
measured from the edge of the lane to the center line of the outermost vehicle wheel).
Fourth, multiple presence of fully correlated heavy trucks side-by-side on bridge lanes (two
lanes). According to Nowak [11] data survey, it is observed that every 30 trucks on the
bridge is accompanied simultaneously by another truck side by side. It is assumed that,
traffic growth takes place at each year increment and traffic simulation from year to the
next is statistically independent. Monte-Carlo random number generation was used to
simulate daily traffic loads pass through lanes in a way to get extreme event statistical
model. Simulation was done in three steps. First, for each day in the year, ADTT trucks per
lane randomly generated. Finite element program SAP2000 was used to determine truck
event to avoid girder distribution factor uncertainty. Day extreme event is recorded and
taken equal to the maximum of; maximum daily event in the first lane, maximum daily
event in second lane, and maximum of N summation of two events randomly chosen.
Second, extreme events obtained over the year in step 1 are averaged and their standard
deviation is calculated. Third, step 1 and 2 were repeated till variation in parameters
converge a chosen tolerances (0.01 and 0.005 for bias and coefficient of variation
respectively), 10 repetitions were found sufficient to achieve the assumed tolerances.
Generalized extreme value GEV distribution best fits the generated data. Fig. 2 presents
Live load extreme event statistics. It is noted that, live load statistical parameters depend on
time and ADTT. The model provides more accurate traffic simulation than Nowaks model,
as the author neglects the effect of ADTT and load growth.
(a)

coefficient of variation

bias

1.5
1

0.5
0

50

250

500

750

1000

(b)

time=1 yr
time=25 yr
time=50 yr
time=75 yr
time=100 yr

0.1

0.05

50

ADTT (truck/day)

250

500

750

1000

ADTT (truck/day)

Fig. 2. Live load extreme event statistics, (a) Bias; mean/nominal. (b) Coefficient of
variation.

504

4. CFRP DEGRADATION AND STRUCUTRAL MODELING


The long term performance of CFRP laminates is affected by field conditions (humidity,
temperature, curing ... etc). Karbhari and Abanilla [13] have studied experimentally
different sizes of FRP specimens under varous exposure conditions of accelerated tests
range between two and three years. The authors formulate durability model based on
Arrhenius acceleration law. For a specific grade of CFRP, strength and modulus can be
expressed as
(4a)
f fu (t ) f fuo ( 3.366 ln(t ) 106.07)
(4b)

E f (t ) E fo (0.418 ln(t ) 106.07)

where, ffuo is the initial FRP ultimate strength, Efo is the initial FRP modulus, t is the time in
days. The model was derived for wet layup carbon/epoxy for external strengthening.
Assessment of predictive accuracy of the model shows reliable predictions of the model
especially for long periods of exposure. Flexural behavior was evaluated through sectional
analysis. Concrete stress-strain relation was taken according to [14]. Steel bars present
elastic perfect plastic behavior, while CFRP assumed to be linear elastic. Ultimate limit
state is served by three failure modes; concrete crushing, CFRP rupture, and CFRP mid
span debonding. Many analytical models [15,16,17,18] were derived for the latter failure
mode. 41 beams were tested and fully detailed in [19-23], all beams were failed by mid
span debonding. Experimental and theoretical - using the models in the references
previously mentioned - capacities were compared. Chen and Teng model [18] gives best
correlation (=0.958) between experimental and theoretical capacities. So, it was
recommended to be used here as mid span CFRP debonding.
5. BASIC RELIBILITY ASPECTS AND APPLICATION
Reinforced concrete section performance can be expressed in the term of ultimate limit
state G as:
(5)
G m R (t ) S (t ) m R ( X 1 , X 2 ,..... X n ) ( DL DL LL LL )
where, m random variable reflects the uncertainty in theoretical resistance, R is the
resistance as function of random variables Xi (geometric and material properties). S is
expresses random loads. DL and LL are random variables express uncertainty in dead DL
and live LL loads respectively. FORM was used to estimate reliability index which can be
determined from solution of constrained minimization; =(u*Tu*)1/2 under G=0. where, u* is
the vector of the most probable design point in the standard normal space. Table 1, shows
the statistical variables considered in the analysis. In order to take into account dependency
between concrete compressive strength fc and concrete tensile strength fct. It is assumed that
fct can be expressed via fc as fct=ct fc2/3, thus fc and the coefficient ct could be assumed
independent [9]. Concrete properties gain with time were considered [14]. Atadero et al
2005 [24] study variability of CFRP properties (thickness, modulus, and strength). The
study concludes that, CFRP modulus and strength are statistically correlated to CFRP
thickness. CFRP strength and modulus statistical distributions were fitted by the authors.
Plevris et al [3] proposed an algorithm to evaluate section resistance model error
uncertainty m based on Monte-Carlo simulation, the algorithm was limited to CFRP
rupture failure mode only. In the present study, the algorithm was generalized for concrete
crushing, debonding, and CFRP rupture failure modes using sectional analysis instead of
505

concrete block principle. Different laminates layers were considered. Fig. 3 concludes
simulation results. Lognormal distribution was found best fits the generated data for both
concrete crushing and CFRP debonding, while normal distribution best fit CFRP rupture
data.
cofficient of variation

1.1
(a)

bias

1.05
1
c onc rete c rus hing
CF RP debonding
CF RP rupture

0.95
0.9

CFRP laye r num be r

(b)

0.1
0.09
0.08
0.07
0.06
0.05

CFRP layer num be r

Fig. 3. Section resistance model uncertainty (m) statistics, (a) Bias. (b) Coefficient of
variation.
The model was applied to check the reliability of a simple span bridge interior girder.
Bridge deck: width=7.2m, span=10m, slab thickness=250mm, asphalt thickness=90mm,
girder spacing=2.2m. the Bridge was designed according to AASHTO [10]. Girder design
dimensions required: width=400mm, steel depth ds=700mm with tension and compression
steel 725 and 320mm respectively. The altered LRFD design equation is
(6)

D DL wWL L LL Rn (.......,x FRP )

where D, W, and L are load factors for dead and wearing surface, and Live load
respectively [10]. WL is the wearing surface load, is the general resistance factor, Rn is the
nominal resistance, is CFRP resistance factor, and xCFRP is the CFRP contribution to
resistance. RC section should produce an initial reliability index: =4.2. It is assumed that,
strengthening time corresponds to the time required for to reach a minimum value:
min=3. Strengthening has to increase to a value greater than a target value T. Choice of
min and T values depend on cost, consequences of failure probability allowed, and required
service life. According to JCSS Model [25], moderate consequences of failure and cost
requires: T=4.2. After strengthening, two options were considered. First, option I:
corrosion will continue regardless strengthening effect on corrosion activity [26]. Second,
option II: corroded steel rebars were fully cleaned and concrete cover assumed to be
repaired. Corrosion mechanism is prevented using a system of corrosion inhibitors, thus As
statistical parameters assumed to be invariant after strengthening. CFRP resistance safety
factor CFRP was calibrated. Two target reliability indices were considered: T=3.5 and 4.2.
Strength reduction safety factor algorithm was used to calculate CFRP; for an assumed
target reliability index and CFRP, CFRP thickness is calculated incrementally according to
Eq. (6) even: =T.
6. RESULTS AND DESCUSSION
Fig. 4a presents reliability index for non-strengthened section, neglecting corrosion effect.
It is noted that, initial reliability index decreases with increasing ADTT. Moving from
ADTT=1000 to 50, drops 13%. at time; t=100years drops 50 and 60% for ADTT=50 and
1000 respectively of initial indices, however girder still reliable for about 65% - as an
506

average - of the total life. Considering corrosion effect in addition to live load growth,
reliability profiles of the most extreme cases (c=45mm, ADTT=50 truck/day & c=25,
ADTT=1000) were plotted in Fig. 4b. It can be inferred that, corrosion is the most important
factor affects reliability especially pitting type, as it turns the section in non reliable
condition: < min at t=23 years for worst case. While in case of uniform corrosion, the
section still reliable for more than 34% of total bridge life (100 years).
(a) no corrosion

4
3

ADTT=50
ADTT=250
ADTT=500
ADTT=750
ADTT=1000

2
1
0

20

40

60

80

min

3
2

strengthening
time

1
0

100

u,c=45,ADTT=50
p,c=45,ADTT=50
u,c=25,ADTT=1e3
p,c=25,ADTT=1e3

(b) with corrosion

reliability index

Reliability index

20

40

time (years)

60

80

100

time (years)

Fig. 4. Time dependent reliability index. u; uniform corrosion. p; pitting corrosion.


Appling strengthened considering anchorage and non-anchorage end laminates. Two layers
of CFRP laminates were used, each layer has thickness: tp=1.27mm and width bp=300mm.
It is found that, debonding is the control failure mode for non-anchorage laminates, rupture
yet occurs. This is probably due to the choice of CFRP thickness. Fig. 5 shows the
reliability profile of strengthened girder under the worst deterioration condition: c=25 &
ADTT=1000. It is observed that, anchorage are more and reliable than non-anchorage
laminates. Option II could maintain longer the concrete element in reliable condition, this
reflects resistance losses role due corrosion in option I on section reliability.
(a - uniform corrosion)

6
additional
service life

min

3
2

original service
life

remaining life

1
0

(b - pitting corrosion)

5
reliability index

reliability index

min

anchorage option I
anchorage option II
non-anchorage option I
non-anchorage option II

2
1

10

20

30

40

50

60

time (years)

70

80

90

100

10

20

30

40

50

60

time (years)

70

80

90

100

Fig. 5. Reliability index of strengthened girder. c=25mm; ADTT=1000 truck/day


Optimization of concrete cover versus strengthening time was plotted in Fig. 5 for uniform
and pitting corrosion. Based on the results, relation between concrete cover and
strengthening time may be considered linear till c=45. Concrete cover effectiveness
decreases after this value especially for high values of ADTT. ADTT and live load growth
with time has a high effect on accelerating strengthening time especially in uniform
corrosion.

507

60

(a) Pitting corrosion


ADTT=1000
ADTT=750
ADTT=500
ADTT=250
ADTT=50

50
40

live load
growth
no live load
growth

30
20

30

40

Cover (mm)

Cover (mm)

60

50
60
70
Strengthening time (yrs)

80

(b) Uniform corrosion

live load
growth

50
40
30

90

100

20

30

40

50
60
70
Strengthening time (yrs)

80

no live load
growth
90
100

Fig. 6. Concrete cover versus strengthening time; min=3.


Optimization of CFRP laminates thickness versus a given total service life was performed.
Anchorage and non-anchorage laminates end were considered. Laminates thickness were
optimized for three level of deterioration after strengthening; option II, option I, and option
I in addition to CFRP durability. Optimization results were plotted in Fig. 7 and 8. It can be
noted that, results of anchorage and non-anchorage laminates are identical for small values
CFRP thickness. As CFRP rupture is the control failure mode for small thickness.
Increasing thickness considering debonding CFRP is inefficient as in CFRP rupture
especially when considering corrosion after strengthening and CFRP durability. CFRP
thickness is linearly proportional with total service life. It is also to be noted that, CFRP
durability has negligible effect on debonding failure mode while significant effect is
observed when using anchorage laminates. This returns to the high contribution of CFRP in
rupture failure mode and dependency of rupture failure mode on both strength and
modulus. While modulus ECFRP only affects the debonding failure mode[18]. Although,
CFRP rupture failure mode is significantly affected by CFRP durability, but it gives longer
service life than debonding failure mode for the same CFRP dimensions. It could be
concluded that, for a certain desired service life rupture failure mode using anchorage
ends - is safer, more economical in CFRP quantities, and more durable than debonding
failure mode when use the same CFRP laminates dimensions (restricted design).
8

non-anc horaged

FRP thickness (mm)

ADT T =1000
ADT T =750
ADT T =500
ADT T =250
ADT T =50

6
5
4
3

anc horaged

2
1
20

40

60

80

Service life (yrs)

100

(b) option I

7
6

FRP thickness (mm)

(a) option II

FRP thickness (mm)

non-anc horaged

5
4
3
2
1
20

anc horaged
40

60

80

(c ) option I & CFRP durability

7
6

non-anc horaged

5
4
3
2
1
20

100

Service life (yrs)

anc horaged
40

60

80

100

Service life (yrs)

Fig. 7. Laminate thickness versus total service life, uniform corrosion, c=25mm.
8

non-anc horaged

7
6
5
4

AD T T =1000
AD T T =750
AD T T =500
AD T T =250
AD T T =50

3
2
1
20

anc horaged
40

60

80

Service life (yr)

100

(b) option I

non-anc horaged

6
5
4

anc horaged

3
2
1
20

40

60

80

Service life (yr)

FRP thickness (mm)

(a) option II

FRP thickness (mm)

FRP thickness (mm)

100

(c ) option I & CF RP durability

non-anc horaged

6
5
4
3

anc horaged

2
1
20

40

60

80

Service life (yr)

Fig. 8. Laminate thickness versus total service life, pitting corrosion, c=25mm.

508

100

Table 2 presents the results of CFRP resistance safety factor CFRP calibration. Two values
of target indices were considered T=3.5, and 4.2. The former corresponds the AASHTO
[10]. The second corresponds moderate consequences of failure and cost [24]. It could be
observed that, CFRP depends on target value, ADTT, steel losses, and anchorage system .
Corrosion type has significant effect on CFRP. Small values of CFRP is recorded for high
value c: c=45mm, as strengthening time is delayed, and the need of strengthening returns to
live load growth as a principle deterioration factor. Therefore, it could be concluded that
strengthening effectiveness in increasing girder reliability index is more significant for steel
losses than live load growth. Also it could be observed that, for free design (no restriction
in calculating CFRP amounts) non-anchorage CFRP laminates is more durable (1.5 to 3
times) than end anchorage laminates, and gives higher amounts of CFRP (1 to 1.7 times)
than anchorage. The choice of T is very important in calibrating CFRP. The recorded
(CFRP=1) correspond to case satisfy T with CFRP amounts less than the presented in
Table 2, but the need for these amounts of CFRP is to satisfy Eq. 6. This is more evident in
the case of (T =3.5) which is proposed by AASHTO provision [10]. This may reflect the
over estimation exists in AASHTO live load or its load factors. Also it is noted that, for
pitting corrosion strengthening with option II provides additional service life Lstr (2 to 3.5)
times service life provided by option I depending on concrete cover value regardless
ADTT. Lower ranges (1.1 to 1.5) are observed for general corrosion. Strengthening with
(T =4.2) provide better durability than T proposed by AASHTO provision with no
significant differences in CFRP amounts required. Thus, CFRP safety factor corresponds to
high corrosion losses may taken in CFRP=0.85 and CFRP=0.6 for CFRP non-anchorage and
anchorage laminates respectively. While for live load growth and slight corrosion losses
safety factor can be taken CFRP=0.6 and CFRP=0.4 for non-anchorage and anchorage
laminates respectively.
7. CONCLUSIONS
The study presents CFRP strengthened RC girder reliability affected by corrosion, live load
traffic growth, and CFRP durability. Failure modes considered are concrete crushing, CFRP
debonding, and CFRP rupture. Results have shown that, traffic affect initial reliability
index. Pitting corrosion is more aggressive than uniform corrosion. Deterioration factors
can be ordered according to their influence as corrosion, live load growth, and CFRP
durability. No significant effect is observed in service life after strengthening due to CFRP
durability with debonding failure mode while significant effect is observed for anchoring
laminates. CFRP strengthening effectiveness in increasing girder reliability is more
significant for steel losses than live load growth. AASHTO design equation seems to be
over estimated for strengthening purpose. Strengthening with (T =4.2) provide better
reliability than T proposed by AASHTO with no significant differences in CFRP amounts
required.

509

Table 2. Summary of safety factor CFRP.

, ADTT=50
,
ADTT=250

T = 4.2

ADTT=500

ADTT=750

ADTT=1000

tp (mm)
Lstr
(I+dur)
Lstr(II+d
ur)

, ADTT=50
,
ADTT=250

T = 3.5

ADTT=500

ADTT=750

ADTT=1000

tp (mm)
Lstr
(I+dur)
Lstr(II+d
ur)

Uniform corrosion
anchorage end
non-anchorage end
c=45
35
25
45
35
25
0.37
0.51
0.67
0.57
0.75
0.92

Pitting corrosion
anchorage end
non-anchorage end
45
35
25
45
35
25
0.4
0.64
0.76
0.59
0.9
1

0.31

0.45

0.61

0.50

0.69

0.85

0.28

0.52

0.67

0.48

0.77

0.95

0.28

0.42

0.58

0.47

0.65

0.82

0.25

0.47

0.64

0.4

0.71

0.91

0.27

0.41

0.57

0.45

0.64

0.81

0.25

0.44

0.62

0.4

0.68

0.89

0.26

0.39

0.57

0.43

0.63

0.97

0.25

0.42

0.61

0.4

0.65

0.88

1.921.89
23.923.4
33.131.2
0.81

1.891.87
22.922
3331.9
1

1.751.72
19.818.7
30.729.2
1

3.653.22
31.830.7
35.342
1

3.493.17
30.429
43.842.8
1

3.273.04
28.427.3
44.943.3
1

1.911.91
18.217.9
34.933.5
0.88

1.871.89
16.416
35.834.6
1

1.851.84
12.111.4
38.835.3
1

3.843.6
23.723.2
44.944.9
1

3.613.37
21.120.2
48.445.8
1

4.023.36
15.314.1
53.847.7
1

0.68

0.95

0.65

0.62

0.89

0.57

0.96

0.59

0.87

0.57

0.97

0.96

0.57

0.84

0.97

0.57

0.92

0.96

0.980.97

1.030.97
9.88.3
14.312.4

1.221.01
12.18.6
18.813.8

1.085
-0.6
20.34
28.85.2

1.921.15
24.419.8
36.330.1

2.761.84
26.722.6
42.436.5

0.950.96

1.240.95

1.441.17

7-7.1

9.6-6

8.9-6

13.713.1

21.313.6

28.819.4

1.230.6
15.33.7
31.86.2

2.91.34
19.514.1
45.133.4

4.012.55
15.312.7
53.843.6

9.1-9
12.612

Lstr additional service life (yrs) after strengthening (see Fig. 5). I & II are first & second
options; dur; CFRP durability. tp laminates thickness.
8. REFERENCES
1.
2.
3.
4.

Cairns, J., Plizzari, G.A., Du, Y., Law, D.W., and Franzoni, C. 2005. Mechanical
Properties of Corrosion-Damaged Reinforcement. ACI Material Journal 102(4): 256264.
Li, C., Melchers, R.E., and Zheng, J. 2006. Analytical model for corrosion-Induced
Crack Width in Reinforced Concrete Structures. ACI Structural Journal 103(4): 479487.
Plevris, N., Triantafillou, T., and Veneziano, D. 1995, Reliability of RC members
strengthened with CFRP Laminates. Journal of Structural Engineering, ASCE, 121(7):
1037-1044.
Pham, H.P., and Al-Mahidi, R. 2008. Reliability analysis of bridge beams retrofitted
with fibre reinforced polymers. Composite Structures, 82: 177-184.

510

5.
6.

7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.

Atadero, A.A., and Karbhari, V.M. 2008. Calibration of resistance factor for reliability
based design of externally-bonded FRP composites. Composites: Part B, 39(4): 665679.
Ali, O., and Bigaud, D. 2010. Time-variant flexural reliability of RC beams with
externally bonded CFRP under coupled fatigue-corrosion action. Proceedings of the 3rd
international conference on accelerated life testing, ALT2010 Clermont-Ferrand,
France, pp. 205-212.
Val, D.V., Stewart, M.G., and Melchers, R.E. 2000. Life-Cycle Performance of RC
Bridges Probabilistic Approach. Computer-Aided Civil Infrastructure Engineering, 15:
14-25.
Stewart, M.G., and Al-Harthy, A. 2008. Pitting corrosion and structural reliability of
corrosion RC structures: Experimental data and Probabilistic analysis. Reliability
Engineering and System Safety, 93: 373-382.
Vu, K.A.T., and Stewart, M.G. 2000. Structural reliability of concrete bridges
including improved chloride-induced corrosion models. Structural safety 23: 313-33.
American Association of state of Highway Transportations Officials (AASHTO).
2005. LRFD Bridge Design Specifications. Washington DC, USA.
Nowak, A.S. 1993. Live load model for high way bridges. Structural Safety, 13: 5366.
Nowak, A.S. 2004. System reliability models for bridge structures. Bulletin of the
Polish Academy of Science, Technical Science, 52(4): 321-328.
Karbhari, V.M., and Abanilla, M.A. 2007. Design factors, reliability, and durability
predictions of wet layup carbon/epoxy used in external strengthening. Composites: Part
B, 38: 10-23.
CEB-FIP CEB-FIP, Model Code 1990. London: Thomas Telford Services Ltd. 460 p.
ACI Committee 440. 2002. Guide for the design and construction of externally bonded
FRP system for strengthening concrete structures. ACI 440.2R-02, American Concrete
Institute, Farmington Hills, MI, USA.
Fib bulletin 14, FIB TG 9.3 FRPEBR. 2001. Externally bonded FRP reinforcement for
RC structures. Fderation international du bton (Fib), Task Group 9.3 FRP, 130 p.
Concrete Society Technical Report 55. 2000. Design guidance for strengthening
concrete structures using fiber composite material, The Concrete Society, Crowthorne,
UK.
Chen, J.F., and Teng, J.G. 2001. Anchorage strength model for FRP and steel plates
bonded to concrete. Journal of Structural Engineering, ASCE, 127 (7): 784-791.
Ashour, A.F., El-Refaie, S.A., and Garrity, S.W., 2004. Flexural strengthening of RC
continuous beams using CFRP laminates. Cement & Concrete Composites, 26: 765775.
Bogas, J.A., and Gomes, A. 2008. Analysis of the CFRP flexural strengthening
reinforcement approaches proposed in fib bulletin 14. Construction and Building
Materials, 22(10): 2130-2140.
Aram, M.R., and Czaderski, C. 2008. Debonding failure modes of flexural FRPstrengthened RC beams. Composites: Part B, 39: 826-841.
Pham, H., and Al-Mahaidi, R. 2004. Experimental investigation into flexural
retrofitting of reinforced concrete beams using FRP composites. Composites
Structures, 66: 617-625.

511

23. Teng, J.G., Smith, S.T., Yao, J., and Chen, J.F. 2003. Intermediate crack-induced
debonding in RC beams and slabs. Construction and Building, 17: 447-462.
24. Atadero, A.A., Lee, L., and Karbhari, V.M. 2005. Consideration of material variability
in reliability analysis of FRP strengthened bridge decks. Composites Structures, 70:
430-43.
25. JCSS, Joint Committee of Structural Safety. 2001. Probabilistic Model Code. (Internet
Publication: www.jcss.ethz.ch).
26. Gadve, S., Mukherjee, A., and Malhotra, S.N. 2009. Corrosion of steel reinforcement
embedded in FRP wrapped concrete. Construction and Building Material, 23: 153-161.

512

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

AXIAL BEHAVIOUR OF CONCRETE-FILLED


COLUMNS AFTER 300-FREEZE-THAW CYCLES

FRP

TUBE

H. El-Zefzafy1, H.M. Mohamed2 and R. Masmoudi3


1

Ph.D. Candidate, Department of Civil Engineering, University of Sherbrooke, Sherbrooke, QC,


Canada
2
NSERC-Industrial Postdoctoral Fellow, Department of Civil Engineering, University of
Sherbrooke, Sherbrooke, QC, Canada
3
Professor , Department of Civil Engineering, University of Sherbrooke, Sherbrooke, QC, Canada

ABSTRACT
Deterioration of concrete structure caused by aging and weathering is a major problem in
harsh environments cold region. Concrete-filled fiber-reinforced polymer (FRP) tubes
(CFFTs) system is one of the most promising techniques to protect the reinforced concrete
(RC) structures from aggressive environmental conditions. However, there are several
concerns related to the behavior of the CFFT as a protective jacket against harsh
environmental effects which have limited its full implement. In this study an experimental
work was performed to investigate the long term effect of freeze-thaw cycles on the
behavior of reinforced CFFT columns. Test parameters in this investigation include the
effect of type of longitudinal reinforcement rebars (steel, GFRP, CFRP) and type long-term
freeze-thaw cycles exposure (dry air, fresh and salt water). Thirty two CFFT and RC
columns were manufactured in the laboratory and conditioned to 300 freeze-thaw cycles.
The identical specimens were kept in the room temperature. Then, pure axial tests were
conducted in order to evaluate the change of mechanical properties of the test columns due
to the freeze-thaw exposure. The test results indicated that the freeze-thaw cycles
considered in this study did not adversely affect the compressive behavior of CFFT
columns. Some of the conditioned specimens showed an increase in the compressive
strength as compared to the room CFFT specimens.
1. INTRODUCTION
Corrosion of steel reinforcement causes continual degradation to the worldwide
infrastructures and it has prompted the need for challenges to those involved with
reinforced concrete (RC) structures. Recently, the use of fiber-reinforced polymers (FRP)
tubes as structurally integrated stay-in-place forms for concrete members, such as beams,
columns, bridge piers, piles and fender piles has emerged as an innovative solution to the
corrosion problem. In such integrated systems, the FRP tubes may act as a permanent form,
often as a protective jacket for concrete, and especially as external reinforcement in the
513

primary and secondary directions such as for confinement. Furthermore, the use of
concrete-filled FRP tubes (CFFT) technique is predicated on performance attributes linked
to their high strength-to-weight ratios, expand the service life of structures, enhance
corrosion resistance, and potentially high durability.
In fact, extensive research programs have been conducted to investigate the behavior of
concrete columns confined with FRP sheets and FRP tubes under pure compression load
(Mohamed and Masmoudi 2010; Mohamed 2010; Fam and Rizkalla 2002). However,
relatively few studies have focused on the durability of CFFT columns. Short and long term
durability of the CFFT is the most important factor needed for CFFT to become widely
accepted for application in new constructions. The majority of research studies regarding
the durability of RC structures strengthened or reinforced with FRP composite materials
have been conducted with focus on the effect of freeze-thaw cycles(Callery, Green, and
Archibald 2000; EL-Hacha and Green and Wight 2010). This is because the freeze-thaw
action has been thought to be the most harmful to concrete structures and FRP composite
materials as well. Problems with deteriorated infrastructures in marine settings and cold
regions are mainly due to exposure to salt water and subject to the force of water freezing
and expanding into ice (Saenz and Pantelides 2006; Toutanji, Zhao, and Isaacs 2007;
Micelli and Myers 2008).
A significant few amount of research has been investigated the effect of freeze-thaw on the
behavior of CFFT columns. It is obvious that further experimental research regarding to the
durability effect is required on CFFT specimens. The present study attempts to enrich the
database of axial behavior of conditioned CFFT columns to freeze-thaw cycles. Thirty two
medium scale RC and CFFT specimens were conditioned to 300 freeze-thaw cycles and
then were tested under pure axial load. Filament-wound glass FRP tubes are used as an
FRP-stay in place structural formwork for concrete columns. The following sections
describe briefly the experimental program and test results of this study.
2. EXPERIMENTAL WORK
2.1

Materials

2.1.1

Filament-wound glass FRP tubes

All specimens constructed from the same type of the glass fibre-reinforced polymers
(GFRP) tubes. The internal diameter of the tubes is equal to 152 mm. The GFRP tubes were
fabricated using filament winding technique; E-glass fibre and Epoxy resin were used for
manufacturing these tubes. The GFRP tubes consist of four (60) layers oriented in the
hoop direction with respect to the longitudinal axis of the tubes, the total thickness is 2.65
mm. The split-disk test and coupon tensile test were performed according to ASTM D2290-08 and ASTM D 638-08 standard, respectively, on five specimens of the tube. Figure
1.a shows the average stress-strain relationship for the split-disk test, the load-strain curve
for the split-disk test was linear up to failure for all specimens. Figure 1.b presents the axial
tensile stress-strain responses for the average results obtained from the five specimens for
the coupon tests.

514

2.1.2

Concrete

All specimens were constructed from the same batch of normal strength ready mix
concrete. Water reducing admixture with super plasticizer was used to increase the
workability of the concrete mix. Ten plain concrete cylinders (152305 mm) were tested at
28-days under axial load. The average concrete compressive strength for ten cylinders was
found 330.4 MPa.

(a) Split-disk test


2.1.3

(b) Coupon tensile test


Fig. 1. stress-strain curve for coupon and ring test

Reinforcing steel

Two different steel bar diameters were used to reinforce the CFFT and control specimens.
Mild steel bars of 3.4 mm diameter were used as spiral reinforcement for the control
specimens. Deformed steel bars No. 13M were used as a longitudinal reinforcement for
CFFT and control columns. The mechanical properties of the steel bars were obtained from
standard tests that were carried out according to ASTM A615/A615M-09, on five
specimens for each type of the steel bars. The actual properties are given in Table 1 in
terms of diameter, nominal area, yield and ultimate strength and youngs modulus.
2.1.4

FRP Bars

Sand-coated glass and carbon FRP bars manufactured by a Canadian company [Pultrall
Inc., Thetford Mines, Quebec], with a fibre content of 73% in vinyl ester resin, were used.
The bars were made of continuous fibres (glass or carbon) impregnated in a vinyl ester
resin using the pultrusion process. GFRP and CFRP bars #3 were used as longitudinal
reinforcement for the CFFT columns. Figure 2 shows a photo of the FRP bars used in this
research. Table 2 presents the mechanical properties of the FRP bars.

515

Table 1. Properties of reinforcing steel bars.


Modulus
Bar type
Nominal Yield Ultimate
Diameter
of
area
strength strength
(mm)
elasticity
Size
Type
(mm)
(MPa)
(MPa)
(GPa)
Mild
Wire
3.2
675
850
221
Steel
13
13M
Deformed
12.7
129
356
536
200
Table 2 Properties of reinforcing FRP bars.
Nominal Nominal Tensile modulus
Size
Type
diameter
area
of elasticity
(mm)
(mm)
(GPa)
#3
GFRP
9.525
71
45.4
#3
CFRP
9.525
71
207,4

Ultimate tensile
Ultimate
strength
strain (%)
(MPa)
765
1.60.03
1431
1.200.09

(a) Glass FRP Bars


(b) Carbon FRP Bars
Fig. 2. Sand coated FRP reinforcing bars.
2.2

Specimens Details and Preparation

In this study, a total of thirty two CFFT and RC columns (152 x 912 mm) were prepared
and tested under pure axial load. The specimens were divided into four groups; each group
consists of two steel-spiral control RC and six CFFT columns, each two CFFT columns
were reinforced either with steel (A90S), carbon FRP (A90C) or glass FRP (A90G) rebars.
The first group was used as reference and kept at room temperature for a period equivalent
to the 300 freeze-thaw (F/T) cycles. The other three groups were exposed to 300
freeze/thaw cycles inside an environmental chamber. The second group was kept in air
inside the environmental room, while the third and forth groups were submerged in fresh
and slat water, respectively.
The RC and CFFT columns were reinforced with six deformed bars, see Figure 3.a. The
bars were distributed uniformly inside the cross section of the tube, see Figure 3.b. The
distance between the bars and the tubes was 8 mm. In addition, a concrete cover of 10 mm
was provided between the ends of the longitudinal steel bars and the top and bottom
surfaces of the specimens to avoid the stress concentration at the steel bars area. In
addition, mild steel bars of diameter 3.2 mm are used for the spiral reinforcement. The
pitches of the spiral are 50.6 mm through all the length of the specimens, while it was
reduced to 25 mm a distance of 125 mm at both ends of the column. This is to avoid the
local failure at the end regions of loading. All the specimens were tested in uni-axial
compression after the environmental conditioning. The specimens were tested using a 6,000

516

kN capacity FORNEY machine, where the CFFT and RC columns were setup vertically at
the center of loading plates of the machine. The FORNEY machine, strain gauges and
LVDTs were connected by a 20 channels Data Acquisition System and the data were
recorded every second during the test.

(a) Reinforcement cages


(b) Overview for cages inside tube
Fig. 3. Reinforcement cages and Overview for cages inside tube
2.3
Freeze-Thaw Exposure
The dry freeze-thaw specimens were placed in the freeze-thaw chambers. Whereas, those
for saturated freeze-thaw were left in fresh and salt water bath and placed in the same
environmental chamber using the isolated wooden tanks. According to the different types
of environmental conditions (freezethaw cycles in dry, fresh and salt water) existed inside
the same environmental chamber, it was hard to reach the same temperature inside the
saturated and non-saturated specimens. Therefore, a temperature acquisition system was
used to monitor the temperature during cycling. The freezing cycles consisted of lowering
the temperature in the middle of the saturated specimens from 4.4 to -17.8C in a period of
16.5 h. The thawing cycles consisted of raising the temperature in the middle of the
saturated specimens from -17.8 to 4.4C. Those freeze-thaw hours were sufficient to vary
the temperature of non-saturated specimens between (+28C to -28C) as shown in Figure
4. Thus, the specimens underwent one freeze-thaw cycles per 27 hours, rather than in
accelerated shorter cycles.

Fig. 4. Temperature variation inside the specimens during two complete F/T cycles

517

3. RESULT AND DISSECTIONS


3.1

Effect of Confinement

Table 4 presents the failure loads of control and CFFT columns subjected to different
environmental condition. Confining using the GFRP tubes showed a large increase in the
axial compressive capacity of CFFT columns in comparison with RC specimens, as a result
of the confinement in the radial direction. The test results showed that the axial
compressive capacities of steel, CFRP and GFRP reinforced CFFT columns of Group No. 1
were increased by 57, 36 and 65% respectively, than that of the control RC column
specimens.
3.2

Effects of Freeze- Thaw Cycles

It is important that we review results here to explain the effects of freezing and thawing on
the confined concrete. During the freeze cycle water in the pores turns into ice and increases
in volume by approximately 9%. Thus, stress is induced on the surrounding concrete. Similar
pressure will be generated if water is present in spaces between the FRP tubes and the
confined concrete core. During the thaw cycles this induced stress is released. If the stress
induced during the freeze cycles is higher than the tensile strength of the concrete, with
increasing the freeze thaw cycles, damage such as cracks in the concrete occurs. Thus, as
long as the F/T cycles are repeated more damage will happen due to water penetration into
the cracks.
In this section the test results of all specimens will be compared to the reference CFFT
specimens kept at room temperature. An insignificant decrease 2.1 and 3.9% in the
compressive capacitates of steel-reinforced CFFT columns that conditioned to 300 F/T
cycles at air dray and submerged in salt water, respectively, was observed. It should be
noted at this point that the reduction occurred although air entrainment concrete was used in
this study to reduce the freeze-thaw effect as much as possible. However, the steelreinforced CFFT columns tested after exposure to the 300 dry F/T cycles in fresh water
showed insignificant increase in the average compressive strength (2.5%). On the other
hand, all the CFRP-CFFT columns conditioned either in dray air, fresh and salt water
presented on average a reduction in axial capacity 4.7, 6.9 and 7.2% respectively. The
GFRP-CFFT columns showed a similar effect and the corresponding values were 4.5, 5.8
and 6.7%, respectively. The control RC-columns exhibited a significant reduction in the
axial capacity as the concrete exposed directly to the severe environmental condition. On
average a reductions of 19.93, 13 and 20.25% have been observed for specimens,
respectively, conditioned at dry air, fresh and salt water.
The test results indicated that the combined environmental cycles such as freeze-thaw
cycles in fresh water or dray air did not show a significant effect on the CFFT specimens.
This was probably because the adverse effect, such as plasticization of matrix and microcracking at matrix-fibre interface, induced by the low temperature during the freeze-thaw
cycles, compromised the positive effect, such as matrix hardening effect. On the other
hand, the freeze-thaw cycles affect the compressive behavior of RC columns. The possible

518

reason for the decrease could be attributed in large part to the cracks, spalling, damage,
steel corrosion and deterioration of the concrete cover.
Table 4. Failure loads (kN) of control and CFFT columns subjected to different conditions
Specimen
Group No.
A90S
A90C
A90G
Cont-90
No.
1
1454
1335
1556
935
Group (1)
2
1498
1218
1558
942
22.5C
Average
1476
1276
1557
938
1
1327
1231
1553
741
2
1564
1200
1421
762
Group (2)
Thaw in
Average
1445
1215
1487
751
dry air
Difference
-2.1
-4.7
-4.5
-19.93
%*
1
1464
1178
1434
836
2
1563
1197
1498
799
Group (3)
Thaw in
Average
1513
1187
1466
817
fresh water Difference
+2.5
-6.9
-5.8
-13
%*
1
1450
1189
1463
755
2
1384
1179
1441
741
Group (4)
Thaw in
Average
1417
1184
1452
748
salt water
Difference
-3.9
-7.2
-6.7
-20.25
%*
*as compared to the average results of Group No. 1
3.3

Failure Modes

From carful observation during the freeze-thaw cycles, it should be noted that the
unconfined cylinders exposed to freeze/thaw conditions started to showed visual signs of
deterioration in the concrete. Several cracks were observed after 40 F/T cycles. By the end
of the 300 cycles prior to testing, all the specimens were visually inspected for evidence of
degradation in of micro-cracks. More signs of deterioration in the saturated plain concrete
cylinders after F/T exposure were reported, rather than specimens exposed to F/T cycles in
air. Figure 5.a shows the damage and deterioration of the concrete cover of RC. On the
other hand, corrosion of steel reinforcements of control RC columns generated the internal
pressure around the steel and concrete interface, eventually causing the cracks and spalling
of the concrete cover.
In general, the F/T cycles considered in this study did not adversely affect the modes of
failure of specimens that exposed to different condition. The failure modes of CFFT
columns were a combination of rupture of the confining FRP tubes and local buckling of
internal steel bars at the column mid-height. Typical failure was generally observed by
rupture of the tubes between one end and the mid-height of the specimen, see Figure 5.b
and 5.c. The rupture of the tube extended to the third height of the specimen. The failure
occurred due to local buckling of the steel bars immediately followed by the rupture of the
519

tube. The failure of the control specimen started by vertical cracks at the bottom region of
the column distributed uniformly around the hoop direction, and no buckling was observed.

a. Cont-90 (fresh water)

b. A90S (dry air)

c. A90S (fresh water)

Fig. 5. Different failure modes of conditioned RC and CFFT columns


4. SUMMARY AND CONCLUSIONS
The study presented herein intended to examine experimentally the long term effects of
freeze-thaw cycling on the compressive behaviour of reinforced concrete-filled FRP tubes
columns. In summary, it was found that the confinement using a GFRP tube appear to
provide excellent protection against freeze thaw cycles. Based on the test results so far, the
following conclusions are drawn:
Confinement using GFRP tubes was found to enhance the compressive capacities of
CFFT columns from 36 to 65% that of RC columns.
Regardless of the type of the freeze-thaw cycles on the long term (300 freeze-thaw
cycles), the freeze-thaw cycles have almost no affect on the average axial compressive
strength of the CFFT columns (less than 5%).
The axial capacity of the CFFT columns conditioned under the freeze-thaw cycles in
fresh water slightly increased (2.5%). The possible reason for the increase could be
attributed to the matrix-hardening effect resulting from the extremely low temperature
during the freeze-thaw cycles.
The RC columns exposed to 300 freeze-thaw cycles showed a significant reduction (on
average 17.75 %) in the axial compressive capacity due to the cracks and spalling of the
concrete cover.
5. ACKNOWLEDGEMENTS
The research reported in this paper was partially sponsored by the Natural Sciences and
Engineering Research Council of Canada (NSERC). The authors also acknowledge the
contribution of the Canadian Foundation for Innovation (CFI) for the infrastructure used to
conduct testing. Special thanks to the FRE Composites Inc, Qc, Canada, and Pultrall Inc.
Qc, for providing the FRP materials.

520

6. REFERENCES
1.

ASTM. 2008a. Standard test method for apparent hoop tensile strength of plastic or
reinforced plastic pipe by split disk method. D 2290-08, West Conshohocken, Pa.
2. ASTM. 2008b. Standard test method for tensile properties of plastics. D638-08, West
Conshohocken, Pa.
3. Callery, K., Green, M.F, and Archibald, J.F. 2000. Environmental effects on the behavior
of wrapped concrete cylinders. Proceedings of the 3rd international conference on
Advanced Composite Materials in Bridges and Structure, Canada, pp. 759-766.
4. EL-Hacha, R., Green, M.F., and Wight, G.R. 2010. Effect of severe environmental
exposures on CFRP wrapped concrete columns. Journal of Composites for Construction,
ASCE, 14(1): 83-93.
5. Fam, ZA, and Rizkalla, S. 2002. Flexural behavior of concrete-filled fiber reinforced
polymer circular tubes. Journal of Composites for Construction, ASCE, 6(2): 23-32.
6. Mohamed, H.M. 2010. Axial and flexural behavior of reinforced concrete-filled fiber
reinforced polymer tubes: experimental and theoretical studies. Ph.D. thesis. University
of Sherbrooke, Sherbrooke, QC, Canada.
7. Karbhari, V.M, and Eckel, D.A. 1994. Effect of cold region climate composite jacketed
concrete columns. Journal of Cold Region Engineering, 8(3): 73-86.
8. Karbhari, V.M, Rivera, J., and Dutta, P.K. 2000. Effect of short-term freeze-thaw cycling
on composite confined concrete. Journal of Composites for Construction, ASCE, 4(4):
191-197.
9. Micelli, F. and Myers, J.J. 2008. Durability of FRP confined concrete. Construction
Materials 6(4).
10. Mohamed, H., and Masmoudi, R. 2010. Axial load capacity of reinforced concrete-filled
FRP tubes columns: experimental versus theoretical predictions. Journal of Composites
for Construction, ASCE, 14(2): 231-243.
11. Saenz, N., and Pantelides, C.P. 2006. Short and medium term durability evaluation of
FRP-confined circular concrete. Journal of Composites for Construction, ASCE, 10(3):
242-253.
12. Toutanji, H., Zhao, L., and Isaacs, G. 2007. Durability studies on concrete columns
confined with advanced fiber composites. International Journal of materials and product
technology, 28: 8-28.

521

522

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

CFRP-STRENGTHENED BEAMS UNDER REPEATED IMPACT


LOADING
A. Parvin1 and T.A. Mohammed 2
1
2

Associate professor, Dept. of Civil Engg., The Univ. of Toledo, USA


PhD candidate, Dept. of Civil Engg., The Univ. of Toledo, USA

ABSTRACT
The behavior of fiber reinforced polymer (FRP)-strengthened reinforced concrete (RC)
structures for quasi-static monotonic loadings has been studied thoroughly. However,
limited investigations have been performed on the progressive collapse analysis of FRPstrengthened RC beams under repeated impact loading. In this paper, the response of asbuilt and carbon fiber reinforced polymers (CFRP)-strengthened RC beams subjected to
repeated drop hammer impact loading is investigated. Three-dimensional complex
nonlinear finite element analysis (FEA) models of as-built and CFRP-strengthened beams
were developed using LS-DYNA explicit nonlinear finite element software program at the
Ohio Supercomputer Center. An experimental study on six as-built reinforced concrete
beams reported in the literature was used to validate the accuracy of the proposed FEA
models. CFRP composite strengthening increased beams durability to sustain repeated
drop hammer impact loadings and improved mid-span deflection in the range of 72.47 to
87.57 %. The strengthening also contained progressive crack propagation and brittle
flexural failure mode.
1. INTRODUCTION
Concrete structures such as bridge piers, sediment control dams, harbors for docking ships,
helipad landing and takeoff slabs, decks of aircraft carrier ships must be designed for
repeated impact loadings. Many researchers have studied the behaviors of fiber reinforced
polymer (FRP)-strengthened reinforced concrete (RC) structures for quasi-static monotonic
loadings. However, little is known about as-built and FRP-strengthened RC structures
subjected to repeated impact loadings. Some of the scattered studies on concrete elements
subjected to impact loadings entail the followings.
Impact is characterized by high intensity loading rate applied in a short period of time.
Concrete material behaves differently under such load as compared to static loading (Tang
and Saadatmanesh 2005). As compared to static, in dynamic loading when the rate of
loading increased, the concrete material exhibited an increase in strength, toughness, and
modulus of elasticity and behaved more brittle (Sukontasukkul et al. 2004). Banthia et al.
523

(1987) studied impact load response of normal strength (NS), high strength (HS), and steel
fiber reinforced concrete beams. It was observed that HS concrete had higher impact
strength and brittle failure as compared to NS concrete. Cracks in the HS concrete beams
propagated in moderately straight path through both the paste and the aggregate. Cracks in
the NS concrete beams propagated in winding path through the paste and aggregate failures
were not common. Furthermore, the steel fibers added ductility to the steel fiber reinforced
concrete beams and the beams showed higher impact resistance as compared to NS
concrete. Arslan (1995) investigated the performance of plain and fiber reinforced concrete
cubes under static and repeated impact loadings. It was reported that the fracture energy of
plain concrete under impact loading was ten times higher than in static tests. The author
also concluded that fiber reinforced concrete had higher fracture energy as compared to
plain concrete. Chakradhara et al (2011) studied the behavior of normal and recycled
aggregate concrete beams under repeated drop weight impact loading. They reported that
cracks were initiated at or near the beam bottom face below the impact location. After
cracks initiation, all their beams were failed in subsequent one or two drop weight impact
loadings. The review of the literature indicates lack of studies on strengthening of
reinforced concrete members subjected to repeated impact loadings. The present paper
involves investigation on progressive collapse of as-built RC beams and use of CFRP
sheets to strengthen and to modify the response of as-built RC beams subjected to repeated
drop hammer impact loadings. The proposed CFRP-strengthening technique, which
consists of complete wrapping of the beam, targets to contain both the developed shear and
flexural cracks and to mitigate concrete crushing failure at the impact location on the top
face of the beam. Repeated drop hammer impact loadings were applied until total collapse
of the beams. Key response variables such as structural force demand, mid-span defection,
beams damage pattern, crack propagation and failure modes were extracted and presented
as a function of time and number of repeated impact loading. Detailed finite element
analysis procedures and results are presented in the following sections.
2. FINITE ELEMENT ANALYSIS MODELING OF BEAMS
Nonlinear finite element software program ANSYS was employed as a preprocessor to
create the mesh and geometry of the reinforced concrete beams (ANSYS 2009). LSDYNA and LS-PrePost software programs were used to perform nonlinear explicit finite
element analysis and to define all material models, respectively (LSTC 2007a,b).
2.1 Element Types
A three-dimensional eight-node brick element, Solid164 was used in ANSYS to model the
concrete, the impacting drop hammer, and the steel plates at the support locations. The
element has three displacement, velocity, and acceleration degrees of freedom per node in
x, y and z directions. In the analysis, the element was used as reduced one point integration
with viscous hourglass control for faster element formulation. Hourglass is a zero-energy
mode with no physical meaning and generates a zigzag mesh form in the FE model.
Solid164 element supports material and geometric nonlinearities in the explicit dynamic
analysis.

524

Reinforcement bars were modeled using a three-dimensional spar element, Link160. The
geometry of the element was defined with one-node at each element end and additional
orientation node at element center. Link160 element has three degrees of freedom at each
node for displacement, velocity and acceleration in x, y and z directions. Link160 element
supports material and geometric nonlinearities and is compatible with Solid164 brick
element.
CFRP sheets were modeled using a four-node membrane element, Shell163. The element
has three degrees of freedom at each node for displacement, velocity and acceleration in x,
y and z directions, compatible with Solid164 brick element. CFRP layers were modeled
using equal spacing integration points with options to define fiber orientation for each
layer. Beytschko-Tsay element formulation was used for membrane Shell163 element.
2.2 Material Models
LS-DYNA material model 159 was employed to characterize the concrete behavior. The
model has a smooth intersection between the shear yield surface and hardening cap. The
initial damage surface coincides with the yield surface. The rate effects are modeled with
visco-plasticity. An element loses its strength and stiffness once the damage accumulation
would be equal to unity. The model is mesh insensitive and maintains constant fracture
energy regardless of the element size.
LS-DYNA material model 3 (MAT_PLASTIC_KINEMATIC) was employed to model the
steel bars. It has the options to integrate rate effects and kinematic or isotropic hardening
rules. Failure strain value of 0.16 was used to erode failed elements. Perfect bond
assumption between the steel bars and concrete was forced by coinciding Link160 and
Soid164 element nodes so that the two materials share the same nodes.
LS-DYNA material model 22 (MAT_COMPOSITE_DAMAGE) was selected to model the
CFRP sheets. The model is orthotropic material with brittle type failure option. The CFRP
sheet failure criterion was set as the maximum tensile strength in the fiber direction.
Perfect bond assumption between the CFRP sheets and the concrete was implemented by
connecting adjacent nodes of Shell163 and Solid164 element to share the same nodes.
2.3 Mesh Generation and Boundary Conditions
The concrete beams were meshed with element sizes ranging from 22.5 to 25 mm
following a preliminary analysis to optimize the mesh density for accuracy and for
flexibility to enforce the perfect bond assumption between the concrete, the steel bars, and
the CFRP sheets. The beams were simply-supported and steel plates were added at the
support locations for even stress distributions.
2.4 Nonlinear Finite Element Analysis Computation
LS-DYNA single precision software program on Glenn (IBM 1350) Linux Cluster at the
Ohio Supercomputer Center was used to perform nonlinear explicit finite element analysis

525

of as-built and CFRP-strengthened concrete beams subjected to repeated drop hammer


impact loads.
Using 16 processors, the simulation time run for the CFRP-strengthened beams were 0.32
second where as for the as-built RC beams, the simulation times were 0.1 and 0.15 second,
for 1200 and 600 mm impacting block dropping height, respectively. The simulation times
varied due to the fact that as-built and CFRP-strengthened beams failed at different level of
repeated impact loading. LS-DYNA outputs were written for every 1xE10-6 second and
total CPU time to complete each run was 2.1 hours for each dropping impact load. The
value of 0.1 percent damping ratio was used to minimize computational time and to
stabilize the drop hammer during the rebound.
3. FINITE ELEMENT ANALYSIS RESULTS OF VALIDATED BEAM MODELS
Fujikake et al. (2009) conducted experimental study on the impact response of reinforced
concrete beams. The beam specimens labeled as S1322, S1616 and S2222 each with 600
mm and 1200 mm hammer drop heights were selected for validation of the proposed finite
element models. The concrete compressive strength of S1322, S1616 and S2222 beams
were 42 MPa with 10 mm maximum aggregate size. The steel bars were D13, D16, D22
and D10 and had yield strength of 397 MPa, 426 MPa, 418 MPa and 295 MPa,
respectively. D10 steel bars were used as stirrups. Beam S2222 was reinforced with four
D22 bars, two at the top and two at the bottom. Beam S1616 had similar arrangement of
reinforcement except the bars size was D16. Beam S1322 had two D13 compression and
two D22 tensile steel bars. The geometry and reinforcement details of validated FE models
of tested beams and CFRP layout are presented in Fig. 1.

Fig. 1. FE validated models of beams (a) geometry and reinforcement details; (b) CFRP
reinforcement layout (mm)
Dynamic impact loading was generated by freely dropping hammer mass of 400 kg on the
top face of the beam at mid-span location. Impacting speed of the hammer mass was
calculated at a height 10 mm away from the top face of the beam using the equation of
motion. The calculated speed was assigned as an initial speed of the drop hammer in FEA.
This approach was utilized to minimize the significant number and size of output files in
LS-DYNA, for such small time step of 1xE10-6 second when the hammer was dropped
from a height ranging between 0.6 to 4 m measured from the top of the beam. A shift in
abscissa of time history plots was applied to reflect the aforementioned approach. The
striking block had semi-spherical head with 90 mm radius as shown in Fig. 2.

526

Fig. 2. FEA mesh of reinforced concrete beam and impacting hammer


S1322

Force (KN)

Force (KN)

300

Test DIF
FEA DIF

200
100

0.01

0.02

0
0.00

0.03

Force (KN)

Test DIF
FEA DIF

100

0.01

0.02

0.02

0.03

(b) 1200 mm drop height

S1616

200

0.01

Time (Sec)

(a) 600 mm drop height


300
Force (KN)

Test DIF
FEA DIF

200

Time (Sec)

0
0.00

S1322

100

0
0.00

300

Test DIF
FEA DIF

200
100
0
0.00

0.03

S1616

300

Time (Sec)

0.01

0.02

0.03

0.04

Time (Sec)

(d) 1200 mm drop height

(c) 600 mm drop height

S2222

Force (KN)

Force (KN)

300
Test DIF
FEA DIF

200
100
0
0.00

300

S2222

200

Test DIF
FEA DIF

100

0.01

0.02

0.03

Time (Sec)

0
0.00

0.01

0.02

Time (Sec)

(e) 600 mm drop height

(f) 1200 mm drop height

Fig. 3. Dynamic impact force versus time history of the reinforced concrete beams
Experimentally measured contact impact forces between the reinforced concrete beams and
the drop hammer were used to validate the accuracy of the proposed finite element models.
Fig. 3 shows comparison of the experiment and FEA results of dynamic impact force
versus time history of S1322, S1616 and S222 beams for 600 mm and 1200 mm drop
heights. The dynamic impact time history plots are characterized by high amplitude pulse
527

followed by transient relatively low amplitude waveforms. The comparison of the finite
element analysis and experimental results showed that the peak dynamic impact forces
varied by 5, 6.5 and 8.71 % when the drop hammer was 600 mm and by 2.64, 3.68 and 3.12
% when the drop hammer changed to 1200 mm for S1322, S1666 and S2222 beams,
respectively. In general, the FE results showed more pronounced drops after the peak, due
to the hammer rebound. Finite element simulation and experimental results were in good
agreement with respect to the behavior and the peak response.
4. FINITE ELEMENT ANALYSIS RESULTS OF AS-BUILT AND CFRPSTRENGTHENED RC BEAMS UNDER REPEATED IMPACT LOADING
A complete CFRP wrap retrofit technique was applied along the span length to improve the
shear and flexural capacities of the beams and to mitigate concrete crushing failure at the
impact location on the top face of the beam. TyfoSCH-41 (Tyfo 2010a), a unidirectional
FRP composite with 0 carbon fiber orientation, was applied on A faces, at the bottom of
the beam for flexural strengthening, and at the top to protect the concrete from crushing at
the impact location (Fig. 1). TyfoSCH-41 sheet has a 1 mm thickness with a tensile
strength of 986 MPa in the fiber direction, tensile modulus of 95.8 GPa, and rupture strain
of 1 %. TyfoBCC (Tyfo 2010b) sheet with 0.86 mm thickness and 45 carbon fiber
orientation was applied on the beam side faces, B, for shear strengthening (Fig. 1).
TyfoBCC has a modulus of elasticity of 47.9 GPa, tensile strength of 661 MPa in the
fiber direction, and rupture strain of 1.4 %. The complete CFRP retrofit consisted of four
layers of TyfoSCH-41 and TyfoBCC CFRP wraps with 0/0/0/0 and +45/45/+45/-45 fiber orientations, respectively.
Table 1 shows the FEA results of as-built and CFEP-strengthened beams. Number of drop
hammerhead repeated impact loadings shown for as-built RC beams refers to the
occurrence of the flexural failure ,where as for the CFRP-retrofitted beams, it refers to the
number of applied repeated drop hammerhead impact loading before the drop hammer lost
its stability in FEA to exert further repeated impact loading. Mid-span deflection results for
as-built RC beams were extracted at the impact load level right before the failure.
In as-built beams, as compared to S1322 and S2222 models, S1616 beam experienced 33.5
and 47.13 % additional mid-span deflection at the drop height of 0.6 m, and 76.8 and
56.93% at the drop height of 1.2 m, respectively. This is due to higher tensile steel
reinforcement ratios of 2.46 % for S1322 and S2222 beams as oppose to 1.26 % value for
S1616 beam.
As compared to corresponding as-built beams, S1322, S1666 and S2222 CFRPstrengthened FEA models mid-span deflections were reduced by 84.21, 87.57 and 82.51 %
for 600 mm drop height and by 72.47, 83.79 and 75.88 % for 1200 mm drop height,
respectively. The reduced mid-span deflections are attributed to the CFRP strengthening.

528

Table 7. Comparison of as-built and CFRP-strengthened beams


Number of
Maximum
Hammerhead
drop
mid-span
Failure
Specimen type
dropping
hammerhead
deflection
mode
height (mm) repeated impact
(mm)
loadings
Flexure
S1322
600
3
37.49
S1322
1200
2
25.43
Flexure
600
3
50.04
Flexure
As-built RC S1616
beams
S1616
1200
2
44.96
Flexure
S2222
600
3
34.01
Flexure
S2222
1200
2
28.65
Flexure
S1322
600
16
14.16 (5.92)a
None
a
S1322
1200
16
25.42 (7.00)
None
CFRPa
S1616
600
16
15.73 (6.22)
None
strengthened
S1616
1200
16
25.28 (7.29)a
None
RC beams
a
S2222
600
16
13.9 (5.95)
None
S2222
1200
16
27.8 (6.91)a
None
a
values purposely are extracted at the same number of repeated loadings as for
corresponding as-built beams for comparison purposes.
Figure 4 shows the deflected shape and failure mode of as-built and CFRP-strengthened
S1322 beams. Eroded elements represent elements that were eminently cracked and no
longer had the capacity to carry any load. The color code ranging from blue to red
represents the damage level from none to severe, respectively. Flexural cracks at the center
of the beam and crushing of the concrete at the impact location were observed at the first
hammerhead impact loading for the as-built S1322, S1666 and S2222 beams. Rapid crack
propagation along the beam span and severe concrete crushing at the impact location
occurred at the subsequent repeated hammerhead impact loadings until the beams failed in
flexure. In the case of the CFRP-strengthened beams, only concrete cover was damaged at
the location of the hammerhead drop. No CFRP rupture at the bottom face or puncture at
the impact location was observed for these beams.

Fig. 4. S1322 beam damage profile at 600 mm dropping height (a) as-built beam at 3rd and
(b) CFRP-strengthened beam at 16th hammerhead drop.

529

5. CONCLUSIONS
The present study has shown that complete CFRP-wrapping is effective in improving the
response of concrete beams subjected to repeated drop hammerhead impact loading. The
CFRP-strengthening technique contained flexural cracks and concrete crushing damages
induced by repeated drop hammerhead impact loading. The CFRP reinforcement improved
mid-span deflection serviceability requirement in the range of 72.47 to 87.57 % and
increased flexural capacity of the beams. The CFRP-strengthening also eliminated brittle
failure mode that was observed in as-built RC beams and increased the durability of the
beams to sustain repeated drop hammerhead impact loading.
6. REFERENCES
1.

Tang, T., and Saadatmanesh, H. 2005. Analytical and experimental studies of fiberreinforced polymer-strengthened concrete beams under impact loading. ACI Structural
Journal, 102(1): 139-149.
2. Sukontasukkul, P., Nimityongskul, P., and Mindess, S. 2004. Effect of loading rate on
damage of concrete. Cement Concrete Research, 34(11): 2127-2134.
3. Banthia, N.P., Mindess, S., and Bentur, A. 1987. Impact behaviour of concrete beams.
Materieral and Structurres, 20(4): 293-302.
4. Arslan, A. 1995. Mixed-Mode Fracture Performance of Fiber-Reinforced Concrete
under Impact Loading. Materieral and Structurres, 28(8): 473-478.
5. Chakradhara Rao, M., Bhattacharyya, S.K., and Barai, S.V. 2011. Behaviour of
recycled aggregate concrete under drop weight impact load. Construction and Building
Materials, 25(1): 69-80.
6. ANSYS. 2009. ANSYS Users Manual, version 11, ANSYS, Inc., Canonsburg, PA.
7. LSTC. 2007a. LS-DYNA Users Manual, Version 971, Livermore Software
Technology Corporation, Livermore, CA.
8. LSTC. 2007b. LS-PREPOST 2.1 Software Program, Livermore Software Technology
Corporation, Livermore, CA.
9. Fujikake, K., Li, B., and Soeun, S. 2009. Impact Response of Reinforced Concrete
Beam and Its Analytical Evaluation. Journal of Structural Engineering, 135(8): 938950.
10. TYFO SCH-41 Composite. 2010a. Technical Data Sheet.
(http://fyfeco.com/products/pdf/tyfo%20sch-41%20comp.pdf).
11. TYFO BCC Composite. 2010b. Technical Data Sheet.
http://fyfeco.com/products/pdf/tyfo%20bcc%20comp.pdf.

530

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

BOND STRESS-SLIP MODEL FOR FRP REBARS IN CONCRETE


S. Quayyum1 and A. Rteil2
1
2

PhD Student, Department of Civil Engrg, North Carolina State University, USA
Assistant Professor, School of Engineering, The University of British Columbia, Kelowna, Canada

ABSTRACT
Fibre reinforced polymer (FRP) rebars have been used in the construction of reinforced
concrete structures instead of steel rebars due to their non-corrosive nature and high tensile
strength. However, a direct substitution between steel and FRP rebars is not possible due to
the difference in the physical and the mechanical properties of the two materials. In
particular, the bond behaviour of the FRP rebars in concrete is significantly different from
that of the steel rebars. Bond is a critical design parameter that controls the performance of
reinforced concrete members at both the serviceability and the ultimate limit states. In order
to prevent bond failure, an adequate anchorage length of the FRP rebar in concrete should
be provided. The anchorage length is derived using bond stress-slip law. In this study, a
bond stress-slip model was developed for helical lugged and spiral wrapped FRP rebars in
concrete. The proposed model was developed based on the available beam bond test data
reported in the literature, where the specimens failed by splitting of the concrete cover. All
the significant parameters were considered in deriving the model, including the effect of
concrete confinement provided by the transverse reinforcement.
1. INTRODUCTION
Recently, fibre reinforced polymer (FRP) reinforcing bars have been used instead of steel
reinforcing bars in the construction of reinforced concrete structures due to their high
tensile strength and non-corrosive nature. Still a direct substitution between FRP and steel
rebar is not possible due to the various differences in the mechanical and the physical
properties of the two materials. One of the main differences is the bond behaviour of FRP
rebars with concrete due to the non-isotropic material properties and the different surface
texture of the FRP rebars (ACI 440.1R-06).
To prevent bond failure in reinforced concrete members and to ensure complete transfer of
forces between the reinforcement and the concrete, the reinforcement should be adequately
anchored in the concrete. To determine the required anchorage length of the rebar, bond
stress-slip ( s ) law is needed. Although many formulations for bond stress-slip law were
proposed for steel rebars, the difference between steel and FRP dictates that an extensive
research effort is still needed to establish a bond stress-slip relationship for FRP rebars.
531

Moreover, the formulations of bond stress-slip relationship proposed so far for FRP rebars
have to be validated by experimental investigation and curve fitting of the experimental
data. Some experimental models for FRP rebars were reported in the literature (Malvar,
1994; Faoro, 1992; Rossetti et al., 1995; Cosenza et al., 1997; Tighiouart et al., 1998;
Focacci et al., 2000; Pecce et al., 2001; Aiello et al., 2007; Baena et al., 2009). It should be
emphasized that the analytical models available in the literature for the s relationship are
aimed at identifying a general law, which holds always by determining its parameters by
curve fitting of the experimental data. No specific formulations for different types of rebars
have been developed so far. Therefore, a generalized bond stress-slip law, which can be
applied to different types of FRP rebars has not been established yet (Cosenza et al., 1997).
The study reported herein aims at developing a general bond stress-slip relationship for
helical lugged and spiral wrapped FRP rebars based on the available experimental data
from the technical literature on beam bond test.
2. DESCRIPTION OF THE DATABASE
In this study, a database of 91 beam-type specimens was created on the bond stress-slip
relationship based on the available literature. A complete listing of all the data can be found
in Quayyum (2010). Only beam-type specimens that resulted in the concrete surrounding
the reinforcement being in tension were considered in the analysis. The data included glass,
carbon and aramid FRP rebars with different surface textures such as sand coating, spiral
wound and helical lugs. Of these 91 specimens, 23 specimens had bond failure through
splitting of concrete and 68 specimens had bond failure through rebar pullout. The 23
specimens that failed by splitting of concrete were used to develop the bond stress-slip law.
Of the 23 beam-type specimens that failed by splitting of concrete, 11 had helical lugged
FRP rebars and 12 had spiral wrapped FRP rebars. There was no reported specimen with
sand coated rebars which failed by concrete splitting. All of bars were cast as bottom bars.
Of the 23 beam specimens, 17 had transverse reinforcement. With respect to the type of
fibre, 7 specimens had AFRP rebars, 11 had CFRP rebars and 5 had GFRP rebars. The
concrete cover of all the specimens were between one and three bar diameters (
d b c 3d b ) and the embedment length of the specimens were less than twenty five bar
diameters ( lembed 25d b ).
3. BOND STRESS-SLIP MODEL
From the analysis of the database, it was observed that the bond stress-slip curve of the
FRP rebars in concrete shows two distinct branches-an initial ascending branch up to the
peak bond stress and a post-peak descending branch. Therefore, the bond stress-slip data
was splitted into two parts-one for the ascending branch of the bond stress-slip curve up to
the peak bond stress and the other one is for the descending post-peak branch of the bond
stress-slip curve. It was noted that to predict the bond stress-slip relationship of the FRP
rebar in concrete, it is necessary to know the peak bond stress ( m ) and the corresponding
slip ( sm ), because the ascending part ends at that point ( sm , m ) and the descending part
starts from the same point. In the present study, the peak bond stress was determined by
using the equation proposed by Quayyum and Rteil (2010). On the other hand, an

532

expression for the slip corresponding to the peak bond stress was derived by performing
regression analysis on the available data from the literature for which the slip
corresponding to the peak bond stress was reported.
3.1 Peak Bond Stress
Quayyum and Rteil (2010) accumulated a database of 177 beam-type specimens from the
available literature, which had bond failure through splitting of concrete cover. From the
analysis of the database, it was observed that the type of fibre and the rebar surface do not
have significant effect on the peak bond stress of FRP rebars in concrete, whereas
confinement provided by the transverse reinforcement increased the peak bond stress.
Using the same approach as Orangun et al. (1975), a linear regression analysis was
performed on the database to develop an equation to determine the peak bond stress by
considering the effect of the transverse reinforcement, which is expressed as follows in SI
units:
m
d
A
c
(1)
0.03 0.14 9.0 b 2.9 tr
db
lembed
snd b
f c
where, m is the FRP rebar-concrete peak bond stress, f c is the compressive strength of
concrete, c is the lesser of the cover to the center of the bar or one-half of the center-tocenter spacing of the bars being developed, d b is the bar diameter, lembed is the embedment
length of the bar in concrete, Atr is the area of transverse reinforcement normal to the plane
of splitting through the bars, s is the center to center spacing of the transverse reinforcement
and n is the number of bars being developed along the plane of splitting.
3.2 Slip Corresponding to Peak Bond Stress
To derive an expression for predicting the slip corresponding to the peak bond stress, 40
beam bond test data was accumulated from the available literature, where the specimens
failed by splitting of concrete. Of the 40 specimens, 6 specimens had AFRP bars, 24 had
CFRP bars and 10 had GFRP bars. There were 15 specimens, which had helical lugged
bars, 3 had sand coated bars and 22 had spiral wrapped bars. All the specimens had
concrete cover of between one and three bar diameters ( d b c 3d b ) and the embedment
length of the specimens ranged between 4 to 28 bar diameters ( 4d b lembed 28d b ). Of the
40 specimens, 28 had transverse reinforcement.
From the analysis of the experimental data for the slip corresponding to the peak bond
stress, it was observed that the type of fibre and the confinement provided by the transverse
reinforcement do not affect the slip corresponding to the peak bond stress of FRP rebars in
concrete, but the surface type of FRP rebars has significant effect on the slip corresponding
to the peak bond stress (Fig. 5). Therefore, it was necessary to split the data according to
the rebar surface. In addition, it was observed that the slip corresponding to the peak bond
stress is a function of the concrete cover, the bar diameter and the embedment length. A
linear regression analysis on the normalized cover (cover to the center of the bar divided by
the nominal bar diameter) and the normalized slip corresponding to the peak bond stress

533

was used to develop Equation (2) in SI units by considering the effect of the surface type of
FRP rebar.
sm

c
3 0.12 f c 0.6
(2)

lembed 100
db

where, sm is the slip corresponding to the peak bond stress and is a surface dependent
factor, which equals to 1 if the bar surface is helical lugged, 0.4 if it is spiral wrapped and
0.35 if it is sand coated. The proposed model showed good agreement with the
experimental data as shown in Fig. 6.
0.025

0.025

Helical Lugged

0.02

Helical Lugged

0.02

Spiral Wrapped

Spiral Wrapped

sm
l embed

Sand Coated

sm
l embed

Predicted

0.015

0.01

Sand Coated
0.015

0.01

0.005

0.005

0
0

c
db

0.005

0.01

0.015

Experimental

Fig. 5. Comparison of the normalized slip


corresponding to the peak bond stress for
FRP rebars having different surface texture.

0.02

0.025

sm
l embed

Fig. 6. Test vs. predicted normalized slip


corresponding to the peak bond stress.

3.3 Analytical Modeling of Bond Stress-Slip Relationship


The formulations for bond stress-slip relationship available in the literature are based on the
specimens which failed by rebar pullout from pullout specimens. In the present study, the
bond stress-slip relationship was proposed based on the beam bond test specimens which
failed by splitting of concrete. Splitting mode of failure was selected as splitting failure is
more common for the development length ( ld 30d b ) and the concrete cover ( d b c 3d b
) used in practice. A nonlinear regression analysis was performed on the normalized bond

s
stress ( ) and the normalized slip ( ) of the accumulated database to develop a
m
sm
generalized bond stress-slip relationship for the 23 beam-type specimens which failed by
concrete splitting.
Fig. 7 and Fig. 8 represent the experimental data along with the nonlinear regression results
for all the specimens having helical lugged FRP bars and spiral wrapped FRP bars
respectively. It was observed that for the ascending part of the bond stress-slip curve, both
helical lugged and spiral wrapped FRP bars showed the same behaviour (Fig. 7a and Fig.
8a) and therefore, the following equation was proposed for the ascending part of the bond
stress-slip relationship 0 s sm :

534

s

m sm

0.45

(3)

On the other hand, for the descending part of the bond stress-slip curves ( s sm ), there was
a slight difference in the behaviour of helical lugged and spiral wrapped FRP bars (Fig. 7b
and Fig. 8b). It was also noted that the bond stress-slip behaviour of FRP rebars for the
descending part of the bond stress-slip curve was nonlinear. Therefore, one generalized
equation was proposed for the descending part of the bond stress-slip relationship based on
a nonlinear regression analysis of the experimental data and it is expressed as:

s

m sm

(4)

where, is dependent on the rebar surface and its value is -0.56 for helical lugged FRP
rebars and -0.60 for spiral wrapped FRP rebars.
1.2

1.2

Regression Line

y x 0.45

Regression Line

0.8

y x 0.5582

0.8

0.6
Experimental Data
0.4

0.6

0.4

0.2

Experimental Data

0.2

0
0

0.2

0.4

0.6

0.8

1.2

s
sm

(a) Ascending Branch

s
sm

(b) Descending Branch

Fig. 7. Nonlinear regression of the experimental data of the bond stress-slip curves for
specimens with helical lugged FRP rebars failed by splitting of concrete.
1.2

1.2

Regression Line

y x 0.6034

0.8

Regression Line

y x 0.45

0.8

0.6
Experimental Data
0.4

0.6

0.4

0.2

0.2

0
0

0.2

0.4

0.6

0.8

1.2

s
sm

(b) Ascending Branch

Experimental Data

0
0

s
sm

(b) Descending Branch

Fig. 8. Nonlinear regression of the experimental data of the bond stress-slip curves for

535

specimens with spiral wrapped FRP rebars failed by splitting of concrete.


Therefore, based on the experimental data and the nonlinear regression results, the
proposed bond stress-slip relationship of helical lugged and spiral wrapped FRP rebars in
concrete is:
s 0.45
when 0 s sm
sm

(5

m s
when s sm
sm
where, m and sm are calculated from Equations (1) and (2) respectively, and
0.56 For helical lugged/ribbed bars
0.6 For spiral wrapped bars

Fig. 9 and Fig. 10 show a typical comparison of the predicted bond stress-slip curves with
the experimental results for four beam-type specimens of the database. The comparison of
the predicted and the experimental bond stress-slip curves for all the database is presented
in Quayyum (2010). It was observed that the predicted values showed good agreement with
the experimental data, especially for the ascending part of the bond stress-slip curve up to
the peak bond stress and the proposed relationship could capture the peak bond stress in
each case.
For helical lugged FRP rebars, the proposed equation for the ascending part of the bond
stress-slip relation showed a high adjusted determination coefficient (adjusted R-square)
value of 0.962 explaining 96.2% of the variability of the response. On the contrary, the
proposed equation for the descending part of the bond stress-slip relation showed a
moderate adjusted determination coefficient (adjusted R-square) value of 0.524 explaining
52.4% of the variability of the response. For spiral wrapped FRP rebars, the proposed
equation for the ascending part of the bond stress-slip relation showed a high adjusted
determination coefficient (adjusted R-square) value of 0.963 explaining 96.3% of the
variability of the response. On the contrary, the proposed equation for the descending part
of the bond stress-slip relation showed a moderate adjusted determination coefficient
(adjusted R-square) value of 0.663 explaining 66.3% of the variability of the response.
Therefore, based on the results of the analysis and comparison with the experimental
results, it can be concluded that the proposed bond stress-slip relationship can give a good
prediction of the bond stress-slip behaviour of helical lugged and spiral wrapped FRP
rebars in concrete, when the failure is initiated by splitting of concrete.

536

4.5

4.5
Experimental Beam 1

Experimental Beam 2

Predicted

Predicted

3.5

3.5

Bond Stress (MPa)

Bond Stress (MPa)

2.5
2
1.5

2.5
2
1.5

0.5

0.5
0

0
0

Slip (mm)

Slip (mm)

Fig. 9. Predicted vs. the experimental s curves for specimens with helical lugged
surface having splitting failure
4

5
4.5

Experimental Beam 2

3.5

Experimental Beam 1
4

Predicted

Predicted

3
Bond Stress (MPa)

Bond Stress (MPa)

3.5
3
2.5
2

2.5
2
1.5

1.5
1

1
0.5

0.5
0

Slip (mm)

Slip (mm)

Fig. 10. Predicted vs. the experimental s curves for specimens with spiral wrapped
surface having splitting failure
4. CONCLUSION
The present study emanated with an aim to develop a bond stress-slip law for helical lugged
and spiral wrapped FRP reinforcing bars in concrete environment. For this purpose, all the
experimental data on the bond stress-slip of FRP rebars in concrete was accumulated from
the literature and the database was analysed statistically. Based on the analysis of the
experimental data, a general bond stress-slip law was formulated. Moreover, to define the
formulated bond stress-slip law, a relationship to predict the slip corresponding to the peak
bond stress was proposed. The effect of the confinement provided by the transverse
reinforcement was taken into consideration and modification factors were proposed for
different surface characteristics of the FRP rebars. Based on the proposed bond stress-slip
relationship, a design equation can be derived to better evaluate the required development
length of FRP rebars in concrete, thus resulting in a more efficient and reliable design of
FRP-reinforced concrete structures.

537

5. REFERENCES
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.

ACI Committee 440., 2006. Guide for the design and construction of structural
concrete reinforced with FRP bars. ACI 440.1R-06, American Concrete Institute,
Farmington Hills, MI, 44 p.
Malvar, L.J. 1994. Bond stress-slip characteristics of FRP rebars. Report TR- 2013SHR, Naval facilities Engineering Service Center, Port Hueneme, California, 1994.
Faoro, M. 1992. Bearing and deformation behaviour of structural components with
reinforcements comprising resin bounded glass fibre bars and conventional ribbed steel
bars. Proceedings of the International Conference on Bond in Concrete.
Alunno Rossetti, V., Galeota, D., and Giammatteo, M.M. 1995. Local Bond Stress-Slip
Relationships of Glass Fibre Reinforced Plastic Bars Embedded in Concrete. Materials
and Structures, 28: 340-344.
Cosenza, E., Manfredi, G., and Realfonzo, R. 1997. Behaviour and modelling of bond
of FRP rebars to concrete. Journal of Composites for Construction, ASCE, 1(2): 40-51.
Tighiouart, B., Benmokrane, B., and Gao, D. 1998. Investigation of bond in concrete
member with fibre reinforced polymer (FRP) bars. Construction and Building
Materials, 12: 453-462.
Focacci, F., Nanni, A, and Bakis, C.E. 2000. Local bond-slip relationship for FRP
reinforcement in concrete. Journal of Composites for Construction, ASCE, 4(1): 24-31.
Pecce, M., Manfredi, G., Realfonzo, R., and Cosenza, E. 2001. Experimental and
analytical evaluation of bond properties of GFRP bars. Journal of Materials in Civil
Engineering, 13(4): 282-290.
Aiello, M.A., Leone, M., and Pecce, M. 2007. Bond performances of FRP rebarsreinforced concrete. Journal of Materials in Civil Engineering, 19(3): 205-213.
Baena, M., Torres, L., Turon, A., and Barris, C. 2000. Experimental study of bond
behaviour between concrete and FRP bars using a pull-out test. Composite: Part B, 40:
784-797.
Quayyum, S. 2010. Bond behaviour of fibre reinforced polymer (FRP) rebars in
concrete. MS Thesis, the University of British Columbia, Kelowna, BC, Canada.
Quayyum, S., and Rteil, A. 2010. Evaluation of bond strength of FRP rebars to
concrete from beam-type specimens. Journal of Composites for Construction, ASCE,
Submitted.
Orangun, C.O., Jirsa, J.O., and Breen, J.E. 1975. The strength of anchor bars: a
reevaluation of test data on development length and splices. Research Report 154-3F,
Center for Highway Research, The University of Texas Austin, Austin, TX, 78 p.

538

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

A CASE STUDY ON THE USE OF ADVANCED FIBER WRAP


COMPOSITES FOR REINFORCED CONCRETE REPAIR OF PORT
TERMINAL WHARF STRUCTURES
J.C. Percival1 and T.T. Jimenez2
1
2

Vice President Fibrwrap Construction Canada Limited


Vice President, Waterfront Structures, Fyfe Co. LLC

ABSTRACT
The use of Advanced Fibre Wrap Composite systems has been widely used in structures
during the past two decades. As the structural engineering community has adopted this
technology, so has the waterfront infrastructure community. This paper will cover the use
of this technology in the repair and strengthening of large wharf caissons. The structure was
severely deteriorated due to corrosive environmental conditions. Advanced Fibre Wrap
Composite Systems are currently being utilized to provide confinement and supplement the
flexural and shear capacities to existing caisson piles, beams and diagonal elements. The
use of special epoxy resins and proven concrete repair methods which were implemented as
part of the retrofit will be discussed. The aim of this paper will be to discuss the Advanced
Fibre Wrap Composite Systems method of repair to marine structures, and the associated
benefits in cost, time, and life cycle costing in relation to the financial challenges facing
marine infrastructure owners.
1. INTRODUCTION
Canadian port facilities have been operating now in many areas for over 100 years, while
changes have been made over time, many of the countries major port facilities have
operating structures that are over 65 years old. Some of these structures are of heavy timber
construction, some are rock filled timber caissons that have been encased in concrete, and
many others are of reinforced concrete construction. The technology of concrete for use in
marine applications has been around since the Romans, what has changed is the
composition of concrete, and the methodology of reinforcing large structures. The Romans
used rocks and plant fibres as reinforcement, these in more recent years were replaced with
iron, and then steel. Now we are using steel, stainless steel, epoxy coated steel, galvanized
steel, and FRP bars. Other than the man made FRP bars, all other methodologies have one
major flaw, given time, moisture, chlorides and other contaminants, they will oxidize. As
this happens, expansion of the materials causes the concrete to crack and expand, allowing

539

water to enter, which in turn accelerates the problem. In time spalling and delaminating of
concrete will occur, and structural failure will follow. See Fig. 1.

Fig. 1. Severely corroded ocean marine concrete structure within the tidal zone
2. OWNERS OPTIONS
2.1 Do Nothing
This may seem a ridiculous suggestion, but it has been the option of choice for many years,
and continues to be the easiest and most immediate cost effective measure available. The
reason we have deteriorated structures is solely due to this option. Each year facility
operators across the globe are expected to do more on shrinking capital and maintenance
budgets. Major expenditures are put on hold, and what is available is used to keep the daily
operation running. It is this situation that has made the do nothing approach the reality for
so many.
2.2 Traditional Repair
When the decision is finally made to invest funds to repair and retrofit, the infrastructure is
typically found in poor condition. Traditional repair methods depend on two factors, above
water, below water, or both. Above water repairs have been tackled by replacing lost
material with material of the same variety, sometimes with the inclusion of stainless steel,
galvanized, or epoxy coated bars. Given time stainless will corrode again, and galvanized
and epoxy coated bars are good when leaving the manufacturer, but after banging together
during shipping, and cutting and bending on site, much of the coating is lost as shown in
Fig. 2. When repairing below water the added expense of coffer dams have to be factored
in. These projects would typically require loss of use of portions or all of the facility,
creating significant loss to the clients revenue stream.

540

Fig. 2. Damage to epoxy coated rebar


2.3 The Option of FRP
Advanced fibre reinforced polymers were adopted from the military and aerospace
industries and into the civil infrastructure several decades ago. Timber, steel, and concrete
elements can all be wrapped in FRP, in many cases as a strengthening element to
deteriorated structures, and increasingly now as a environmental protection to structures old
and new that are susceptible not only to repeated fresh or salt water immersion, but also
freeze thaw cycles, ice, vessel impact, and attack from marine organisms. FRP is used as an
externally bonded element to supplement tension (steel) the engineer or record wishes was
already in the section. The additional capacity can be quantified in terms of shear or
flexural capacity, confinement (lateral confining pressure) and also provides protection by
encapsulating the structure and significantly reducing deterioration by preventing
contaminants from entering the repaired section. Fyfe has developed marine specific epoxy
resins that may be applied under water by specialized dive teams, certified and trained by
the manufacturer. All Fyfe Tyfo products are approved by Fisheries and Oceans Canada,
have passed Aquatic Toxicity Tests, and are free of volatile organic compounds.
3. GRAIN HANDLING WHARF, VANCOUVER CANADA
3.1 The Problem
In 2009 the owner operator of a major Canadian grain facility located in Vancouver
recognized the need to repair the major structural elements of their grain loading wharf.
The wharf was built pre WWII, it was expanded post war, and repaired in the 1970s and
80s using traditional methods. By mid 2005 these repairs had failed completely (Fig. 3 &
Fig. 4)

541

Fig. 3. Pre Repair Cross Beam Condition

Fig. 4. Pre Repair Caisson Condition

3.2 The Solution


The Engineer of Record for the Project was Sandwell Engineering of Vancouver, now
known as Ausenco Sandwell. Fibrwrap Construction was approached to review and offer
alternative repair technology based on an FRP system. After careful site investigation and
first hand reviews of previous projects with the owners representative the decision was
made by the owner and the engineer-of-record to repair the structure using FRP.
A preliminary design was created using CFRP longitudinally to replace lost tension steel
and GFRP to replace the lost hoop steel in the cross bracing. CFRP was designed to
supplement confinement of the pier corner piles. This design was based on the known
information prior to cleaning and demolition of all loose material.
Prior to installation of the composites, the structure had to undergo cleaning using high
pressure water jets, and then demolition of all unsound material. This was carried out from
the underside of the dock to the mud line. All removed material was collected and
transported offsite to protect the environment (Fig. 5 & Fig. 6). Upon completion of this
stage of the work the remaining steel was measured to determine the section loss and
adjustments were made to the CFRP and GFRP design.

Fig. 5. Cleaned X Brace Section

Fig. 6. Cleaned Caisson Section

542

To ensure no further deterioration would take place it was decided that the entire structure
would be wrapped in CFRP and coated with the Tyfo SW-1(S) marine epoxy which is
approved by Fisheries and Oceans Canada for use in sensitive marine environments.
3.3 The Repair
Upon completion of cleaning and determination of sectional loss of steel the repair process
could begin. All remaining steel was treated with a Corrosion Inhibiting System, Tyfo CIS
Type 1 & 2, designed to clean concrete, increase its pH and concrete compressive strength
(fc) as well as reform the passivity film on the steel reinforcement and protect it from
further corrosion. The sectional properties of the structure were then reformed with high
performance marine grout materials that were poured both above and below water.
To prepare the rebuilt sections for the application of the Tyfo composite system final
preparation was performed by hand using air driven grinders with diamond blades (Fig. 7).

Fig. 7. Final Preparation


The Tyfo S epoxy was used for saturation of fabrics above water, and the Tyfo SW-1S
for saturation of fabrics below water. Both epoxies are impregnated into the fabric using the
same saturation machine. It is imperative that fabrics are saturated correctly at this stage, if
the epoxy does not coat each FRP filament, the filaments that confirm the FRP sheet will
not be able to share the tension forces that they are designed to supplement and it may
cause fibre separation and potentially bonding failures. This will not be discovered prior to
bond pull tests to the ASTM D4541 (Fig. 8), or by sound testing. By using the Tyfo
saturator this process is guaranteed (Fig. 9).

543

Fig. 8. ASTM D4541 Test Apparatus

Fig. 9. Saturation of Fabric SW-1S

The Tyfo SCH 41 CFRP was applied to the cross bracing in the longitudinal direction first
using the Tyfo S epoxy. (Fig. 10) This application was above water during low tide
conditions and application was from working floats. Immediately after the Tyfo SEH-51A
GFRP was applied in the hope direction using the Tyfo SW-1(S) epoxy (Fig.11) The main
Caissons were wrapped using the same materials however no CFRP was required on these
sections (Fig.12).

Fig. 10. Application of Tyfo SCH-41

Fig. 11. Application of Tyfo SEH-51A

Fig. 12. Application of Tyfo SEH-51A to Caissons

544

3.4 The Results


The completed project has returned to the owner a structurally sound, environmentally
protected, structure that can look forward to many years of grain handling with little or no
maintenance. By encapsulating the structure the effects of salt water, abrasion by floating
debris such as logs, and marine organisms, has been significantly reduced. The life
expectancy of the finished structure can now be measured in decades once again.

Fig.13. the Completed caisson and cross beam sections.


The use of the Tyfo composite system also ensured a reduction in construction costs and no
disruption to the working wharf, further saving the client money. Additionally the
environmental approval from Fisheries and Oceans Canada ensured that the client had no
risk of any environmental mishap that could have negative consequences both financially
and socially.
4. REFERENCES
1.
2.
3.
4.
5.

TYFO. CIS. 2010. Technical Data Sheet. (http://fyfeco.com/products/pdf/tyfo%20cis1.pdf; http://fyfeco.com/products/pdf/tyfo%20cis-2.pdf ).


Tyfo S Epoxy. 2010. Technical Data Sheet. (http://fyfeco.com/products/pdf/tyfo%20s
epoxy.pdf).
TYFO SW-1S. 2010. Technical Data Sheet.
(http://www.fyfeasia.com/documents/datasheets/TYFO%20SW-1S%20Underwater
Saturant%20Epoxy%20Data%20Sheet.pdf).
TYFO SCH-41 Composite. 2010. Technical Data Sheet.
(http://fyfeco.com/products/pdf/tyfo%20sch-41%20comp.pdf).
TYFO SEH-51A Composite. 2010. Technical Data Sheet.
(http://fyfeco.com/products/pdf/tyfo%20seh-51a%20comp.pdf).

545

546

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

PROBABALISTIC DEVELOPMENT OF A LIFE CYCLE


INVENTORY (LCI) DATASET FOR PULTRUDED FIBER
REINFORCED POLYMER (FRP) COMPOSITES
S.M. Ali1, M.D. Lepech2 and J.P. Basbagill1
1
2

Graduate Student, Stanford University, Stanford, CA, USA


Assistant Professor, Stanford University, Stanford, CA, USA

ABSTRACT
Sustainability is increasingly coming to the forefront of many building designs and
infrastructure plans incorporating FRP composites. However, the metrics used to quantify
the environmental impacts of FRPs are not fully understood by manufacturers or engineers.
Although pultruded composite sections comprise the most widespread application of FRPs,
the underlying datasets necessary for determining the full life cycle impacts of these
materials have not been developed for North America. In this paper, the probabilistic
development of a life cycle inventory (LCI) dataset for pultruded FRPs is presented. A
survey of North American pultruders was conducted to measure the industry wide
material and energy inputs into the FRP pultrusion process. These were statistically
analysed to construct predictive models for material and energy consumption. Coupling
existing literature on component material impacts with results from this study, an LCI
model was developed for pultruded sections produced in North America including
structural shapes and lineals.
1. INTRODUCTION
The proliferation of building metrics that encourage sustainable design and construction has
been evidenced by widespread adoption of green building standards such as Leadership in
Energy and Environmental Design (LEED) [1]. However, many materials that go into these
buildings have environmental profiles that are unaccounted on a life cycle basis [2]. Life
Cycle Assessment (LCA) is a methodology that analyzes the various stages of a product or
process that includes raw material acquisition, material processing, manufacture and
assembly, use and service, retirement and recovery, and disposal to assess environmental,
social and economic impacts. Guidelines for performing an LCA are prescribed by the ISO
14040 4 series [3].
Responding to this need for more sustainable construction materials, fiber reinforced
polymer composites (FRPs) have found adoption in civil engineering infrastructure systems

547

[4]. Due to their high strength to weight ratio, slender sections, and good durability,
some view FRPs as a more sustainable material alternative to steel, concrete or aluminum.
Pultrusion is a low cost manufacturing process that has emerged as a mainstream method
for producing FRP composites for civil infrastructure. Pultrusion involves physically
pulling material through a manufacturing line. Fiber threads are pulled into a resin
impregnation system (open bath or direct injection) where resin is exposed to the fiber at
the polymerization temperature. The resin-filled fiber is then pulled through a die and cured
to produce a solid with constant cross section and desired mechanical properties. At the end
of the line, the continuous product is cut to length [5].
The ISO 14040 series outlines the four primary phases of an LCA; goal/scope definition,
life cycle inventory analysis (LCI), life cycle impact assessment (LCIA) and interpretation.
The life cycle inventory phase quantitatively measures the systems material and energy
inputs as well as emissions, wastes and byproducts that result [6]. An accurate life cycle
inventory is necessary to accurately quantify the environmental impacts of pultruded FRPs
[7]. The current primary dataset for pultruded FRPs was produced by the Association of
Plastic Manufacturers of Europe (APME), which has been adopted into the EcoInvent
database [8]. Peereboom et al. conducted a comparative study on PVC using six life cycle
inventory databases and found that such LCI datasets can have an enormous effect on the
LCA results and significant attention must be given to properly ensuring that the datasets
being used accurately represent the physical system being modeled [9]. As such, using
European LCI datasets is incorrect when modeling North American pultrusion due to
concerns of analysis boundaries, outdated process models, geographic differences and raw
material mismatches. This paper details the development of a probabilistic LCI dataset for
pultruded FRPs in North America.
2. MATERIALS AND METHODS
The United States pultrusion industry comprises roughly 5% of the total composites
industry by material output [10]. The Pultrusion Industry Council (PIC) represents North
American composites manufacturers and suppliers whose products are used in pultrusion.
For this LCI, a survey was conducted of PIC members for primary data collection. The
pultruders participating produce a total annual output in excess of 37 million kilograms of
material. This amount represented over 50% of the total yearly output of the pultrusion
industry [4]. In addition to the pultruders, material suppliers for both unsaturated polyester
resin and E glass, the two largest components of pultruded FRP, provided primary data for
their respective products.
The primary materials going into the pultrusion line and considered in this inventory
include reinforcing fiberglass, resin, filler, fire retardant, styrene, pigment, mold release
agent and catalyst. Unsaturated polyester resin (UPR) was used by all pultruders surveyed.
Styrene, needed for processing, comprises roughly a third of the UPR and is also added in
separately at the line. The filler material is calcium carbonate (limestone) and a small
amount of clay. Fire retardant additives include decabromodiphenyl oxide, antimony
trioxide, and aluminum trihydrate. Due to the small percentage of fire retardant additives
used (typically 4.6% by weight), and based on previous work indicating the minimal

548

toxicological effects of these materials, they were excluded from the analysis boundary.
[11]
Beyond quantifying the basic material inputs, the ISO 14040 framework prescribes
incorporating uncertainty analysis into LCA. Including uncertainty in the inventory analysis
helps to understand both natural variations in the measurements of the data (aleatory
uncertainty or variability) as well as underlying sources of systemic fluctuation that result
from not fully understanding the system under analysis (epistemic uncertainty) [12].
Although composites manufacturing incorporates the use of statistical techniques to ensure
high quality control standards that produce a final product that is as close to intended
design as possible, variability in material input quantities and uncertainty in impact factors
are primary contributors to uncertainty in this probabilistic life cycle inventory [13].
In addition to material inputs, a model was developed to calculate pultrusion process
energy. The total process energy is comprised of the resistive heat energy, the pulling
energy, and the saw energy. The resistive heat energy is used to heat the resin and keep it at
polymerization temperature. The pulling energy physically pulls the material through the
line. The saw energy is required to cut the finished product to length. The two primary
types of pultrusion equipment used are reciprocator and caterpillar machines. The
electricity requirements for a reciprocator machine are higher than that for a caterpillar due
to higher pulling energy. A reciprocator type machine was used to calculate the total
process energy. Equation 1 shows the relationship between the energy consumption and the
pultruded sections cross-sectional perimeter.
4.95

10

8.52

10

0.25 m P 6.5 m

(1)

where E is the total process energy (Joules/meter) and P is the perimeter (meter) of the
pultruded cross-section. This equation exhibited an R2 value of 0.965 based on an ANOVA
analysis including parameters of line speed, cross-sectional area, product weight and
perimeter. Energy consumption was controlled by the perimeter of the cross-section being
pultruded (P) with line speed, cross-sectional area and product weight providing little
additional correlation. Thus these parameters were excluded from the model.
Using the inventory data collected from the North American pultrusion industry, a Monte
Carlo simulation was performed in order to probabilistically characterize the total impacts
of the pultruded FRP supply chain. Figure 1 shows an overview of the variable
distributions, their relation in the supply chain and uncertainty propagation. This
hierarchical supply chain model is mathematically described in equations 2, 3 and 4.
IM
IPUL
II

pRM ,
IF RM
IM
pPUL ,
IPUL p

(2)
(3)
(4)

Where IMat is the impact distribution for various materials going into pultrusion.
pRMi,j is the proportion that individual raw materials are in pultrusion materials.
IFi is the impact factor for individual raw materials.
549

RMi is the weight distribution of individual raw materials.


IPUL is the impact distribution for different pultruders.
pPULk,l is the amount of each pultrusion material that specific pultruders use.
IIndustry is the impact distribution for the entire industry.
pl is the market share by volume for different pultruders.
Equation 2 (IMat)
Impact Distribution
Raw Material 1
for E-Glass

pR

M
1,
1

Impact Distributions
Raw Material i
for E-Glass
Impact Distribution
Raw Material 1
for Resin
(IF1RM1)
Impact Distribution
Raw Material i
for Resin
(IFiRMi)

RM

1 ,j

Equation 3 (IPUL)

E-Glass Impact
Distribution
(Ie-glass)

,1
L1
PU

Equation 4 (IIndustry)

Pultruder 1
(IPUL1)

Pultruder 2
(IPUL2)

pR

M
2 ,1
2,
M

pR

Pultruder 3
(IPUL3)

Resin Impact
Distribution
(Iresin)
Other Pultrusion
Materials
Impact Distribution
(IMatk)

Lk,
PU

Industry Average
Profile (IIndustry)

Pultruder l
(IPULl)

Fig. 1. Schematic representation of the probabilistic pultrusion supply chain used for Monte
Carlo simulation.
Table 1. Unsaturated polyester resin (UPR) material inputs with uncertainty assumption
parameters for triangular distributions based on industry survey of North American resin
production.
Material Component
Mode
Triangular Distribution
(kg/100 kg of resin)
Min/Max Parameters
Styrene
41.4
+/ 1.5%
Isophthalic Acid
20.7
+/ 0.75%
Propylene Glycol
14.3
+/ 0.75%
Maleic Anhydride
13.4
+/ 1.5%
Diethylene Glycol
8.3
+/ 0.75%
Ethylene Glycol
1.2
+/ 0.75%
The raw material distribution (RMi) characterized the quantity variation of each of the raw
materials for resin. The parameters for this distribution are shown in Table 1. The variable
proportion at which these individual raw materials are batched to formulate the resin is
represented by pRMi,j. The impact factor (IFi) is the midpoint indicator impact associated
with the production of specific materials used in pultrusion [14]. It is characterized by a
lognormal distribution derived from uncertainty parameters [15]. The impact of E-glass was
550

quantified using a similar method based on industry data. For all other materials, n and
pRMi,j were equal to 1. The uncertainty in the impact factor distributions (IFi) was derived
from equation 5.
(5)
where the uncertainty U factors were characterized using the Eco-Invent methodology:
U1 (reliability), U2 (sample size), U3 (technological correlation), U4 (geographic
correlation), U5 (temporal correlation), U6 (completeness), Ub (basic uncertainty) [16].
Table 2 below shows values that were used from the Eco-Invent database for characterizing
the uncertainty in impacts. This uncertainty parameter provides the 5% and 95% value from
the deterministic impact factor to create a lognormal impact factor distribution (IFi).
Table 2. Uncertainty parameters for environmental impact distributions of input materials.
Material Component
g 2
Fiberglass (E Glass)
1.25
Unsaturated Polyester Resin
1.45
Filler (Limestone)
1.33
Filler (Clay)
1.17
Styrene
1.70
Pigment
1.13
Mold Release Agent
1.20
Catalyst
3.16
Table 3. Raw Material Input Distribution parameters based on North American pultrusion
industry survey
Material Component
Material Input Distribution (kg/100kg of composite)
Fiberglass (E Glass)
Normal ( = 59.3, = 1.3)
Unsaturated Polyester Resin
Normal ( = 27.0, = 1.2)
Limestone: Normal ( = 6.8, = 0.3)
Filler
Clay: Uniform (1.0 7.0)
Styrene
Uniform (0.5 2.0)
Pigment
Normal ( = 0.75, = 0.05)
Mold Release Agent
Normal ( = 0.42, = 0.04)
Catalyst
Normal ( = 0.38, = 0.04)
Equation 2 was used to calculate impact distributions (IMat) for each of the materials going
into the pultrusion process. These distributions were then used in Equation 3 to create
individual material profiles for the various pultruders surveyed. Equation 3 shows this
calculation. The specific amount of each pultrusion material that pultruder l used was
characterized by pPULk,l. Industry distributions for pPULk,l values are shown in Table 3. The
variable pPULk,l was then multiplied by the impact distribution (IMatk) to calculate the impact
contribution of each input material towards the total FRP for pultruder l. This process was
repeated for each raw material and the sum of all of these products results in an impact
distribution for pultruder l (IPULl). The same was done for each pultruder. With individual
551

impact distributions for each pultruder (IPULl), the total impacts for the entire industry were
characterized by summing the product of the individual pultruder distributions (IPULl) and
the market share of the corresponding pultruder (pl). The industry average impacts were
calculated using Equation 5.
3. RESULTS & DISCUSSION
Monte Carlo simulation was used to produce distributions for Eco-Indicator 95 impact
categories. Specifically, greenhouse gas, winter smog, summer smog, ozone layer, heavy
metals, eutrophication, energy resources, carcinogens and acidification were calculated.
Figure 2 shows each of these impact categories Monte Carlo values as well as a fitted
lognormal distribution. Each of these impacts are on a per kilogram of pultruded FRP basis.
For example, figure 2(a) shows that production of one kilogram of FRP would have a mean
value of 0.94 kg of carbon dioxide emitted with a standard deviation of 0.12 kg. The
statistical parameters that characterize the normal distributions are shown in Table 4.

(a)

(b)

(c)

(d)

(e)

(f)

(g)
(h)
(i)
Fig. 2. Monte Carlo distributions for impact categories for pultruded FRP per kilogram
material output; (a) greenhouse gas; (b) winter smog; (c) summer smog; (d) ozone layer; (e)
heavy metals; (f) eutrophication; (g) energy resources; (h) carcinogens; (i) acidification
For every impact category, 100,000 trials were run. Previous Monte Carlo simulations for
LCA indicate using a number of trials between 100 and 30,000 to produce stable impact
distributions. [17-19] The fitted distribution for each impact category was lognormal. Table
4 also shows the Kolmogorov Smirnov (KS) parameter for measuring lognormality. [20]
KS values less than 0.03 indicate an acceptable match between the fitted lognormal
distribution and Monte Carlo values. This threshold was achieved for all distributions.

552

Table 4. Impact categories for 1 kilogram of composites produced in North America.


Standard
KS
Material Component
Distribution
Mean,
Deviation,
Parameter
Greenhouse Gas (kg CO2)
Lognormal
0.94
0.12
0.0038
7
8
Ozone Layer (kg CFC 11)
1.5

10
1.7

10
0.0027
Lognormal
Acidification (kg SO2)
0.0032
0.0004
0.0040
Lognormal
4
5
Eutrophication (kg PO4)
5.9 10
5.0 10
0.0024
Lognormal
Heavy Metals (kg Pb)
1.2 10 5
1.6 10 6
0.0056
Lognormal
8
9
Carcinogens (kg B(a)P)
3.9 10
3.2 10
0.0020
Lognormal
4
5
Summer Smog (kg C2H4)
4.1

10
6.0

10
0.0046
Lognormal
Winter Smog (kg SPM)
0.0024
0.0003
0.0038
Lognormal
Energy Resources (MJ LHV)
11.0
1.6
0.0035
Lognormal
4. CONCLUSIONS AND OUTLOOK
LCI data collection, probabilistic characterization of the dataset, and calculation of impacts
are critical steps towards more fully understanding uncertainty within LCA and in
disseminating data for future LCAs of North American pultruded composites. Regardless of
whether or not an LCA is being performed deterministically or probabilistically, this
dataset serves as a significant improvement over the existing European datasets that are
currently used to characterize pultruded FRPs in North America.
Significant further work can be done to build on this research. Further investigation of input
parameters for probabilistic LCA is necessary. The probabilistic model supply chain should
be refined to account for conditional probabilities that occur as a result of existing producer
and consumer contracts. This work can also be further improved by increasing the number
of industry participants. While 50% of the pultrusion industry was surveyed, the remaining
portion could be surveyed in order to produce an inventory that better represents the
industry.
5. REFERENCES
1.
2.
3.
4.
5.
6.
7.

U.S. Green Building Council (USGBC). 2009. Green Building Facts. (


http://www.usgbc.org/)
Horvath, A. 2004. Construction Materials and the Environment. Annual Review of
Environment and Resources, 29: 181-204.
ISO. 1997. ISO 14040: Environmental Management Life Cycle Assessment
Principles and Framework.
Busel, J. 2008. State of the North American Pultrusion Industry. Composites
Manufacturing, pp. 28-33; pp. 53-54.
T.F. Starr, Pultrusion for engineers, Woodhead Publishing, 2000.
ISO. 1998. ISO 14041: Environmental management Life Cycle Assessment Goal
and Scope Definition Inventory Analysis, Springer.
Coulon, R., Camobreco, V., Teulon, H., and Besnainou, J. 1997. Data quality and
uncertainty in LCI. The International Journal of Life Cycle Assessment, 2: 178182.

553

8.
9.
10.
11.
12.
13.
14.
15.
16.

17.
18.
19.
20.

APME. 2010. Eco-profiles and Environmental Declarations: LCI Methodology and


PCR for uncompounded polymer resins and reactive polymer precursors.
Peereboom, E.C., Kleijn, R., Lemkowitz, S., and Lundie, S. 1998. Influence of
Inventory Data Sets on Life-Cycle Assessment Results: A Case Study on PVC. Journal
of Industrial Ecology, 2: 109-130.
Stewart, R. 2002. Pultrusion industry grows steadily in US. Reinforced Plastics, 46:
3639.
Basbagill, J.P., Lepech, M.D., and Ali, S.M. 2010. Human Health Impact as a
Boundary Selection Criterion in the Life Cycle Assessment of Pultruded Fiber
Reinforced Polymer Composite Materials. 30 p.
Baker, J.W. and Lepech, M.D. 2009. Treatment of Uncertainties in Life Cycle
Assessment. Proceedings of the 10th International Conference on Structural Safety and
Reliability (ICOSSAR09), 8 p.
Li, H., Foschi, R., Vaziri, R., Fernlund, G., and Poursartip, A. 1991. Probability-Based
Modelling of Composites Manufacturing and its Application to Optimal Process
Design, Journal of Composite Materials, 36: 1967-1991.
Owens, J.W. 1996. LCA Impact Assessment Categories. The International Journal of
Life Cycle Assessment, 1: 151158.
Reap, J., Roman, F., Duncan, S., and Bras, B. 2008. A survey of unresolved problems
in Life Cycle Assessment. Part 2: Impact Assessment and Interpretation. The
International Journal of Life Cycle Assessment, 13: 374388.
Frischknecht, R., Jungbluth, N., Althaus, H.J., Doka, G., Dones, R., Heck, T. Hellweg,
S., Hischier, R., Nemecek, T., Rebitzer, G., and Spielmann, M. 2005. The ecoinvent
Database: Overview and Methodological Framework. The International Journal of Life
Cycle Assessment, 10: 39.
Sonnemann, G.W., Schuhmacher, M., and Castells, F. 2003. Uncertainty assessment by
a Monte Carlo simulation in a life cycle inventory of electricity produced by a waste
incinerator. Journal of Cleaner Production, 11: 279292.
Ciroth, A., Fleischer, G., and Steinbach, J. 2004. Uncertainty calculation in life cycle
assessments. The International Journal of Life Cycle Assessment, 9: 216226.
Lloyd, S.M. and Ries, R. 2007. Characterizing, Propagating, and Analyzing
Uncertainty in Life-Cycle Assessment: A Survey of Quantitative Approaches. Journal
of Industrial Ecology, 11: 161-179.
Lilliefors, H.W. 1967. On the Kolmogorov-Smirnov test for normality with mean and
variance unknown, Journal of the American Statistical Association, 62: 399402.

554

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

STUDY ON MOISTURE FICKIAN DIFFUSION PROCESS OF A


PULTRUDED FRP COMPOSITE MATERIAL UNDER HOT/WET
ENVIRONMENT
X. Jiang1, H. Kolstein2 and F.S.K. Bijlaard3
1

PhD Candidate, Faculty of Civil Engineering and Geosciences, TU Delft, the Netherlands
Associate Professor, Faculty of Civil Engineering and Geosciences, TU Delft, the Netherlands
3
Professor, Faculty of Civil Engineering and Geosciences, TU Delft, the Netherlands
2

ABSTRACT
Pultruded FRP composite material was increasingly used in the applications of civil
infrastructures, especially for bridge decks. Thus, the long-term life of these materials in
hostile environments is critical. In this paper, experimental studies have been undertaken to
characterise the moisture diffusion of a pultruded FRP composite material. Gravimetric
experiments were carried out for the rectangular FRP composite specimens, which were
exposed to the 40C and 96% RH environmental aging condition. Moisture uptake curves
as a function of time were obtained. Correspondingly, based on the moisture diffusion
theory, the diffusion coefficients of FRP composite material parallel to the pultrusion
direction were identified. It was found that, the equilibrium moisture contents of the
different part specimens are close to each other, from 0.35% to 0.38%. But the moisture
diffusion coefficients Dz differ between sealed specimens and unsealed specimens and also
differ in the specimens from different parts. Depending on the area ratios, the moisture
diffusion process of unsealed specimens is three-dimensional, while sealed specimens
behaved as a one-dimensional diffusion. All in all, the pultruded FRP composite material in
this study has the higher diffusion coefficient and much lower moisture saturation level,
which prove them a good resistance of moisture corrosion in a hot/wet environment.
1. INTRODUCTION
Fiber-reinforced polymer (FRP) materials, especially glass-fiber-reinforced polymer
(GFRP) pultruded profiles are being increasingly used in the application of civil
engineering as a competitive alternative to traditional materials, such as concrete, timber
and steel. The growing acceptance of FRP structures can be attributed to their pronounced
advantages: the high strength to weight ratio can allow them well to automated fabrication
techniques, easy and rapid installation, as well as reducing self- weight; robust and durable
FRP material with the strong resistance of corrosion can be economical and minimizes

555

maintenance requirements. One conspicuous application of FRP material is FRP bridge


decks (Fig. 1).

(a) ASSET (Fiberline, Denmark)

(b) DuraSpan (Martin Marietta Composite, USA)

(c) Kansas Structural Composites (d) Hardcore Composites VARTM cellcore deck
Fig. 1. FRP bridge decks
With regard to durability of FRP composite material, moisture ingression is an important
factor. The mechanical properties of FRP materials may suffer when the material is
exposed to moisture for long periods of time. Moisture is absorbed and transported via
diffusion through the matrix as well as interface between fibers and matrix. Absorbed
moisture can cause pronounced changes in modulus, strength, and strain to failure [1]. So,
understanding the whole diffusion process by which moisture enters a FRP composite is
critical to identify the location of damage, analysis the mechanical degradation and predict
the residual strength and service life of FRP structures. Understanding the moisture
diffusion process means to determine the moisture concentration distribution throughout
sections of FRP composite material as a function of time. Typically, gravimetric
experiments [2] [3] are employed to measure the moisture diffusion coefficients by
recording the weight of absorbed water in a specimen as a function of time.
The objective of this investigation was to understand the moisture diffusion process of a
pultruded FRP composite material with a short-term exposure to a hot/wet environment.
2. MOISTURE DIFFUSION THEORY
Berens and Hopfenberg [4] proposed a moisture absorption model for polymeric materials,
which was a linear superposition of independent contributions from Fickian diffusion and
polymeric relaxation. The total amount of absorption at time t, Mt, may be expressed as
M t M t ,F M t ,R

556

(1)

where Mt,F and Mt,R are the contributions of the Fickian and relaxation processes,
respectively, at time t.
Most of the studies on moisture diffusion in FRP composites rely on one-dimensional
Fickian process, whose equation is expressed as:
c
2c
Dz 2
t
x

(2)

where c represents the moisture concentration, x the space coordinate measured parallel to
the diffusion, and Dz the moisture diffusion coefficient in the x direction. Dz is supposed to
be independent of the spatial and temporal coordinates. For a plate of infinite dimensions,
the total mass of moisture absorbed at time t can be expressed:

8
M t , F M , F 1 2

(2n 1)
n 0

Dt

exp z2 2 (2n 1) 2
h

(3)

where M,F is the equilibrium amount of absorption in the unrelaxed polymer, and h is the
normal thickness in direction of diffusion. The theoretical Fickian diffusion process is
shown in Fig.2.

Fig. 2. Fickian diffusion process


For the initial linear part of Fickian diffusion curve ( Dz t / h 0.28 ), the identification of
the moisture diffusion coefficient Dz is performed as follows. The gravimetric curve (Fig.
2) is used to determine M,F. Then, Dz is calculated from the linear part using two points at
times t1 and t2.

h
Dz
4M
, F

M 1, F M 2, F

t t
1
2

(4)

This simple procedure is mostly used in practice, although the hypothesis on which it relies,
i.e., the fact that the plate is infinite in the plane directions, is incorrect for the generally

557

small specimens used in the conditioning chambers. The aforementioned calculation


procedures can be used only if the moisture diffusion is by a typical Fickian process.
Generally, Fickian diffusion takes place at low temperatures and for materials exposed to
humid air. Deviations from Fickian diffusion occur at elevated temperatures and for
materials immersed in liquids, which could be attributed to polymer relaxation. The
relaxation process is assumed to be first order in the concentration difference which drives
the relaxation. The differential equation for the relaxation process is therefore:

dM t , R
dt

k ( M , R M t , R )

(5)

where k is the relaxation-rate constant and M,R is the ultimate amount of sorption due to
relaxation. Integration of equation (3) leads to:
M t , R M , R 1 exp(kt )

(6)

Substitution of equations (3) and (6) into equation (1) results in:

8
M t M , F 1 2

(2n 1)
n 0

Dt

exp z2 2 (2n 1) 2 M , R 1 exp(kt ) (7)


h

Equation (7) is plotted in Fig. 3 as a linear superposition of the first term (Fickian diffusion)
and the second term (polymeric relaxation) of the equation. The linear portion of the total
absorption is almost identical to that of Fickian absorption until the Fickian curve starts to
bend over, indicating that the effect of relaxation on the initial absorption is not significant.
The second part of the combined absorption starts at about the maximum Fickian
absorption, when the diffusion rate is substantially decreased and the contribution of the
polymeric relaxation to the absorption is steadily increased. The final saturation at
equilibrium is represented by the sum of the maximum Fickian absorption, Mt,F and the
maximum relaxation absorption, Mt,R [5].

Fig. 3. Theoretical absorption curves due to Fickian diffusion and polymeric relaxation

558

3. EXPERIMENTAL STUDY
3.1 Specimens and Test Process
FRP specimens for the gravimetric experiments are cut from the ASSET FRP deck element
(Fig. 4) produced by Fiberline Composites A/S, which is the same as that used for the 2002
constructed West Mill Bridge, UK [6] and the 2008 constructed road bridge in Friedberg,
Germany [7]. The lay-up consists of longitudinal rovings surrounded, continuous strand
mat and a surfacing veil. The shapes consist of approximately 62% E-glass fibers (volume
fraction) and an isophthalic polyester matrix. The specimens are cut in the transverse
direction of pultrusion. According to the different lay-up thicknesses of inner web, outer
web and flange parts, the roving volume fraction and the relative volume of surface mat
layer are not identical. It can be predicted that the diffusion properties of different sections
should differ. Thus, it is necessary to conduct separate diffusion experiments on the inner
web, outer web and flange parts. Base on the moisture diffusion theory, one-dimensional
diffusion will only behave on the infinite plate with the thickness of h. But for this
experiment, small specimens were employed, due to the limited geometry of FRP bridge
deck element. So another group of specimens whose surrounding surfaces were sealed with
the sealant Sikafloor-156 were also involved in this experiment. The diffusion coefficients
of sealant material are several orders of magnitude less than those of pultruded FRP
composite material. Thus, it can be assumed that the sealed surfaces are impermeable to
moisture. In this way, the sealed specimens could be considered to behave as onedimensional moisture diffusion. Results of sealed specimens can be compared with the
unsealed specimens to find the difference between one-dimensional diffusion and threedimensional diffusion. Four replicates of each type of specimens were used in the specific
aging condition to minimize the deviation of experimental data. The aging condition with
the temperature of 40
and relative humidity of 96% was selected as a hot/wet
environment for the application of bridge structures, which was kept to be constant during
the whole process of testing. The average dimension of each sample type is given in Table
1. Prior to putting specimens into the environmental chamber, all the specimens were dried
in an oven and the weight of specimens were periodically checked until no changes in
weight occur.
Table 1. Average dimension of specimens
Section

Length(mm)

Width(mm)

Thickness(mm)

Inner web
Outer web
Flange

100.13
100.08
100.08

8.73
8.14
16.03

2.70
2.76
2.82

559

a) Outer web

b) Inner web

c) Flange

Fig. 4. FRP ASSET bridge deck element and test specimens


To measure the change in weight, each specimen was removed from the environmental
conditioning chamber, weighed quickly using a precision balance and then returned to the
chamber. In the first day, the specimens were weighed frequently, but the time interval was
changed to be one or two days as the rate of weight growing slowly
3.2 Results and Discussion
The average moisture uptake curves of sealed and unsealed specimens are shown in Fig.5
where Mt/ M is plotted vs. t to show the initial linear diffusion curve. All of the FRP
specimens aged in the 40 and 96% RH environmental condition reached the saturation
level within two days. During this period, the linear trends of moisture ingression into FRP
material are obvious in the data, which comply well with the aforementioned Ficks
moisture diffusion theory. After analyzing these data, the diffusion coefficients and
maximum moisture content at equilibrium were calculated by equation (4) and summarized
in the Table 2.
As shown in table 2, there is no considerable variability in the data of moisture equilibrium
contents of the different part specimens, with the flange absorbing about 0.35% and the
inner and outer webs absorbing about 0.38%. These values are extremely low, comparing
with other types of polyester-glass composite materials summarized by [3], which proved a
good resistance to moisture corrosion. And for the specimens from the same part, the
moisture equilibrium contents are very close to each other. It indicates that, no matter the
560

surrounding surfaces of specimens are sealed or unsealed, the moisture equilibrium content
will not change. It only depends on the FRP material itself. For the diffusion coefficients,
they are much higher than other types of FRP materials [3]. The diffusion process only took
two days to achieve the moisture equilibrium levels of every specimen. For both sealed
specimens and unsealed specimens, the moisture diffusion processes in the flange part are
always faster than in the other two parts, which indicate that larger roving area in the flange
specimens can contribute more to fast moisture diffusion. For the specimens from the same
part, there is a significant difference in moisture diffusion coefficients between sealed and
unsealed specimen for all components, which means the unsealed specimens behaved as a
typical three-dimensional diffusion. The moisture diffusion coefficient Dz of sealed flange
specimens is 30.4% lower than that of unsealed specimens, the same comparative values
for inner web and outer web parts are 34.3% and 42.3% respectively. These results are
reasonable when comparing with area ratios (obtained from equation 8) of each specimen.
Since, specimens with higher area ratio (0.830497, flange) should result in a more likely
one-dimensional moisture diffusive behavior and vice versa. From the Fig. 5, a short part of
the contribution from polymeric relaxation can be attained, but more long-term
experimental data are needed to draw other conclusions.
Area ratio =

Area transverse to pultrusion direction


Area of all the surfaces

(8)

Table 2. Experimentally obtained moisture diffusion coefficients


Specimen
Inner web
Inner web sealed
Outer web
Outer web sealed
Flange
Flange sealed

Dz (mm2/s)
-5

8.75910
5.05710-5
9.22610-5
6.05810-5
9.54410-5
6.64410-5

M (%)

Area ratio

0.383
0.379
0.376
0.393
0.348
0.351

0.748367
1
0.731719
1
0.830497
1

Fig. 5. Moisture uptake curves

561

4. CONCLUSION
The main objective of this paper was to investigate the moisture Fickian diffusion process
parallel to the pultrusion direction of FRP composite material, by using short-term
gravimetric experiments exposed to a hot/wet environmental condition. The moisture
diffusion theory of FRP composite material is introduced succinctly, which form the basic
background for the calculation of moisture diffusion coefficients. The experimental results
show that the equilibrium moisture contents of the different part specimens are close to
each other, from 0.35% to 0.38%. But the moisture diffusion coefficients Dz differ between
sealed specimens and unsealed specimens and also differ in the specimens from different
parts. Depending on the area ratios, the moisture diffusion process of unsealed specimens is
three-dimensional, while sealed specimens behaved as a one-dimensional diffusion. All in
all, the pultruded FRP composite material used in this experimental study has the higher
diffusion coefficient and much lower moisture saturation level, which prove them a good
resistance of moisture corrosion in a hot/wet environment. More experimental
investigations on the mechanical degradations and physical degradations of FRP composite
material exposed to hot/wet environments will be conducted in the near future.
5. REFERENCES
1.
2.
3.
4.
5.
6.
7.

Springer, G.S., Sanders, B.A., and Tung, R.W. 1980. Environmental effects on GFRP
polyester and vinyl ester composites. Journal of Composite Materials, 14(3): 213-223.
Pierron, F., Poirette Y. and Vautrin A. 2002. A Novel Procedure for Identification of
3D Moisture Diffusion Parameters on Thick Composites: Theory, Validation and
Experimental Results. Journal of Composite Materials, 36(19): 2219-2243.
Post, N.L., Riebel, F., Zhou, A., Keller, T., Case, S.W. and Lesko, J.J. 2009.
Investigation of 3D Moisture Diffusion Coefficients and Damage in a Pultuded E/glass
Polyester Structural Composite. Journal of Composite Materials, 43: 75-96.
Berebs, A.R. and Hopfenberg, H.B. 1977. Diffusion and relaxation in glassy polymer
powders: 2. Separation of diffusion and relaxation parameters. POLYMER, 19: 489495.
Shao, Y. and Kouadio, S. 2002. Durability of Fiberglass Composite Sheet Piles in
Water. Journal of Composites for Construction, ASCE, 6(4): 280-287.
Luke, S., Canning, L., Collins, S., Knudsen, E., Brown, P., Taljsten, B., and Olofsson,
I. 2002. Advanced composite bridge decking systemproject ASSET. Structural
Engineering International, 12(2): 76-79.
Knippers, J., Pelke, E., Gabler, M. and Berger, D. 2010. Bridges with Glass FibreReinforced Polymer Decks: The Road Bridge in Friedberg, Germany. Structural
Engineering International, 4: 400-404.

562

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

DURABILITY OF FRP-STRENGTHENED RC MEMBERS UNDER


VARIOUS LOADINGS
A. Parvin1 and A. Kulikowski2
1
2

Associate Professor, Department of Civil Eng., The University of Toledo, Toledo, USA
M.S. Student, Department of Civil Eng., The University of Toledo, Toledo, USA

ABSTRACT
The investigations in the last two decades have shown that the fiber reinforced polymer
(FRP) materials offer an ideal and durable solution as a replacement of steel bars or as an
external reinforcement for steel reinforced concrete members. The durability of structures is
influenced by various factors including the effect of environment and the applied load
types. The deterioration due to environmental changes, fatigue, natural or manmade
extreme events may result in serious damage to infrastructures and loss of life. This paper
presents an overview of use of FRP materials for reinforced concrete (RC) members with
particular emphasis on RC columns that are subjected to environmental changes, fatigue,
and extreme loadings such as vehicular collisions, blast, and fire. Both experimental and
finite element analysis results of available studies in the literature are discussed.
Additionally, areas where more research is needed to be performed are identified.
1. INTRODUCTION
Columns in bridges and frame structures can experience large loads during extreme events
such as vehicular collisions, blast, fires, earthquakes, and tsunamis. Additionally, extreme
environmental changes including freeze thaw and exposure to excessive heat can have
negative effects on the strength and performance of reinforced concrete columns as well.
Fatigue and cyclic loadings also contribute to RC columns weakening and failure. The
present paper involves the review of recent studies that have taken place concerning the
durability of FRP applications for RC members with focus on columns under extreme
loading conditions.
2. DURABILITY OF FRP-STRENGTENED STRUCTURES
In the following sections, effects of extreme loadings, environmental changes, and fatigue
on the durability of FRP-strengthened structures are discussed in detail.

563

2.1. Vehicular Collisions


With consistently increasing traffic in recent years, vehicular collisions with bridge
columns have become more of a prevalent issue. Most column designs account for static
loading only, while an impact load due to a vehicle collision is highly dynamic. Several
studies have been conducted concerning the dynamic effects of a high impact vehicle
collision with bridge piers and columns (El-Tawil et al 2005; Ferrier and Hamelin 2005;
Tsang and Lam 2008; Thilakarathna et al 2010).
El-Tawil et al (2005) utilized finite element models to study the effects of an accidental
collision between a truck and a bridge pier. The finite element simulation utilized a 14-kN
Chevy Truck and a 66-kN Ford Truck travelling at varying speeds between 55 and 135 kph
and colliding with two circular RC columns and piers with different sizes. The finite
element results indicated the shear failure of smaller circular piers subjected to the impact
load of the 60-kN Ford truck. This study revealed that there are certainly many existing
bridges that could be deficiently designed in the case of vehicular impact.
In recent years, besides rehabilitation of deficient RC columns, FRP materials have also
been used as more of a protective measure against vehicular collisions. Ferrier and
Hamelin (2005) performed experimental investigation on as-built and CFRP-strengthened
RC beams and columns. The specimens were subjected to static and dynamic impact loads.
For the static load tests that were conducted on three RC beams, the ultimate load of the
CFRP-strengthened specimen was 62 % higher than the as-built specimen. In the dynamic
test, it was observed that the CFRP-strengthened RC column load capacity was 88% higher
than an as-built specimen. Through both the static and dynamic tests, it was found that the
use of CFRP material significantly increased the strength of RC columns under impact
loading.
Tsang and Lam (2008) considered the fact that there are many structures with very slender
columns that are located along the roadways that can be possibly subjected to vehicle
collisions. This study utilized an energy absorption method, taking into account the impact
energy absorbed by the column and by the vehicle itself. Through the dynamic analysis, it
was found that in the case of an impact, the moment capacity of a slender RC column was
doubled if the dimensions of the column were increased by 20 %, while keeping the
reinforcement ratio constant. Furthermore, the finite element RC column model showed
greater ultimate capacity when the dynamic energy method was utilized to simulate the
vehicular collision rather than a simple static calculation.
Thilakarathna et al (2010) realized the fact that the dynamic response due to a collision is
different for every single model of vehicle on the road, making it a necessity to develop an
impact model that could cover a wider range of crash response data.
2.2. Blast
In recent years, there has been a shift in blast proofing new and existing civil infrastructure.
Studies have been carried out with both accidental and terrorist based explosions in mind
and how to protect these structures. FRP-wrapping and steel-jacketing have both been
564

found to be effective in blast resistance of RC columns. For a while, it was fairly common
to retrofit columns with steel jacketing for blast protection. In recent years, FRP composite
material has been utilized more frequently due to its high strength to weight ratio, ease of
application, corrosion resistance, and longer life span. In blast proofing, the safety and the
aesthetic integrity of the structure must remain intact and the retrofitting must not inhibit
any of the former structures functionality or accessibility (Malvar et al 2007; Bao and Li
2009; Buchan and Chen 2007; Muszynski and Purcell 2003; Crawford 2006; Tan and
Patoary 2009).
Malvar et al (2007) noted that in seismic regions, steel jacketing of columns for blast
resistance could make the member too stiff for seismic design. In this situation, FRP
wrapping was found very effective in increasing the shear strength of RC columns and
therefore preventing the diagonal shear failure that is often present in the case of blast.
Furthermore, it was found that the extreme heat from a blast did not seem to affect the FRP
retrofitting material, but secondary debris flying through the air could cause its failure. A
veneer covering the wrapping material for protection against such debris was
recommended.
Considering that exterior structural columns are very susceptible to any form of terrorist
attack, Bao, and Li (2009) carried out a finite element analysis of 12 different columns
undergoing various blast conditions using LS-DYNA. The finite element models were
validated with experimental studies. The study assumed a blast load ranging from an
equivalent of 0 to 1 ton of TNT at a stand off distance of 5 meters from an RC column. The
RC column models were subjected to axial and blast loads. The blast load considered
consisted of the impact of primary fragments and debris, impact of secondary debris,
overpressure, and reflected pressure. In order for the columns to survive such impulse
loadings and display a ductile behaviour, the shear capacity of the column must exceed the
flexural capacity. This means that the amount and distribution of the transverse
reinforcement within the plastic region plays a key role in the behaviour of columns under
blast loading. Their study revealed that the RC columns with lower amount of transverse
reinforcement would fail in shear followed by failure due to the axial load. The shear
failure in the columns could lead to the collapse of the entire structure. One way to enhance
the shear capacity and to provide confinement for the concrete core and lateral restraint
against buckling of longitudinal reinforcement in order to endure the blast impulse loads is
by FRP shear strengthening of columns.
Muszynski and Purcell (2003) along with The U.S. Army Waterways Experiment Station
conducted field blast experimentation on two different reinforced concrete structures in
Israel. Structure A in this experimentation included two walls retrofitted with FRP material
and two non-retrofitted columns. Half of the interior wall was retrofitted with biaxial
knitted E-glass fabric, while the other half with three layers of CFRP. Structure B included
two columns that were retrofitted with FRP material. One column was wrapped with biaxial
knitted E-glass fabric and the other with three layers of CFRP. 860 kg of TNT was utilized
in the full-scale detonations. Each structure was subjected to four equal blasts from various
stand-off positions around the building. In Structure A, both of the non-retrofitted RC
columns failed in tension after the third blast. The retrofitted columns in Structure B
remained intact through the third trial. In the fourth blast situation, Structure B experienced
565

complete demolition. The columns were completely sheared off at the connections at both
the top and bottom of the column. Column failure in Structure B in this experimentation
was blamed on poor casting and construction. It was found that both the CFRP and E-glass
retrofitting did indeed increase the columns ability to withstand blast impact for most of the
trials. Furthermore, the FRP-strengthening was highly successful in the wall retrofitting
case.
2.3 Fire
During a fire situation the structure must be able to support the required service load for
sufficient amount of time to allow occupants to escape safely. Structural members that are
retrofitted with composite material are no exception, and the load carrying integrity of the
retrofitted member must remain intact throughout a fire. One major drawback that has
prevented many engineers from utilizing FRP in structural strengthening is the lack of
sufficient data on the durability of composites during a fire. In recent years, several
researchers have looked at the effects of fire on FRP retrofitting materials (Chowdhury et al
2007; Bisby et al 2002; Bisby et al 2005; Kodur and Ahmed 2010; Kodur and Bisby 2005;
Kodur et al 2004; Williams et al 2004; Bisby et al 2005a; Bisby et al 2004).
Chowdhury et al (2007) performed fire endurance tests on full-scale FRP-wrapped RC
columns. Two axially loaded circular columns were exposed to fire. One specimen was
wrapped with two layers of FRP material with no thermal insulation. The second specimen
was wrapped with two layers of CFRP material and then sprayed with a cement based
insulating material. In the furnace where the testing was carried out, both columns were
subjected to a 2,635 kN axial compressive load during the fire exposure. The FRP-wrapped
column without the insulation failed under the axial load after 210 minutes in the furnace.
The FRP-wrapped column with the insulation survived 300 minutes in the furnace until the
axial load was gradually increased to 4,575 kN when failure occurred. It can be seen that
the column with no insulation for the FRP material experienced a loss in the strength due to
the thermal expansion in the member. In both cases, the FRP wraps surpassed the surface
temperature recommended for the composite material integrity (glass transition
temperature).
As noted by Bisby et al (2002), carbon fibers can maintain their strength through
temperatures in excess of 1000o C. However, the resin utilized in FRP reinforcing makes
the composite ineffective after the glass transition temperature which can range between
65o and 150o C. In the case of the insulated specimen, the CFRP wrap became structurally
ineffective and surpassed its glass transition temperature after 34 minutes. The addition of
the wrap coupled with insulation extended the life of the column under fire load and the
wrap remained intact during the entire fire test. For fire endurance, the use of insulating
material was recommended when designing FRP-retrofitted structures.
The most important factors to determining the fire endurance of a concrete member is the
thickness and thermal conductivity of the thermal insulation applied on top of the FRP
material (Bisby et al 2004). The failure of a retrofitted column during a fire exposure can
be categorized in three ways. The member can no longer handle its design load, the
temperature of the FRP material has surpassed its glass transition temperature, or the FRP
566

material ignites with the heat. An experiment was carried out to explore the effect of
thermal insulation thickness (Bisby et al 2005). One column specimen was wrapped with
one layer of CFRP wrap and 57 mm cement based insulation. A second column specimen
was confined with one layer of CFRP wrap and 32 mm cement based insulation. The
insulated columns were then covered with insulated epoxy paint. Both specimens
experienced very similar results when subjected to the fire endurance test. The specimen
with 57 mm of insulation experienced less axial deformation due to the thermal expansion
and longer survival time compared to the specimen with 32 mm of insulation. However, it
was noted that both columns experienced minimal axial deformation as compared to CFRPwrapped with no insulation due to the addition of the thermal protection. It was found
through numerical analysis that an unwrapped column of the same size had a fire endurance
period of 245 minutes. The column model was then wrapped with CFRP sheet and was
able to carry a higher service load, but had a fire endurance period of 100 minutes.
However, the CFRP-retrofitted column with insulation could endure over 6 hours of fire
exposure. It was concluded that that the use of FRP material enhances the load carrying
capacity, not the fire survivability. For fire endurance it is necessary to use thermal
insulation with the FRP material.
2.4 Earthquake
In seismic regions, FRP material has been emerging as a popular choice for structural
retrofitting and rehabilitation due to the ease of application, long life span, and its capability
to provide strength and ductility to structural members. Numerous studies have been
carried out on seismic retrofit and the durability of FRP material. However, few are
discussed in this paper (Gu et al 2010; Li and Kai 2010; Wu et al 2007).
Many older concrete structures are unfit to handle cyclic seismic loading and that structural
retrofitting is of dire need. Particularly column drift capacity and column hinging both
become prevalent issues. An experimental study was carried out on axially loaded FRPretrofitted concrete columns under cyclic lateral loading (Gu et al 2010). It was found that
the columns with a higher axial load required more FRP layers for confinement in order to
be effective in seismic regions. With the greater axial loads, an increase in FRP wrap
stiffness allowed for greater seismic energy dissipation. The amount of energy dissipation
for the columns was proportional to the FRP stiffness. In another study involving seismic
retrofit, the possibility of using fiber anchors to prevent the premature delamination and
debonding of FRP sheet from the concrete members was discussed (Li and Kai 2010).
2.5 Tsunami
Offshore infrastructures are subject to extreme loading conditions including tidal waves and
tsunami. Most structural members that are underwater are columns. FRP materials are
becoming increasingly popular as underwater column retrofitting material due to their long
lifespan and its durability under harsh aquatic conditions. However, there has been very
little research carried out on the effects of tsunami on FRP-retrofitted RC concrete columns
in offshore infrastructure. Most aging offshore columns are fully submerged in the water.
The bonding and curing process is probably the most challenging aspect when dealing with
underwater structural FRP-retrofitting (Wu et al 2007). It has been found that the
utilization of CFRP grids along with an underwater epoxy resin is the best retrofitting
567

solution for offshore infrastructure. Using typical FRP sheets has been deemed not
effective as it is difficult to form an adequate bond between the FRP sheet and concrete
because of the interference of the water. Another method of strengthening offshore
infrastructure is by encasing the column with an FRP tube as confinement. Much more
research is needed to be carried out on the topic of CFRP grid durability under tsunami
loadings.
2.6 Environmental Conditions
The FRP-retrofitted members are subject to outside exposure and are affected by the
environmental changes. These changes can consist of extreme temperature variations,
freeze-thaw situations, ultraviolet radiation exposure, and wet-dry cycles.
Numerous investigations have been performed on individual environmental variables
affecting the FRP systems with the exception of few studies which considered the influence
of combined environmental parameters (Bae and Belarbi 2008; Ray 2009; Robert et al
2010; El Hacha et al 2010; Karbhari 2009; Buyukozturk et al 2003; Nordin 2010). These
studies revealed that CFRP materials had much lower loss of strength due to extreme
temperature changes and wet-dry cycles than GFRP materials. In some cases, GFRPwrapped columns failed at stress levels as low as one third of the ultimate strength of the
material due to the thermal fatigue (Ray 2009). Additional study on the GFRP exposure to
high temperatures is necessary to evaluate the accuracy of the glass transition temperature
for such material (Robert et al 2010). Humidity might cause composite expansion and
deterioration of the polymer and fiber system which could lead to decreased material
durability (Ray 2009; Karbhari 2009).
Very few studies have considered combinations of environmental variables in their
experimentation (Bae and Belarbi 2008; El Hacha et al 2010). The CFRP and GFRP
materials behaved differently under extreme environmental changes. With freeze-thaw
cycles, it was found that both materials seemed to withhold a high percentage of their
material strength integrity. In moisture and saline conditions, the CFRP material strength
decreased by a much smaller percentage compared to the GFRP material strength that was
severely affected (Bae and Belarbi 2008). Under heating and cooling cycles, the CFRPwrapped column specimens exhibited an increase in strength of up to 74 % as compared to
unconfined specimens. It was concluded that both as-built and CFRP-wrapped cylinders
undergoing freeze-thaw conditions experienced a slight decrease in their compressive
strength (El-Hacha et al. 2010).
2.7 Fatigue
In the presence of fatigue load, the bond strength between composite materials and concrete
is often the limiting strength factor in the design. Various studies have been conducted on
the concrete-FRP sheet bond behavior under fatigue loading (Yun et al 2008; Harries 2005;
Nigro et al 2010; Omar Chaallal et al 2010; Kim and Heffernan 2008; Post et al 2008;
Cosenza et al 1997; Warn and Aref 2010; Choi et al 2010; Debaiky et al 2007; Harries and
Aidoo 2006; Aidoo et al 2004). In flexural strengthening of FRP-externally bonded
concrete specimens four failures may occur. They are (a) Debonding and peeling away
from the concrete member at the ends of the FRP strips, (b) peeling away from the concrete
568

member at the flexural crack locations, (c) peeling away from the concrete member at the
location of shear cracks, and (d) debonding due to unevenness of the concrete surface
(Aidoo et al. 2004).
Under cyclic fatigue loading of bonded FRP-strengthened beams although the steel bars
controlled the failure load, the FRP-reinforcement indeed prolonged the fatigue life of the
beams (Harries 2005; Kim and Heffernan 2008). In a state-of-the-art report several
parameters influencing the bond between FRP rebars and concrete were analyzed. These
parameters included fiber type, bar outer surface shape and type of matrix, bar diameter,
bar sand coating, confining pressure, and compressive concrete strength. It was observed
that the bond decreased as the FRP bar size increased. Additionally, the bond strength did
not depend on the concrete strength but on the strength of the FRP materials (Cosenza et al
1997).
The bond strength between FRP sheets for flexural strengthening applied to concrete prisms
subjected to cyclic loading was investigated. It was revealed that in the first loading cycle
the largest energy was dissipated compared to later cycles and numerous micro cracks
occurred. It was also observed that under fatigue loading, slip increased much more rapidly
during the initial few cycles of loading as compared to later cycles (Yun et al 2008).
Chaallal et al (2010) investigated the effect of fatigue load on shear strengthened RC beams
by CFRP sheets. The beam member strengthened with one layer of CFRP was able to
survive five million cycles. With the addition of an extra layer of CFRP, the member failed
in fatigue earlier than the beam specimen with a single CFRP layer due to the increased
CFRP rigidity.
3. CONCLUSIONS
Further studies are required to fully understand the durability of composite materials under
various load cases. The lack of full understanding prevents the wide use of FRP for
structural repair and reinforcement. The present review paper has offered an updated look
at research that has been conducted on the use of FRP for structural elements under extreme
environmental changes, vehicular impact, blast, fire, and fatigue loading conditions.
4. RESEARCH NEEDS
To increase the wide use of FRP materials in civil infrastructure applications the following
areas are identified for future research.
Full-scale tests on FRP-retrofitted columns subjected to dynamic impact load with varying
intensity will provide realistic accidental scenarios. Most existing columns are designed for
gravity loads or an equivalent static lateral load representing impact. Both eccentric and
concentric axial loads should also be considered.
When dealing with fire durability and FRP composite material, studies have found that the
retrofitting material and resin polymers are highly flammable and offer little to no fire
protection when used by themselves for structural strengthening. Additional experiments
should be carried out looking into combining the insulating and FRP retrofitting processes
569

in to one fire retardant retrofitting material. Adding on an insulating polymer to the current
resins and polymers used in FRP materials could create a composite material that is able to
withstand high temperatures and flames while keeping its strength. Developing such a
material would take experimentation with different resin and polymer combinations under
fire situations. Full-scale laboratory testing should be carried out with loaded structural
members wrapped with varying insulation-FRP material combinations undergoing fire
conditions. If the hybrid composite material would be able to withstand the fire, this would
eliminate a full step of applying extra insulating material on top of retrofitted structural
members. This would in turn save on time and construction costs while at the same time
preserving the safety of the members within the retrofitted structure.
Extremely limited studies have been performed on offshore retrofitting of concrete columns
taking into account possible tsunami situations. Scaled-down models of columns
submerged in water should be tested for different wave magnitudes. Accelerated testing
can be carried out on RC columns retrofitted with the suggested grid CFRP configuration
combined with underwater epoxy to project the effect of water and the roughness of wave
action over a prolonged period of time.
The extreme environmental changes that retrofitted structures may undergo consist of
extreme temperature variances, freeze-thaw situations, ultraviolet radiation exposure, and
wet-dry cycles. Combining two or more of these environmental changes at the same time
can in turn have a magnified effect on the durability of FRP retrofitting material. More
studies need to be carried out with varying the combinations of extreme environmental
changes. Study on combined effects of environmental variables simultaneously would
provide a realistic scenario of the unpredictable nature of the environment.
Additional full-scale laboratory experiments must be carried out on the effect of fatigue
loading on the durability of FRP retrofitting material. Since bonding tends to be the
limiting strength factor in FRP retrofitting design, special consideration must be placed on
the FRP-concrete bond under fatigue loading.
5. REFERENCES
1.
2.

3.
4.

5.

El-Tawil, S., Severino, E., and Fonseca, P. 2005. Vehicle Collision with Bridge Piers.
Journal of Bridge Engineering, ASCE, 10 (3): 345-353.
Ferrier, E., and Hamelin, P. 2005. Dynamic Behaviour of Externally bonded CFRP:
Application to reinforced concrete column externally reinforced under impact loading.
Proceedings of the International Symposium on Bond Behaviour of FRP in Structures,
pp. 509-518.
Tsang, H., and Lam, N.T.K. 2008. Collapse of Reinforced Concrete Column by
Vehicle Impact. Computer-Aided Civil and Infrastructure Engineering, 23: 427-436.
Thilakarathna, H.M.I., Thambiratnam, D.P., Dhandasekar, M., and Perera, N. 2010.
Numerical Simulation of Axially Loaded Concrete Columns under Transverse Impact
and Vulnerability Assessment. International Journal of Impact Engineering, pp. 11001112.
Malvar, L.J., Crawford, J.E., and Morrill, K.B. 2007. Use of Composites to Resist
Blast. Journal of Composites for Construction, 11(6): 601-610.
570

6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.

Bao, X., and Li, B. 2010. Residual Strength of Blast Damaged Reinforced Concrete
Columns. International journal of Impact Engineering, 37: 295-308.
Buchan, P.A., and Chen, J.F. 2007. Blast Resistance of FRP Composites and Polymer
Strengthened Concrete and Masonry Structures- A State of the Art Review. Science
Direct Composites: Part B, 38: 509-522.
Muszynski, L.C., and Purcell, M.R. 2003. Composite Reinforcement to Strengthen
Existing Concrete Structures against Air Blast. Journal of Composites for
Construction, ASCE, 7(2): 93-97.
Crawford, J.E. 2006. Protective Designs for Blast and Impact Threats. 4th World
Conference on Structural Control and Monitoring, Singapore, 8 p.
Tan, K.H., and Patoary, M.K.H. 2009. Blast Resistance of FRP-Strengthened Masonry
Walls. I: Approximate Analysis and Field Explosion Tests. Journal of Composites for
Construction, ASCE, 13(5): 422-430.
Chowdhury, E.U., Bisby, L.A. Green, M.F., and Kodur, V.K.R. 2007. Investigation of
Insulated FRP-Wrapped Reinforced Concrete Columns in Fire. ScienceDirect Fire
Safety Journal, 42: 452-460.
Bisby, L.A., Williams, B.K., Green, M.F., and Kodur, V.K.R. 2002. Studies on the
Fire Behavior of FRP Reinforced and/or Strengthened Concrete Members. NRC
Publications Archive, 12 p.
Bisby, L.A., Kodur, V.R., and Green, M.F. 2005. Fire Endurance of Fiber-Reinforced
Polymer-Confined Concrete Columns. ACI Structural Journal, 102(6): 883-891.
Bisby, L.A., Green, M.F., and Kodur, V.K.R. 2005. Modeling the Behavior of Fiber
reinforced Polymer-Confined Concrete Columns Exposed to Fire. Journal of
Composites for Construction, ASCE, 9(1): 15-24.
Kodur, V.K.R., and Ahmed, A. 2010. Numerical Model for Tracing the Response of
FRP-Strengthened RC Beams Exposed to Fire. Journal of Composites for
Construction, ASCE, 14(6), 730-742.
Kodur, V.K.R., and Bisby, L.A. 2005. Evaluation of Fire Endurance of Concrete Slabs
Reinforced with Fiber-Reinforced Polymer Bars. Journal of Structural Engineering,
131(1): 34-43.
Kodur, V.K.R., Green, M., Bisby, L.A., and Williams, B. 2004. Evaluating the Fire
Performance of FRP-Strengthened Concrete Structures. Concrete Engineering
International, 8(2): 48-50.
Williams, B.K., Kodur, V.K.R., Bisby, L.A., and Green, M.F. 2004. The Performance
of FRP-Strengthened Concrete Slabs in Fire. NRC Publications Archive, pp. 1-8.
Bisby, L.A., Green, M.F., and Kodur, V.K.R. 2005a. Response to Fire of Concrete
Structures that Incorporate FRP. Progress in Structural Engineering and Materials, 7
(3): 136-149.
Bisby, L.A., Kodur, V.K.R., and Green, M.F. 2004. Performance in Fire of FRPConfined Reinforced Concrete Columns. 4th International Conference on Advanced
Composite Materials in Bridges and Structures, Calgary, Alberta, pp. 1-8.
Gu, D., Wu, G., Wu, Z., and Wu, Y. 2010. Confinement Effectiveness of FRP in
Retrofitting Circular Concrete Columns under Simulated Seismic Load. Journal of
Composites for Construction, ASCE, 14(5): 531-540.
Li, B., and Kai, Q. 2010. Seismic Behavior of Reinforced Concrete Interior BeamWide Column Joints Repaired Using FRP. Journal of Composites for Construction,
ASCE, in print.
571

23. Wu, Z., Wang, X., and Iwashita, K. 2007. State-of-the-Art of Advanced FRP
Applications in Civil Infrastructure in Japan. American Composites Manufacturers
Association, 13 p.
24. Bae, S., and Belarbi, A. 2008. Effects of Various Environmental Conditions on RC
Columns Wrapped with FRP Sheets. Journal of Reinforced Plastics and Composites,
pp. 290-309.
25. Ray, B.C. 2009. Impact of Environmental and Experimental Parameters on FRP
Composites. 18th International Symposium on Processing and Fabrication of Advanced
Materials, Sendai, Japan, 11 p.
26. Robert, M., Wang, P., Cousin, P., and Benmokrane, B. 2010. Temperature as an
Accelerating Factor for Long-Term Durability Testing of FRPs: Should there be any
Limitations? Journal of Composites for Construction, ASCE, 14(4): 361-367.
27. El-Hacha, R., Green, M.F., and Wight, G.R. 2010. Effect of Severe Environmental
Exposures on CFRP Wrapped Concrete Columns. Journal of Composites for
Construction, ASCE, 14(1): 83-93.
28. Karbhari, V.M. 2009. Durability Data for FRP Rehabilitation Systems. Structural
Systems Research Project, 182 p.
29. Buyukozturk, O., Gunes, O., and Karaca, E. 2003. Progress on Understanding
Debonding Problems in Reinforced concrete and Steel Members Strengthened Using
FRP Composites. Construction and Building Materials, 18 (1): 9-19
30. Nordin, C., Ma, Z.J., and Penumadu, D. 2010. Combined Effect of Loading and Cold
Temperature on the Stiffness of Glass Fiber Composites. Journal of Composites for
Construction, ASCE, 14(2): 224-230.
31. Yun, Y., Wu, Y., and Tang, W.C. 2008. Performance of FRP Bonding Systems under
Fatigue Loading. Engineering Structures, 30(11): 3129-3140.
32. Harries, K. 2005. Fatigue Behaviour of Bonded FRP Used for Flexural Retrofit.
Proceedings of the International Symposium on Bond Behaviour of FRP in Structures
(BBFS 2005): 547-552.
33. Nigro, E., Ludovico, M.D., and Bilotta, A. 2010. Experimental Investigation on FRPConcrete Debonding Under Cyclic Actions. Journal of Materials in Civil Engineering.
34. Chaallal, O., Boussaha, F., and Bousselham, A. 2010. Fatigue Performance of RC
Beams Strengthened in Shear with CFRP Fabrics. Journal of Composites for
Construction, 14(4): 415-423.
35. Kim, Y.J., and Heffeman, P.J. 2008. Fatigue Behavior of Externally Strengthened
Concrete Beams with Fiber-Reinforced Polymers: State of the Art. Journal of
Composites for Construction, ASCE, 12(3): 246-256.
36. Post, N.L., Case, S.W., and Lesko, J.J. 2008. Modeling the Variable Amplitude
Fatigue of Composite Materials: A Review and Evaluation of the State of the Art for
Spectrum Loading. International Journal of Fatigue, 30(12): 2064-2086.
37. Cosenza, E., Manfredi, G., and Realfonzo, R. 1997. Behavior and Modeling of Bond
of FRP Rebars to Concrete. Journal of Composites for Construction, ASCE, 1(2): 4051.
38. Wam, G.P., and Aref, A.J. 2010. Sustained-Load and Fatigue Performance of a Hybrid
FRP-Concrete Bridge Deck System. Journal of Composites for Construction, ASCE,
14(6): 856-864.

572

39. Choi, K., Reda, Taha, M.M.R., Masia, M.J., Shrive, P.L., and Shrive, N.G. 2010.
Numerical Investigation of Creep Effects on FRP-Strengthened RC Beams. Journal of
Composites for Construction, ASCE, 14(6): 812-822.
40. Debaiky, A.S., Green, M.F., and Hope, B.B. 2007. Modeling of Corroded FRP
Wrapped Reinforced Concrete Columns in Axial Compression. Journal of Composites
for Construction, ASCE, 11(6): 556-564.
41. Harries, K.A., and Aidoo, J. 2006. Debonding and Fatigue Related Strain Limits for
Externally Bonded FRP. Journal of Composites for Construction, ASCE, 10(1): 87-90.
42. Aidoo, J., Harries, K.A., and Petrou, M.F. 2004. Fatigue Behavior of Carbon Fiber
Reinforced Polymer-Strengthened Reinforced Concrete Bridge Girders. Journal of
Composites for Construction, ASCE, 8(6): 501-509.

573

574

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

FATIGUE PERFORMANCE OF HYBRID SFRP-GFRP-UHPC BEAMS


D. Chen1 and R. El-Hacha2
1
2

PhD Candidate, Department of Civil Engineering, University of Calgary, Alberta, Canada


Associate Professor, Department of Civil Engineering, University of Calgary, Alberta, Canada

ABSTRACT
This paper describes the experimental program on innovative hybrid SFRP-GFRP-UHPC
beams under fatigue loading. Three beams were tested, one as the control beam tested
under static loading with the remaining two subjected to variable amplitude cyclic loading.
Debonding occurred at the GFRP-UHPC interface, causing permanent deformations to
occur under fatigue loading, though the amount of damage was minimized through the use
of specially designed clamping/anchorage systems at the beam ends. Investigation into the
flexural stiffness of the beams showed insignificant changes over time until the period
immediately prior to ultimate failure; however, the beams tested under fatigue loading did
not achieve the same flexural stiffness as that of the beam tested under static loading. It is
recommended that further research be performed, using an improved interface connection
mechanism.
1. INTRODUCTION
Research into the behaviour under fatigue loading of high performance materials, such as
Fibre Reinforced Polymers (FRPs) and Ultra-High-Performance-Concrete (UHPC), for
structural purposes initiated mainly in the late 1970s (Reifsnider and Jamison, 1982; DewHuges and Way, 1973; Schutz and Gerharz, 1977). In more recent years, focus has been
shifted towards hybrid structural members that are composed of a combination of high
performance materials as well as conventional materials, such as reinforced concrete and
structural steel (Kitane et al., 2004; Cheng and Karbhari, 2006; Moon et al., 2009). The
advantage of high performance materials is related to their smaller weight-to-strength ratio
as well as their resistance to corrosion and other environmental effects. This paper will
explore the experimental testing under both static and fatigue flexural loading of a new
hybrid member composed of Glass FRP (GFRP), Steel FRP (SFRP) as well as UHPC.
2. DESIGN OF HYBRID SFRP-GFRP-UHPC BEAMS
The hybrid beams were designed to mimic the interaction of forces in conventional
reinforced concrete members, where the reinforced concrete acts to resist compression
forces with the internal steel reinforcement bars providing tensile resistance. The distance
575

between the two forces creates the moment arm required for flexural resistance. In this
design, an optimized cross-section was used. The cross-section consists of a pultruded
GFRP hollow box section beam, strengthened with a UHPC cast-in-place layer on top and a
sheet of SFRP bonded to the bottom of the beam. The compression and tension forces are
held by the UHPC and the SFRP, respectively, while the GFRP beam provides the
necessary moment arm as well as shear resistance. In order to provide shear resistance at
the GFRP-UHPC interface, a combination of both continuous bonding, from epoxy
adhesive applied at the contact surface, as well as discreet mechanical connections, from
GFRP shear studs, was used. The cross-section of the hybrid beam is shown in Figure 1.

(dimension are in mm)

Fig. 1. Cross-section of hybrid SFRP-GFRP-UHPC beam


3. MATERIALS
The pultruded GFRP hollow box section beam was fabricated with glass rovings in the
longitudinal direction and continuous strand glass mats or stitched reinforcements in the
transverse direction for strength. From the manufacturers specifications, the modulus of
elasticity is equal to 17.2 GPa, with flexural and shear resistances of 207 MPa and 31 MPa,
respectively (Strongwell, 2009a). Tests performed on six tension coupon test specimens in
accordance with ASTM D3039 (ASTM, 2000) showed average values for tensile strength
in the bottom flange and side web equal to 321 40 MPa and 230 44 MPa, respectively.
Similarly, average values for the modulus of elasticity were 27558 3379 MPa and 18689
2927 MPa for the bottom flange and side web, respectively.
The UHPC, a proprietary product by Lafarge marketed as Ductal BS1000, had an ultimate
compressive strength at 28 days of between 150 and 180 MPa with an expected
compressive strength equal to 30 MPa after 24 hours, as provided by the manufacturers
specifications (Lafarge, 2007). If subjected to additional heat treatment after curing, a
procedure that was not administered during this research program, the compressive strength
can be expected to reach above 200 MPa. Compression tests performed in accordance with
ASTM C469 (ASTM, 1994) provided average compressive strength of 124 18 MPa with
an average modulus of elasticity equal to 45107 3575 MPa.
The SFRP laminate was a SFRP sheet (Hardwire 3x2-20 Tape) impregnated with
Sikadur 330 epoxy adhesive. The SFRP sheet was formed from uni-directional brasscoated ultra-high strength twisted steel wires, made into cords by twisting two wires at a
high angle around three other straight wires; each wire was 0.35 mm in diameter. Each
576

sheet had twenty steel cords laid side-by-side in every one inch width (7.67 cords/cm)
where a silicon mesh was applied on one face to maintain the correct spacing (Hardwire,
2010). Each steel cord had a breaking load of 1539 N, a modulus of elasticity of 160 GPa
and an ultimate tensile strain of 2.1% (Hardwire, 2010). The final SFRP laminate had a net
cross sectional area of 0.38 mm2/mm and a thickness of 1.23 mm with a tensile strength of
985 MPa and a modulus of elasticity of 66.1 GPa (Hardwire, 2010). Experimental tests
performed with six tension coupon specimen in accordance with ASTM D3039 (ASTM,
2000) showed that the average strength and modulus of elasticity were 963 99 MPa and
67373 4968 MPa, respectively. From manufacturers specifications, the epoxy adhesive
had a tensile strength of 30 MPa, ultimate tensile strain of 1.5% and a modulus of elasticity
equal to 3.8 GPa (Sika, 2007).
At the GFRP-UHPC interface, the epoxy adhesive used was Sikadur 32 HI-MOD.
Manufacturers specifications provided values for the tensile strength equal to 48 MPa,
with a tensile elongation limit of 1.9% and a modulus of elasticity equal to 3726 MPa (Sika,
2008). The GFRP shear studs had minimum resistance under single shear equal to 99 MPa,
according to manufacturers specifications; values for the maximum tensile strength and
modulus of elasticity were given as 66 MPa and 13.8 GPa, respectively (Strongwell,
2009b). Experimental testing performed during a previous phase in this research program
showed the average resistance under single shear to be equal to 141 MPa (Elmahdy, 2010).
4. EXPERIMENTAL PROGRAM
Three hybrid beams were tested, with one tested under static loading as the control beam
(Beam S-S) and the remaining two tested under variable amplitude fatigue loading (Beams
S-F1 and S-F2). All three beams were tested under four-point flexural loading, with the
test load set-up and the locations of the instrumentation shown in Figure 2.

Fig. 2. Test load set-up and location of instrumentation

Due to the lack of available experimental work in the literature performed to investigate the
behaviour of this particular hybrid system under fatigue loading, the fatigue stress limits
used during experimentation were determined based on existing literature on the individual
material components in the cross-section. According to CAN/CSA S6-06, Cl. 16.8.3 (CSA,
2006), the sustained stress limits for GFRP is 25% of the ultimate strength, fu(GFRP). Due to
the fact that SFRP is a relatively new type of FRP material, there is no current limit set in
codes and standards; however, the recommended fatigue stress limit for steel bars, given as
577

80% of its yield stress, was adopted for the purpose of this research program. Since the
ultra-high strength wires do not demonstrate significant yielding, the sustained stress limit
was taken as 80% of the ultimate strength, fu(SFRP). It was expected that the true fatigue
limit of the hybrid beam would be above 0.25fu(GFRP) but below 0.80fu(SFRP) as hybrid
systems obtain their properties from their individual material components. The fatigue
loading test program is shown in Table 1.
Table 1. Fatigue Loading Testing Program
Beam
ID

S-F1

S-F2

Range
Load (kN)

A
73 122

B
73 146

C
98 146

Frequency (Hz)

1.5

1.0

1.5

Load Condition (Min Max)


D
E
F
122 146 98 171 122 171
1.5

G
122 195

H
122 219

1.0

1.5

1.0

1.0

Number of cycles

500,000

250,000

500,000

250,000

250,000

250,000

250,000

250,000

GFRP Stress (%fu(GFRP)

15 25

15 30

20 30

25 30

20 35

25 35

25 40

25 45

SFRP Stress (%f u(SFRP)

12 20

12 25

16 25

20 25

16 29

20 29

20 33

20 - 37

Load (kN)

73 98

73 122

73 146

73 171

73 195

73 219

Frequency (Hz)

2.0

1.5

1.0

1.0

0.5

0.5

Number of cycles

500,000

500,000

250,000

250,000

250,000

250,000

GFRP Stress (%f u(GFRP)

15 20

15 25

15 30

15 35

15 40

15 45

SFRP Stress (%f u(SFRP)

12 16

12 20

12 25

12 29

12 33

12 37

At the start of testing, prior to fatigue loading, as well as at the transition between load
conditions, the hybrid beams were subjected to three quasi-static cycles between the
maximum and minimum loads of the upcoming loading condition. Loading frequency
changes during testing were due to machinery limitations, where larger actuator strokes
required for higher stress amplitude load conditions could only be achieved at a lower
loading frequency.
5. EXPERIMENTAL RESULTS
5.1 Control Beam: Beam S-S
Beam S-S was loaded under deflection control, at a rate of 1 mm/min. In the initial phase
of testing, the beam behaved linear-elastically. Premature debonding at the GFRP-UHPC
interface occurred at an applied load of 124 kN and again at 133 kN upon further loading.
A relative peak load of 174 kN was reached after which significant ductile behaviour was
exhibited, with increasingly larger deflections experienced at constant load intervals.
Ultimate failure occurred at a load of 187 kN. At failure, the strain in the SFRP and UHPC
were 6880 and 780 , respectively, along with a mid-span deflection of 54 mm. The
load-deflection and load-strain curves at mid-span are shown in Figure 3 with the strain
distributions along the bottom of the beam and the strain profiles across the depth at midspan provided in Figure 4. From the data, it is seen that after interface debonding occurred,
two neutral axes developed in the cross-section of Beam S-S, demonstrating the loss of
composite action between the GFRP hollow box section beam and the UHPC layer, where
vertical lifting off of the UHPC layer above the GFRP occurred.

578

200

200

180

180

Debonding #2

160
Debonding #1

140

SG-3

120
100
80

SG-4

140

SG-2

120

Load (kN)

Load (kN)

SG-1

160

100
80

60

60

40

40

20

20

0
0

10

20

30

40

50

60

-2000

Mid-span Deflection (mm)

2000

4000

Mid-span Strain ()

Compression

6000

8000

Tension

Fig. 3. Load-deflection and load-strain curves at mid-span for Beam S-S


In order to prevent lifting off of the UHPC layer in the hybrid beams to be tested under
fatigue loading, a clamping/anchorage system was developed and installed at each end of
the beams. A diagram of the clamping/anchorage system is shown in Figure 5.
8000

300

7000

at 124 kN
at 175 kN
Ultimate (187.4 kN)

6000

Strain ()

Height from Bottom of Section (mm)

Mid-span

5000
4000
3000
2000
1000
0
0

500

1000

1500

2000

2500

3000

Distance from Support (mm)

Top of concrete

250

Top flange of GFRP box

200

150
GFRP box mid-depth
at 124 kN
at 175 kN
Ultimate (187.4 kN)

100

50
Bottom flange of GFRP box
0
-7000

-5000

-3000

Compression

-1000

1000

3000

Mid-span Strain ()

5000

7000

Tension

Fig. 4. Strain distributions along bottom of beam and strain profiles across the depth at midspan for Beam S-S

Fig. 5. Diagram of clamping/anchorage system used in Beams S-F1 and S-F2


5.2 Beam S-F1
Fatigue testing of Beam S-F1 proceeded without significant changes in deflection and strain
until 2,199,664 cycles during load condition G. Ultimate failure occurred at 2,204,023

579

200

200

180

180

160

160

140

140

120

120

Load (kN)

Load (kN)

cycles, where maximum strains in the SFRP and the UHPC were 7730 and 730 ,
respectively. The maximum mid-span deflection reached was 46 mm. The load-deflection
and load-strain curves at mid-span are given in Figure 6. The strain distributions along the
bottom of the beam as well as the strain profiles across the depth of the beam at mid-span at
intervals are provided in Figure 7.

100
80
60

100
80
60

40

40

initial quasi-static cycles


quasi-static after 2000000 cycles
at 2204023 cycles

20
0
0

10

20

30

40

50

initial quasi-static cycles


quasi-static after 2000000 cycles
at 2204023 cycles

20
0
-2000
0
Compression

60

Mid-span Deflection (mm)

2000

4000

Mid-span Strain ()

6000
Tension

8000

Fig. 6. Load-deflection (left) and load-strain curves at mid-span for Beam S-F1
6000

300

Depth from Bottom of Section (mm)

Mid-span

5000

Strain ()

4000

3000

2000
S-S
S-F1 initial quasi-static cycles
S-F1 quasi-static after 2000000 cycles
S-F1 at 2199664 cycles
S-F1 at 2204023 cycles

1000

0
0

500

1000

1500

2000

2500

3000

Distance from Support (mm)

Top of concrete

250

Top flange of GFRP

200

150
GFRP mid-depth
100

50

0
-6000

S-S
S-F1 initial quasi-static cycles
S-F1 quasi-static after 2000000 cycles
S-F1 at 2199664 cycles
S-F1 at 2204023 cycles
-4000
Compression

-2000

Bottom
flange of
GFRP

2000

4000

6000

Mid-span Strain () Tension

Fig. 7. Strain distributions along the bottom of the beam and strain profiles across the depth
of the beam at mid-span for Beam S-F1
It can be seen in Figure 6 and Figure 7 that the debonding occurred after 2,000,000 cycles,
with the formation of two distinct neutral axes as well as permanent deformation in the
mid-span deflection and strains.
5.3 Beam S-F2
After the three initial quasi-static cycles, an error in the machine programming resulted in
an accidental load of 131 kN, with an associated resulting mid-span deflection of 17.9 mm
applied to the beam. The setting was corrected and testing resumed. Beam S-F2 was
subjected to 1,250,000 cycles prior to failure, which occurred during the quasi-static cycles
for the subsequent loading cycle, load condition D. The ultimate load reached was 156 kN,
with strains in the SFRP and UHPC equal to 6340 and 760 and a mid-span deflection
of 50 mm. The load-deflection and load-strain curves at mid-span during fatigue testing for
Beam S-F2 are provided in Figure 8. The strain distributions along the bottom of the beam
580

200

200

180

180

160

160

140

140

120

120

Load (kN)

Load (kN)

as well as the strain profiles across the depth of the beam at mid-span are shown in Figure
9. For Beam S-F2, debonding occurred between 500,000 and 750,000 cycles. The
behaviour of the beam after debonding was similar to Beam S-F1, where two distinct
neutral axes were present in the cross-section and permanent deformation were apparent, as
can be seen by the horizontal offsets in Figure 8.

100
80
60

100
80
60

initial quasi-static cycles


quasi-static after 500000 cycles
quasi-static after 1000000 cycles
at 1250000 cycles

40
20
0
0

10

20

30

40

50

initial quasi-static cycles


quasi-static after 500000 cycles
quasi-static after 1000000 cycles
quasi-static after 1250000 cycles

40
20
0

60

-2000

Mid-span Deflection (mm)

Compression

2000

4000

6000
Tension

Mid-span Strain ()

8000

Fig. 8. Load-deflection (left) and load-strain curves at mid-span for Beam S-F2
300

6000

Depth from Bottom of Section (mm)

Mid-span
5000

Strain ()

4000

3000

2000
S-S
S-F2 quasi-static after 500000 cycles
S-F2 quasi-static after 1000000 cycles
S-F2 quasi-static after 1250000 cycles

1000

0
0

500

1000

1500

2000

2500

3000

Distance from Support (mm)

Top of concrete

250

Top flange of GFRP

200

150
GFRP mid-depth
100

50

0
-6000

Bottom
flange of
GFRP

S-S
S-F2 quasi-static after 500000 cycles
S-F2 quasi-static after 1000000 cycles
S-F2 quasi-static after 1250000 cycles
-4000
Compression

-2000

2000

Mid-span Strain ()

4000

6000

Tension

Fig. 9. Strain distributions along the bottom of the beam and strain profiles across the depth
of the beam at mid-span of Beam S-F2
6. DISCUSSION AND ANALYSIS
To analysis the trend in the flexural stiffnesses of the beams during fatigue testing, the
secant flexural stiffness was calculated from the mid-span deflection data using Equation 1:
Pa
EI
(3 L2 4 a 2 )
(1)
24
where: EI = the secant flexural stiffness (Nmm2);
P = the applied load (N);
a = the shear span, equal to the distance between the support to the point load (mm);
= the mid-span deflection (mm); and,
L = the span length (mm).

581

The changes, as a function of the number of fatigue cycles, in mid-span deflection as well
as flexural stiffness, compared with the average flexural stiffness of Beam S-S, are shown
in Figure 10.
4.5E+12
S-S (static)

S-F1 Max

50

Flexural Stiffness (Nmm2)

Mid-span Deflection (mm)

60

S-F2 Max

40

S-F2 Min

30
20

S-F1 Min

10

4E+12

3.5E+12

3E+12

S-F2

S-F1

2.5E+12

2E+12

0
0

500,000

1,000,000

1,500,000

2,000,000

2,500,000

500,000

1,000,000 1,500,000 2,000,000 2,500,000

Number of Fatigue Cycles

Number of Fatigue Cycles

Fig. 10. Mid-span deflection and flexural stiffness over time for the hybrid beams
The results shown in Figure 10 demonstrate that significant loss in flexural stiffness took
place primarily when ultimate failure occurred. Overall, the trend over time for flexural
stiffness follows two phases, where the first phase consists of a relatively stable and
constant value followed by the second phase, just immediately prior to failure, which is
characterized by a substantial drop in flexural stiffness. As compared with the flexural
stiffness of Beam S-S, it is apparent that neither of the beams tested under fatigue loading
achieved the same stiffness as Beam S-S at any time during testing.
The clamping/anchorage system performed well to prevent the lifting off of the UHPC,
thereby postponing the structural effects of debonding during fatigue testing. Nevertheless,
the influence of debonding must be taken into account and it can be assumed that the
fatigue life of the hybrid beams would have been greater if interface debonding had not
occurred. Further testing under fatigue loading is recommended to obtain a better
understanding of the performance of these hybrid beams; however, an improved connection
mechanism at the interface is required for more exact results.
7. CONCLUSION
The following conclusions were made:
- The hybrid SFRP-GFRP-UHPC behaved in a linear-elastic manner during the majority
of both static and fatigue testing;
- The hybrid SFRP-GFRP-UHPC beams performed well under fatigue loading, with
insignificant losses in flexural stiffness during fatigue testing until ultimate failure;
- The clamping/anchorage system prevented damage caused by lifting off of the UHPC
layer; and,
- Further investigation is required, with the use of a more effective connection
mechanism at the GFRP-UHPC interface.

582

8. ACKNOWLEDGEMENTS
The authors would like to thank Fyfe Co. LLC, Lafarge Canada, Hardwire LLC and Sika
Canada Inc. for their generous donations of material. We would also like acknowledge the
financial support provided by the University of Calgary and the Natural Science and
Engineering Research Council of Canada.
9. REFERENCES
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.

ASTM. 1994. Standard test methods for static modulus of elasticity and Poissons ratio
of concrete in compression, ASTM C459-94. ASTM International, West
Conshohocken, PA. USA.
ASTM. 2000. Standard test method for tensile properties of polymer matrix composite
materials, ASTM D3039-00. ASTM International, West Conshohocken, PA, USA.
Canadian Standards Association (CSA). 2006. Canadian Highway Bridge Design
Code. CAN/CSA S6-06, Mississauga, Ontario.
Cheng, L. and Karbhari, V.M. 2006. Fatigue behavior of a steel-free FRP-concrete
modular bridge deck system. Journal of Bridge Engineering, 11(4): 474-488.
Dew-Hughes, D., Way, J.L. 1973. Fatigue of fibre-reinforced plastics: a review.
Composites, 4(4): 167-173.
Elmahdy, A. 2010. Experimental and analytical study of new hybrid beams constructed
from high performance materials. PhD, University of Calgary, 242 p.
Hardwire. 2010. 32 cord properties metric. Hardwire LLC, Pocomoke City, MD,
USA.
Kitane, Y., Aref, A., and Lee, G.C. 2004. Static and fatigue testing of hybrid fiberreinforced polymer-concrete bridge superstructure. Journal of Composites for
Construction, ASCE, 8(2): 182-190.
Lafarge. 2007. Ductal - Fiche de caractristiques techniques. Lafarge Canada Inc.,
Montreal, Quebec, Canada.
Moon, D.Y., Zi, G., Lee, D.H., Kim, B.M. and Huang, Y.K. 2009. Fatigue behaviour
of the foam-filled GFRP bridge deck. Elsevier Composites Part B: Engineering, 40(2):
141-148.
Reifsnider, K.L., and Jamison, R. 1982. Fracture of fatigue-loaded composite
laminates. International Journal of Fatigue, 4(4): 187-197.
Schtz, D., Gerharz, J.J. 1977. Fatigue strength of a fibre-reinforced material.
Composites, 8(4):245-250.
Sika. 2007. Sikadur 330 product data sheet. Sika Canada Inc., Pointe Claire, Quebec,
Canada.
Sika 228. Sikadur 32 HI-MOD product data sheet. Sika Canada Inc., Pointe Claire,
Quebec, Canada.
Strongwell. 2009a. EXTRENE properties. Strongwell Corporation, Bristol, VA, USA.
Strongwell. 2009a. FIBREBOLT properties. Strongwell Corporation, Bristol, VA,
USA.

583

584

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

SDOF ANALYSIS OF PROTECTIVE HARDENING DESIGN FOR


REINFORCED CONCRETE COLUMNS USING FIBRE
REINFORCED POLYMER WRAP
J. Quek and M.C. Ow
FYFE Asia Pte Ltd, Singapore, Singapore

ABSTRACT
The wrapping of columns with Fiber Reinforced Polymer (FRP) sheets has become a
common technique for the protective hardening of reinforced concrete columns against
blast pressure effects. FRP wrap provides added strength and ductility, resulting in an
economical solution to harden existing columns subject to blast threats.
This paper investigates the effectiveness of FRP wrapping on typical reinforced concrete
columns subject to various blast pressures and with different wrapping configuration.
Single-degree-of-freedom (SDOF) analyses of the unwrapped and wrapped columns are
carried out with FWrapSDOF, a proprietary SDOF analysis code for simulating the
response of FRP wrapped columns to blast loading.
The SDOF results are compared with the actual test results of unwrapped and wrapped
columns subject to various blast pressures and with different wrapping configuration. This
is to evaluate the fidelity of SDOF analyses for FRP wrapped reinforced concrete columns.
1. INTRODUCTION
The aim of this paper is to describe a retrofit design procedure that incorporates
documented enhancement effects, from an FRP retrofit, into a well-established software,
which is based on the SDOF methodology. This design procedure is to aid in the retrofitting
of existing RC columns, where improvement to its lateral resistance is required.
The retrofit design procedure and comparisons described herein is qualitative in nature.
Quantitative measurements and comparisons, for the evaluation of the accuracy and
effectiveness of the design procedure, are to be further assessed and verified with additional
testing in the future.

585

2. BASIS OF FWRAPSDOF SOFTWARE


2.1 Single-Degree-Of-Freedom (SDOF) Analysis
Several different dynamic analysis methods are available for blast resistant design. These
approaches can range from simple hand calculations and graphical solutions, to more
complex computer based applications. The decision on which approach to adopt is usually
based on the consideration between obtaining results of sufficient accuracy and the
complexity of the calculations involved.
The basic analytical model used in most blast design applications is the SDOF analysis
approach. Although all structures possess more than one degree of freedom, many
structures can be adequately represented as a series of SDOF systems for analytical
purposes. Furthermore, sufficiently accurate results can be obtained for primary load
carrying components such as columns.
Various publications prescribe well-documented methods to compute the equivalent SDOF
system properties for various types of structural components.
2.2 SDOF Blast Effects Design Spreadsheets (SBEDS)
2.2.1 Background
The SDOF Blast Effects Design Spreadsheets (SBEDS) is a product of the U.S. Army
Corps of Engineers Protective Design Center and was developed by Baker Engineering and
Risk Consultants, Inc. (BakerRisk). It is an Excel based tool for the design of structural
components subjected to dynamic loads, such as an airblast, using the SDOF methodology.
Contained within SBEDS are worksheets, for common structural components, which allow
a user to input available parameters, related to material properties and structural geometries,
in order to calculate the SDOF properties.

Fig. 1. General Resistance-Deflection diagram with softening


The foundation of SBEDS is an SDOF numerical integration scheme capable of analysing a
resistance function with multi-linear segments for initial response, as well as for rebound
(Fig. 1). The calculations are performed using a constant velocity numerical integration
scheme, as generally recommended, to solve the SDOF equations at each time step.
586

2.2.2 Reinforced Concrete Beam-Columns with Combined Compression and Lateral Load
For an SDOF analysis of a reinforced concrete column that is required to carry vertical
loads, in combination to being subjected to lateral blast loads, the ultimate flexural
resistance of the column can be calculated based on the interaction diagram (Fig. 2)
between axial load and moment capacity in a short column, which is very similar to the
interaction diagram used for static design.

Fig. 2. Reinforced concrete Beam-Column axial load moment capacity interaction diagram
In SBEDS, the ultimate moment capacity of a reinforced concrete beam-column is
calculated by developing the interaction diagram (Fig. 2) based on equations defined in
Chapter 10 of UFC 3-340-01, except that a straight line is conservatively assumed between
Mo, the dynamic moment capacity without axial load, and the balanced condition (Pb,Mb).
2.3 FWrapSDOF
2.3.1 Retrofit Design Procedure
Using the SBEDS and its methodology as the cornerstone, the FWrapSDOF was developed
to retain the user-friendliness of the beam-column spreadsheet in SBEDS, and at the same
time incorporate the retrofit design procedure. The modifications made were heavily
influenced by ACI 440.2R-08, while other FRP guides were referenced for consistency.
The modifications incorporated can be grouped into three main areas of interest: concrete
confinement due to FRP hoop wraps, flexural enhancement due to FRP flexural wrap and
shear contribution from steel links as well as FRP hoop wraps. Effects associated with FRP
bonding to concrete, aging or degradation, which reduces the strength of the FRP
contributing to the ultimate capacities of the column, are addressed by the factors
recommended in ACI 440.2R-08 (Fig. 3 and Fig. 4).

587

Fig. 3. Values of FRP properties and factors-of-safety based on input for flexural layers

Fig. 4. Values of FRP properties and factors-of-safety based on input for hoop layers
With the modifications made to these three areas, the resistance-deflection curve, which
originally only plotted the flexural response of the column, has been modified to reflect the
column shear capacities, as well as the expected deflection from the lateral loads (Fig. 5).

Fig. 5. Resistance-Deflection curve for reinforced concrete component


2.3.3 Confined Concrete Model
The confinement of reinforced concrete columns, by means of FRP jackets, can be used to
enhance their strength and ductility. For the development of the FWrapSDOF for column
retrofit, the stress-strain model developed by Lam and Teng (2003a,b) has been adopted
(Fig. 6). This model characterises the main aspects of the behaviour of FRP-confined
588

concrete in a simple form and its usefulness and accuracy can be seen from its adoption by
the Technical Report No. 55, ACI 440.2R-08 and ICC AC125 documents

Fig. 6. Lam and Tengs Stress-Strain model for FRP-confined concrete


The equations for the FRP-confined concrete model are extracted from ACI 440.2R-08 and
computation of the various components is as illustrated in Fig. 7. As these equations are
computed based on concrete cylinder strength, an option was built into the program to
consider input based on concrete cube strength. An additional option provided allows the
user to decide whether to consider confined concrete enhancement or not (Fig. 8).

Fig. 7. Computations for FRP-confined concrete model


2.3.4 Flexural Capacity
For the consideration of flexural enhancement with FRP layers, the user is required to
select the FRP System and provide the number of layers to apply (Fig. 8). Indication of the
589

exposure condition is for the consideration of the environmental reduction factor (Fig. 3
and Fig. 4).
For simplification of the design, the FRP for flexural wrap is considered to strengthen the
positive moment region only. Enhancement to negative moment region is assumed to be
physically impractical.
FRP on compression face is conservatively taken to not increase moment capacity, moment
of inertia or axial load capacity. The fibre on faces adjacent to the tension face, which may
partially experience tension, is also not taken into consideration. The average sectional
stiffness for the reinforced concrete column is taken as the average of the cracked and gross
sections, as proposed by TM5-1300.

Fig. 8. User-input options for retrofit using TYFO Fibrwrap Systems


2.3.5 Diagonal Shear Capacity
In FWrapSDOF, the user is able to consider the existing steel shear links (Fig. 9) in the
design. The inputs will go towards the computation of the shear capacity provided by the
steel shear links and is computed based on ACI 318-05 (Fig. 10). For the shear strength
reduction factor, , it is provided as an option for the user to decide (Fig. 8).

Fig. 9. User-Input options for provision of steel shear links


For the FRPs contribution to shear strength, all computations are based on the assumption
that the columns are completely wrapped. When hoop wraps are considered, the concrete
shear capacity is based on the full thickness of the cross section, implying that the distance

590

d is based on the fibre reinforcement at the column face rather than at conventional
reinforcing steel with a cover depth.

Fig. 10. Computations for contribution to shear capacity from steel links and hoop layers
2.4 Live Test
A retrofit using TYFO SCH-41-2X carbon horizontal wraps was designed, applied, and
validated with a blast load in a full-scale field test (Fig. 11). The objective of the test was to
explore the effectiveness of the FRP system in protecting the columns against close-in
threats. This test has demonstrated that retrofitting RC columns with the TYFO Fibrwrap
Systems is an efficient means to ensure the survivability of columns subjected to close-in
blast loads.
When the same test parameters were applied into FWrapSDOF, the results were in
consistent agreement that with the provision of the carbon hoop wraps, the columns are
sufficiently protected against potential diagonal shear failure.

Fig. 11. Typical test bed set-up for the full-scale field test
3. CONCLUSIONS
The design methodology adopted for the modification of an existing design program was
described. The modification applied is for the consideration of retrofitting existing RC
columns with composite wrap to improve their resistance to blast loads.

591

Both the flexural and diagonal shear capacities of conventional reinforced concrete
columns can be upgraded to prevent excessive lateral deflections and shear failure. A
qualitative comparison, between observations made in a live test with the runs made on the
modified program, shows congruence.
4. REFERENCES
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.

12.
13.
14.
15.

ICC-ES. 2010. Acceptance Criteria For Concrete And Reinforced And Unreinforced
Masonry Strengthening Using Externally Bonded Fiber-Reinforced Polymer (FRP)
Composite Systems. AC125, March
ACI Committee 440. 2008. Guide for The Design And Construction Of Externally
Bonded FRP Systems For Strengthening Concrete Structures. ACI 440.2R-08,
American Concrete Institute (ACI), Farmington Hill, MI, 76 p.
Concrete Society Committee. 2004. Design Guidance For Strengthening Concrete
Structures Using Composite Materials, Technical Report No. 55, Second Edition.
fib CEB.FIP. 2001. Externally Bonded FRP Reinforcement For RC Structures.
Technical Report Bulletin 14, July.
U.S. Army Corps of Engineers. 2005. Single-Degree-of-Freedom Blast Effects Design
Spreadsheets (SBEDS). PDC TR-05-01, March.
U.S. Army Corps of Engineers. 2006. Methodology Manual for the Single-Degree-ofFreedom Blast Effects Design Spreadsheets (SBEDS). PDC-TR 06-01, September.
U.S. Army Corps of Engineers. 2006. Users Guide for the Single-Degree-of-Freedom
Blast Effects Design Spreadsheets (SBEDS). PDC-TR 06-02, September.
U.S. Army Corps of Engineers 2006. Single Degree of Freedom Structural Response
Limits for Antiterrorism Design. PDC-TR 06-08, October.
Biggs, J. 1964. Introduction to Structural Dynamics, McGraw-Hill, Inc. New York,
NY, USA.
Petrochemical Committee. 1997. Design of Blast Resistant Buildings in Petrochemical
Facilities. ASCE, New York, NY, USA.
Defense Special Weapons Agency (DTRA). 1998. Design and Analysis of Hardened
Structures to Conventional Weapons Effects, Technical Manual for Army TM5-855-1,
Air Force AFPAM 32-1147(I), Navy NAVFAC P-1080, and Defense Special Weapons
Agency DAHSCWEMAN-97, August.
Lam, L., and Teng, J. 2003a. Design-Oriented Stress-Strain Model for FRP-Confined
Concrete. Construction and Building Materials, 17: 471-489
Lam, L., and Teng, J. 2003b. Design-Oriented Stress-Strain Model for FRP-Confined
Concrete in Rectangular Columns. Journal of Reinforced Plastics and Composites,
22(13): 1149-1186
ACI Committee 318. 2004. Building Code Requirements for Structural Concrete. ACI
318-05, American Concrete Institute, Farmington Hill, MI,
FYFE Co. LLC. 2010. Reinforced Concrete (RC) Columns Retrofitted With The
TYFO SCH System Subjected To A Close-In High Explosive. TR-10-33.4,
September.

592

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

EFFECT OF MECHANICAL LOAD ON THE THERMAL BEHAVIOR


OF CONCRETE SLAB REINFORCED WITH FIBRE REINFORCED
POLYMER (FRP) BARS
H. Bellakehal1, R. Masmoudi2, A. Zaidi3 and M. Bouhicha 4
1

PhD Candidate, Department of Civil Engineering, University of Laghouat, Laghouat, Algeria.


Professor, Department of Civil Engineering, University of Sherbrooke, Sherbrooke, Qc, Canada.
3
Professor, Department of Civil Engineering, University of Laghouat, Laghouat, Algeria.
4
Professor, Department of Civil Engineering, University of Laghouat, Laghouat, Algeria.
2

ABSTRACT
The corrosion problem in reinforced concrete structures has motivated researchers to
substitute steel bars with fiber reinforced polymer (FRP) bars because of their numerous
advantages such as high specific strength in tension and best performance in durability.
However, the thermal behavior is the main inconvenience of FRP due to the difference
between transverse and longitudinal coefficients of thermal expansion of these FRP bars.
Several experimental and analytical investigations have been carried out to characterize the
properties and behavior of FRP materials under the independent effect of thermal and
mechanical loads. The main topic of this study is to evaluate the combined effects of
mechanical and thermal loads on the transverse expansion of GFRP bars and concrete. This
paper focus on the effect of increasing the mechanical load beyond the service load of
concrete slabs reinforced on the thermal strain of concrete and bars. The results show that
the effect of a mechanical load of 34% of the ultimate load of reinforced concrete slabs on
glass FRP bar deformations is insignificant compared to that of a thermal load varied from 30 to +60C. The mechanical load increasing has not a significant effect on the thermal
strains of the bars, however, it incite the tensile concrete thermal strain.
1.

INTRODUCTION

Concrete structures are conventionally reinforced with steel bars. However, combinations
of moisture, temperature and chlorides may cause corrosion of steel in concrete structures
leading to the deterioration of concrete. This corrosion is a major problem that reduces the
service life of concrete structures, increases repairing costs and threats the safety of
constructions (Oehlers, D. and Seracino, R., 2009; El-Zaroug et al. 2007). The corrosion of
steel reinforcement in concrete structures has been a problem for years, especially in
bridges and marine structures. In some cases the repair costs can be twice as high as the
original cost. The use of non-metallic fiber reinforced polymer (FRP) reinforcement as an

593

alternative to steel reinforcement in concrete structure, particularly in hostile and aggressive


environments, is gaining acceptance mainly due to its high corrosion resistance, high
strength-to-weight ratio, high stiffness-to-weight ratio, ease of handling and fabrication,
high fatigue endurance, high damping and low thermal coefficient (in fiber direction)
(Pendhari et al. 2007). However, the different mechanical behavior of non-metallic
reinforcement, with respect to steel, involves some drawbacks namely the lack of thermal
compatibility between concrete and FRP reinforcement. It has been found that the
mechanical properties such as strength and stiffness of polymers decrease significantly as
the temperature increases to its glass transition temperature (Mutsuyoshi et al. 2004) and
significantly affects the serviceability of the structure. Experiments on FRP reinforcements
have shown that the transverse coefficient of thermal expansion (TCTE) of FRP bars is
typically much higher than its longitudinal CTE (Masmoudi et al., 2005). For glass fiber
reinforced polymers (GFRP) bars, while the longitudinal coefficient (LCTE) is similar to
that of concrete, the transverse coefficient (TCTE) is 3 to 8 times greater. Because of the
difference between the transverse coefficients of thermal expansion of FRP bars and
concrete, a radial pressure generated at the FRP bar/concrete interface induces tensile
stresses within the concrete under temperature increase. These tensile stresses may cause
splitting cracks within concrete and eventually degradation of the member stiffness. As a
consequence, important thermal strains take place just after apparition of the first micro
cracking concrete which occurs when the thermal stress in the concrete around the GFRP
bars, in different locations, reaches its tensile strength (ft). These induced thermal cracks
cause loss of the bond between GFRP bar and the surrounding concrete and, eventually,
failure of the concrete cover if the confining action of concrete is not sufficient (Zaidi et al.
2006). Although temperature effects on the mechanical properties of FRP are recognized by
all the documents reviewed, no guidance is given for the design of FRP reinforced
structures for temperature (Elbadry et al. 2004). Hence, a better understanding of the
thermal behavior of FRP reinforced concrete structures when subjected to large temperature
variations is still required. This paper presents an experimental study on the effect of
increasing the mechanical load beyond the service load of concrete slabs reinforced with
GFRP bars and subjected simultaneously to temperatures varied from -30 to +60 C.
2.

EXPERIMENTAL PROGRAM

2.1 Materials
2.1.1 Concrete
For each slab, six standard concrete cylinders of 150 x 300 mm were cast and cured with
water for 28 days at room temperature under the same conditions as the slab specimens.
Cylinders were tested to evaluate the compression and tensile strengths of concrete
immediately before slab specimen tests. The average values of compression and tensile
strengths of concrete were found to be equal to 36.5 MPa and 2.2 MPa, respectively.
2.1.2 FRP Bars
Each slab was reinforced with six V-Rod glass FRP (GFRP) bars of 15.9 mm diameter. The
mechanical and thermal properties of these bars are presented in Table 1.
594

Table 1. Mechanical and thermal properties of GFRP reinforcing bars.


Bars
CETT*
longitudinal Ultimate guarantee Ultimate Poisson's
Diameter Modulus of tensile
tensile
tensile
ratio
ft
elasticity, El strength strength
strain
[x10-6]
(GPa)
(MPa)
(MPa)
(%)
(mm)
/C
lt
15,9
48.2
751
683
1,56
0.25
26.0

CETL**
fl
[x10-6]
/C
6.95

* CTTE: Coefficient of Transversal Thermal Expansion.


** CLTE: Coefficient of Longitudinal Thermal Expansion.

2.2. Description and Instrumentation of Slabs


A total of three concrete slab specimens reinforced with GFRP bars were fabricated for this
experimental program. The slabs had dimensions of 500 mm wide, 200 mm thick, 2500
mm total length and 2000 mm span between supports as shown in Figures 1 and 2. The first
slab was subjected to temperature varied from -30 to +60 C (slab B). The second slab was
subjected to the both temperatures varied from -30 to +60 C and mechanical loads which
represents approximately 34%Mr (ultimate moment resistance of the slab) for the 1st
thermal loading cycle and 52%Mr for the 2nd thermal loading cycle (slab A). The third slab
of control was stored under room temperature (Slab C). The temperature was increased by
increments of 10C, each increment required 3 to 6 hours to reach a stable and uniform
temperature distribution throughout the slab. The mechanical load applied on the slab was
distributed on two transverse loading lines where each line received a load P of 34 KN for
the 1st thermal loading cycle and 52 KN for the 2nd thermal loading cycle, as shown in
Figure 1. Except mechanical and thermal loads, all other parameters were kept constant for
all slab specimens.

250
250

250

750
250

500
250

750

250

250

250

2000

250

30

h=200

GFRPbars#5

b=500 mm

Fig. 2. Cross section of the


slab.

Fig. 1. One-way slab simply supported.

The slabs were instrumented with strain gauges, thermocouples and LVDTs in order to
measure deformations, temperatures and deflections, respectively.
Each slab was instrumented as follows:
- One LVDT installed at mid-span to measure deflection; 12 strain gauges installed on the
main rebar at mid-span and 3/8 of the span: 6 gauges are designed to measure the
transverse strains and 6 others to measure the longitudinal strains (Figure 3), two
thermocouples installed on the rebar close to strain gauges to measure temperature at the
interface, and two thermocouples installed on the upper and lower surfaces of concrete

595

slabs, 8 strain gauges installed on the lower tensioned surface of the slab to measure the
tensile concrete strain: 4 gauges installed at mid-span and 4 gauge at 3/8 of the slab span
in transverse and longitudinal directions (Figure 4), 6 strain gauges installed on the upper
compressed surface of the concrete slab: for the 4 longitudinal strain gauges: two located
at mid-span and two installed at 3/8 of the slab span. For the 2 transverse strain gauges:
one installed at mid-span and one installed at 3/8 of the slab span.

Longitudinale strain
gauges

Transversestrain
gauges

35

5x68

35

1250

250

1000

Fig. 3. Position of strain gauges on GFRP bars.

35

215

250

250

1000

Fig. 4. Position of strain gauges on the tension concrete side of slabs


2.3. Test Procedure
Concrete slab specimens (A) and (B) were subjected to temperatures varied from 20 to -30
C and from -30 to +60 C with an increment of 10 C using the thermal room shown in
Figure 5. Once the temperature had stabilized in the concrete and GFRP bars for each
increment, all temperature and strain gauge readings were recorded. Then, the samples
were examined visually to note any cracks. In addition to the thermal load, the slab (A) was
submitted to a mechanical load of 68 KN which represents about 34% of the ultimate load
(Pu) of slabs. At the end of this thermal cycle, the same cycle was repeated but with a
mechanical load of 52%Pu of slabs. After all thermal cycles, the three slab specimens (A, B
and C) were subjected to four-point bending tests up to failure of slab specimens, the results
of the latest test are not presented in this paper.

596

1250
1000

strainofbar(m/m)

750
500
250
0
40

20

250 0

20

40

500
750
1000
1250
1500

60
80
Transvers
strain(Freezing)
longitudinal
starin(Freezing)
Transversstrain
(Heating)
Longitudinal
satrin(Heating)

Temperature(C)

Fig. 6. Transverse and longitudinal thermal strains of

Fig. 5. Thermal room.

GFRP bars in concrete slabs for cooling-heating cycle.


3.

ANALYSIS OF EXPERIMENTAL RESULTS

3.1 Thermo-Mechanical Behavior of FRP Bars


Figure 6 shows that the variation of thermal deformation curves in the cooling cycle from
+20 to -30 C coincide with those in the heating cycle from -30 to +20 C. This proves that
the thermal behavior of GFRP bars embedded in concrete slabs in actual scale is linear
elastic. The figure shows also, that the transverse strains are almost 4 times higher than the
longitudinal strains, which is in agreement with literature. This indicates that the applied
mechanical load does not affect the thermal behavior of the bars.
1500

400

Transversestrainofbar(m/m)

Longitudinalstrainofbar(m/m)

500

300
200
100
0
40

20

100

20

40

60

200

unloaded

300

loadedslab

80

1000
500
0
40

20

20

40

60

80

500
unloaded

1000

loadedslab
1500

400

Temperature(C)

Temperature(C)

Fig. 7. Longitudinal strains of GFRP bars


for the 1st loading cycle.

597

Fig. 8. Transverse strains of GFRP bars


for the 1st loading cycle.

At high temperature (+60 C) and under the effect of mechanical load of 34% of the
ultimate load (Pu) of slabs, the transverse thermal expansion of GFRP bars decreases with
6% compared to the transverse thermal expansion of GFRP bars without mechanical loads.
This is due to the Poisson's effect (Figure 8). Under the same mechanical load, the
longitudinal thermal expansion decreases with 30% compared to the longitudinal thermal
expansion of FRP bars without mechanical loads. This is due to degradation of the bond
between GFRP bars and concrete, particularly at high temperature for the slab (B) under
thermal loading only, because of the appearance of radial cracks within concrete at the
interface. However, for the slab (A) under thermal and mechanical combined loads, the
mechanical load contributes to the reduction of the radial pressure generated at the interface
due to transverse thermal incompatibility between concrete and FRP bar, consequently the
transverse expansion and eventually the radial cracks are reduced which improves the bond
between reinforcement and concrete (Figure 7). At low temperature, the effect of
mechanical load of 34%Pu of the slab has no significant effect on the longitudinal and
transverse contraction of GFRP bars (Figure 7 and Figure 8).
3.2. Effect of Mechanical Load on the Thermal Deformation
Figures 9 and 10 show a comparison between the mechanical load effect of 34% of the
ultimate load (Pu) with the mechanical load effect of 52%Pu on the thermal strain at the
interface FRP/concrete of the slab A (which subjected to thermal and mechanical combined
load). Figure 9 shows that the longitudinal strain increased by 10% with the increasing of
mechanical load in extreme temperatures (-30 and +60 C). Figure 10 shows that the curve
corresponding to 52%Pu was reduced down by 70 ms.
At high temperature, mechanical load acts in the same direction as the thermal load in terms
of longitudinal strain, and in the opposite direction in terms of transverse strain. At low
temperature, mechanical load acts in the same direction as the thermal load in terms of
transverse strain, and in the opposite direction in terms of longitudinal strain. Although it
can be concluded that the mechanical load increasing has no effect on the thermal strains of
the bars.
1500

400
300
200
100
0
40

20 100 0

20

40

60

80

200
300

load30%Mr

400

load50%Mr

Transversstrainofbar(m/m)

Longitudinalestrainofbar(m/m)

500

1000
500
0
40

20

Temperatue(C)

40

60
load30%Mr

1000
1500

500

20

500

load50%Mr
Temperatue(C)

Fig. 10. Transverse strain of bars.

Fig. 9. Longitudinal strain of bars.

598

80

1000

Transversetensileconcretestrain(m/m)

Longitudinaltensileconcretestrain(m/m)

800
600
400
200
0
40

20

20

40

60

80

200
400

load50%Mr

600

load30%Mr

800

800
600
400
200
0
40

20 200 0

20

40

60

80

400
600

load30%Mr

800

load50%Mr

1000

Temperature(C)

Temperature(C)

Fig. 12. Transverse tensile concrete strain.

Fig. 11. Longitudinal tensile concrete strain.

Figures 11 and 12 show a comparison between the mechanical load effect of 34% Pu with
the mechanical load effect of 52%Pu on the tensile concrete thermal strain of the concrete
slab A. These two figures show that the transverse and longitudinal tensile concrete thermal
strain where the slab was loaded by 52%Pu are larger compared to the case where the slab
was loaded by 34%Pu, for low and high temperatures. This implies that mechanical load
incite the tensile concrete thermal strain. It means the increase in mechanical load leads to
increased thermal deformations.
3.3. Effect of Thermal Cycling on Thermal Deformation
Figures 13 and 14 show a comparison between the second thermal cycle effect with the first
thermal cycle effect on the thermal strains at the interface FRP/concrete of the slab B
(which subjected only to the thermal load).
1500

400
200

20

1000
500
0

40

0
40

Transversstrainofbar(m/m)

Longitudinalestrainofbar(m/m)

600

20

40

60

200
1stcycle

400

80

20

20

40

60

80

500
1000
1stcycle
2ndcycle

1500

2ndcycle
600

2000
Temperature(C)

Temperature(C)

Fig. 14. Transverse strain of bars.

Fig. 13. Longitudinal strain of bars

These figures show that the longitudinal and transverse strains of the bars for the first
thermal cycle are higher than those for the 2nd thermal cycle by a constant value of strain,
from the low temperature to high temperature. This difference varies between 80 and

599

120s. This means that the bars keep the same behavior and the same coefficient of thermal
expansion, but they start the second cycle with a residual strain due to the previous cycle.
4.

CONCLUSIONS

The results show that the thermal and thermo-mechanical behavior of GFRP bars
embedded in concrete of actual slabs is linear elastic. This indicates that the applied
mechanical load does not affect the thermal behavior of the bars.

The effect of a mechanical load of 34% of the ultimate load (Pu) of reinforced concrete
slabs on GFRP bar deformations is insignificant compared to that of a thermal load
varied from -30 to +60 C, particularly for low temperatures. While, the longitudinal
thermal expansion of GFRP bars in the slab under thermal and mechanical combined
loads decreases by 30% compared to that of GFRP bars of the slab specimen under
thermal loads only (without mechanical loads). This is due to the degradation of bond
between GFRP bars and concrete for the slab subjected to just temperatures,
particularly at high temperatures.

The mechanical load increasing has not a significant effect on the thermal strains of the
bars, however, it incite the tensile concrete thermal strain.

The ratio c/db = 1.88 used in this study seems to be sufficient to avoid failure of the
concrete cover under combined mechanical load (34% and 52%Pu) and thermal load (30 to +60C).

5. AKNOWLEDGEMENTS

The research reported in this paper was partially sponsored by the Natural Sciences and
Engineering Research Council of Canada (NSERC). The authors also acknowledge the
contribution of the Canadian Foundation for Innovation (CFI) for the infrastructure used to
conduct testing. Special thanks to Pultrall Inc., QC, Canada for providing the FRP
materials.
6.

REFERENCES

1.

Oehlers, D., Griffith, M.C., and Seracino, R. 2009. Proceedings of the 9th International
Symposium on Fiber Reinforced Polymer Reinforcement for Concrete Structures
(FRPRCS 9), Sydney, Australia, 2009.
EL-Zaroug, O., Forth, J., Ye, J., Beeby, A. 2007. Flexural performance of concrete
slabs reinforced with GFRP and subjected to different thermal histories. Proceedings of
the 8th International Symposium on Fiber Reinforced Polymer Reinforcement for
Concrete Structures (FRPRCS-8), University of Patras, Patras, Greece, 16-18 July 10
p.
Pendhari, S.S., Kant, T., and Desai, Y.M. 2007. Application of polymer composites in
civil construction: A general review. Composite Structures, 84: 114124.

2.

3.

600

4.
5.
6.
7.

Mutsuyoshi, H., Zin, T., Sumida, A. 2004. Development of new heat resisting FRP
bars. Advanced Composite Materials in Bridges and Structures. Calgary, Alberta, 2023 July.
Masmoudi, R., Zaidi, A., and Girard, P. 2005. Transverse Thermal Expansion of FRP
Bars Embedded in Concrete. Journal of Composites for Construction, ACSE, 9(5):
377-387.
Zaidi, A. and Masmoudi, R. 2006. Thermal effect on FRP-reinforced concrete slabs. 1st
international Structural Specialty Conference. CSCE. Calgary, Alberta, Canada. ST036, 10 p.
Elbadry, M., Elzaroug, O. 2004. Control of cracking due to temperature in structural
concrete Reinforced with CFRP bars. Composite Structures. 64: 3745.

601

602

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

PROPOSED SHEAR MODEL FOR REINFORCED COCNRETE


BEAMS WRAPPED WITH FRP: A GENETIC ALGORITHM
APPROACH
C. Marshall, A. Rteil and M.S. Alam
School of Engineering, University of British Columbia, Kelowna, B.C., Canada

ABSTRACT
A large portion of the worlds infrastructure is composed of reinforced concrete that is
rapidly coming to the end of its useful service life. The use of externally bonded fibre
reinforced polymers (FRPs) as a repair and rehabilitation technique is quickly gaining
popularity in industry because it is economical, durable and superior material properties
compared to traditional repair materials and systems. Shear behaviour of FRP-repaired
reinforced concrete beams is very complicated and different models exist in the literature to
predict the shear strength of such beams. However, a previous study by the authors
(Marshall et al. 2011) has shown that the available shear design guidelines for reinforced
concrete beams wrapped with FRP (ACI 440.2R-08, CSA S6-06, CSA S806-02, fib 2001,
CIDAR 2006, CNR-DT200 2004) cannot accurately and reliably predict the FRP shear
contribution. In this paper, a new shear equation for the prediction of the FRP shear
contribution is proposed. Genetic algorithm was used to optimize the equation using over
400 beam specimens obtained from a thorough literature review. The proposed model
yielded a better shear strength prediction with less scatter compared to the existing design
codes and guidelines.
1. INTRODUCTION
Much of the worlds infrastructure is composed of reinforced concrete and the deterioration
of these structures is inevitable (NRC, 2010). Corrosion of reinforcing steel, constantly
changing live loads and exposure to different environmental conditions has caused
deterioration of these structures and eventually rendered some of them structurally
deficient. The use of externally applied fiber reinforced polymers (FRPs) can extend the
service life of reinforced concrete structures (Bonacci & Maalej 2000). FRPs are composed
of fibers embedded in a resin matrix. They have a high strength to weight ratio, are
corrosion resistant and can be easily installed making them an economic solution for repair.
FRP can be applied in three different wrapping schemes: side bonded, U wrapped, fully
wrapped or fully wrapping a member which is impractical in most rehabilitation cases.

603

Additionally, FRPs can be applied continuously or discontinuously (strips) along the length
of the member. Two types of fibers are typically used: carbon and glass.
The current design guidelines and codes (ACI 440.2R-08, CSA S6-06, CSA S806-02, fib
2001, CIDAR 2006, CNR-DT200 2004) do not accurately predict the FRP shear
contribution (Marshall et al 2011). These design guidelines are unreliable predicting both
very dangerous and very conservative FRP shear contributions (discussed later in this
paper). Many of these design guidelines have limitations in regards to strengthening and
wrapping schemes or require many detailed parameters that are not easily attained. A
complete and accurate model is needed for the design of reinforced concrete members with
externally applied FRP shear reinforcement.
This paper will focus on studying the shear capacity of a reinforced concrete member
repaired with externally applied FRPs. This will be done by analyzing the experimental
results in a database obtained from a thorough literature review to develop a model to
predict the FRP shear contribution. The model parameters will be optimized using Genetic
Algorithm technique. The concrete (VC), transverse steel (VS) and FRP (VF) shear
contributions will be treated separately, the sum of which is the ultimate shear capacity of
the member (Vu).
2. EXPERIMENTAL DATABASE
An experimental database was obtained from a thorough literature review. A complete list
of references can be found in Marshal et al. (2011). The database consisted of 299
specimens reinforced with externally applied FRP. Specimens with failure modes other
than shear or debonding were excluded; therefore only 224 specimens were considered in
this study. Figure 1 shows the breakdown of the specimens.

U Wrapped

55
102

20

Discontinuously

138

Completely
Wrapped

67

Continuously

86

Side Bonded

2
Carbon

Fiber
Rupture

98

Glass

126

202

Aramid

Fig. 1. Modified Experimental Database Details


604

Debonding

3. GENETIC ALGORITHM
The genetic algorithm (GA), developed by Holland (1975), is a powerful optimization
technique which can optimize complex nonlinear equations. This optimization technique is
based upon the theory of natural selection where superior traits (genes) are passed on to
subsequent generations. This is done by selecting the most fit genes, or elites, to pass
directly onto the next generation or by mating superior genes through either mutation or
crossover; thus creating new generations which are superior to the previous ones. This
process is then preformed until the solution converges.
In this study a multi-objective genetic algorithm, developed using MATLab, was used to
optimize the model. The objective function, Z, to be optimized is present in Equation 1.
Z

VF ,e
goal
VF , p

(1)

Where VF,e and VF,p are the experimentally determined and predicted FRP shear
contributions respectively. The goal variable ranged from 0.85 to 1.05 to ensure the most
accurate model was developed. The model constants were chosen from the generated set of
solutions (Pareto Front) based upon the average ratio of the experimentally determined to
predicted FRP shear contribution (VF,e/VF,p = 1) in addition to the smallest standard
deviation.
4. MODEL DEVELOPMENT
A sensitivity analysis was performed to determine which variables would be the basis for
optimization of the FRP shear contribution equation. This was done graphically in order to
determine the relationship between certain variables and the FRP shear contribution. Based
on this analysis, it was deemed that the concrete strength (fc), the beam web width (bw) and
the height of the FRP sheet or laminate (hF) showed a direct correlation to the FRP shear
contribution while the FRP stiffness (EFF) showed an inverse correlation. The relationship
between the FRP stiffness and FRP shear contribution is more evident when debonding is
the failure mode.
Results from the sensitivity analysis were used to develop a base or skeleton equation.
Numerous models were used in the optimization process as the base equation; however the
equation shown below (Equation 2) showed the best correlation to experimental results.
This base equation is similar to the model proposed by Adhikary et al. (2004). Values for
the constants, C1, C2, C3 and C4, were obtained using the GA optimization process.
VF F E F F ,e bw hF (sin cos )
Where
F ,e C1 ( f c )C 3

C 2 ( f c )2 C 3 1
F ,u F E F C 4

(2)
(3)

605

2 n F t F wF
bw s F
Number of FRP sheets or laminates
FRP thickness
Width of FRP strip
Spacing of FRP strips (center to center)
Beam web width
FRP modulus of elasticity
FRP fiber inclination angle with respect to the longitudinal axis of the beam
FRP effective strain
FRP ultimate strain
Concrete compressive strength (MPa)
FRP stiffness (GPa)

= FRP reinforcement ratio =

nF
tF
wF
sF
bw
EF

F,e
F,u
fc
F EF

=
=
=
=
=
=
=
=
=
=
=

The FRP effective strain was calculated depending on the wrapping scheme: Fully
wrapped, U wrapped or side bonded. It should be noted that fully wrapped beams
typically failed due to FRP rupture (96.4%), beams with side bonded FRP typically failed
due to debonding of the FRP (82.1%) and U wrapped beams generally failed due to
debonding (67.6%). Therefore, the failure mode was implicitly included in these equations.
The FRP effective strain is calculated as shown in Equation 3.
Fully Wrapped:

F ,e 0.0334 ( f c )1 3

0.0195 ( f c )2 3 1
0.414
F ,u
F E F

(4.1)

F ,e 0.0319 ( f c )1 3

1
F ,u
F E F 0.784

(4.2)

F ,e 0.00052 ( f c )2 3
Side Bonded:

1
F ,u
F E F 0.520

(4.3)

U Wrapped:

The concrete strength had little contribution to beams which are U wrapped or side
bonded because debonding of the FRP was the dominant failure mode in these cases.
Therefore, the second term in Equation 3 was dropped. This same trend was shown in the
proposed equation by Adhikary et al. (2004). The proposed equation is less accurate for
U wrapped beams than beams fully wrapped or side bonded with FRP. This is due to a
split in failure modes (Debonding, 67.3% or FRP rupture, 32.7%).
6. DISCUSSION
The average ratio of the experimentally determined to the predicted FRP shear contribution
was 1.03 with a standard deviation of 0.66. The predicted and experimentally determined
FRP shear contributions were plotted (Fig. 2) and a linear regression was performed. The
regression yielded a slope of 0.95.

606

For design purposes, a safety factor can be applied to the general FRP shear contribution
model (Equation 2). Safety factors were optimized to ensure the greatest reliability, or
smallest penalty value, based on the reliability analysis presented below. This resulted in a
safety factor of 1.7 for fully wrapped beams, 2.3 for U wrapped beams and 2.4 for beams
side bonded with FRP showed the best results. These values can be changed to modify the
level of confidence.
The proposed model was compared with design guidelines (ACI 440.2R-08, CSA S6-06,
CSA S806-02, fib 2002, CIDAR 2006 and CNR-DT200 2004). A weighted penalty system,
developed by Collins (2001), was used to rank each model (Table 1). In this system discrete
ranges of VF,e / VF,p are assigned a penalty based on the level of safety. Due to limitations
associated with the presented design guidelines and codes, the design guidelines could not
evaluate all the specimens in the database. For this reason, the total penalty value was
normalized by the number of beam specimens used in the analysis. It should be noted that
all safety or material reduction factors were used in this analysis, including the
aforementioned safety factors. The classifications and penalty values were adopted from
Lima and Barros (2007).

Predicted,V F,P(kN)

300
250
200
150

=0.95
100

SafeSide

50
0
0

50

100
150
200
Experimental,V F,e(kN)

250

Fig. 2. FRP Shear Contribution Predicted Vs. Experimental Results

607

300

Table 1. Comparison of the Proposed Model Vs. Design Guidelines and Codes
Classificatio
n

VF ,e

VF , p

Penalty
ACI440.2R
(2008)
CSA S6
(2006)
CSA S806
(2002)
fib
(2001)
CIDAR
(2006)
CNR-DT200
(2004)
Proposed
Equation

Reduced
Safety

Appropriat
e
Safety

Conservative

0.75 - 1.00

1.00 - 1.25

1.25 - 1.75

1.75 - 3.00

> 3.00

10

29

25

60

101

2.63

27

28

52

107

2.59

38

12

14

46

4.27

61

36

29

41

47

16

4.03

19

10

11

38

80

65

2.12

60

11

20

29

52

57

3.76

23

17

22

45

78

54

2.28

Very
Dangerous

Dangerous

< 0.75

Very
Conservative

Penalt
y/
#Beams

As shown in Table 1, the proposed design equation has a normalized penalty value of 2.28.
Only the CIDAR (2006) design guidelines outperformed the proposed equation having a
normalized penalty value of 2.12. In addition, the proposed equation was able to predict the
FRP shear contribution for 239 beam specimens, the most out of any design guideline
presented (various constraints in these design guidelines limit the number of specimens
used in the analysis). Furthermore, the majority of the predictions made by the proposed
equation were in the very conservative, conservative or appropriate safety range while most
of the predicted values of the presented design guidelines were in the very conservative or
very dangerous FRP shear contributions zone.
The main limitation of the proposed equation lies in its semi-empirical nature; its accuracy
is highly dependent on the accuracy and size of the database used in the optimization
process. For this reason, further experimental research is required in this field to develop a
larger database. Additional research will not only allow one to gain greater understanding
and insight into this topic but will also aid in the development of more accurate shear
design equations.
7. CONCLUSIONS
Genetic algorithm was effectively used to develop an equation to accurately predict the
FRP shear contribution of reinforced concrete beams wrapped with FRP. The results show
that there is a good correlation between experimental and predicted results. Furthermore,
the proposed equation better predicted the experimental results compared to the available
design guidelines and was the most versatile (utilized the greatest number of beam
specimens in the analysis) with most FRP shear contribution predictions within the
conservative and appropriate safety range.
608

8. REFERENCES
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.

Marshall, C., Rteil, A., and Alam, M.S. 2011. Critical Review of Shear Prediction
Models for Beams Externally Wrapped with Fiber Reinforced Polymers. Submitted to
Journal of Composites for Construction, ASCE, February, 33 p.
National Research Council (NRC). 2010. Highlights - Toward shock-proof
infrastructure. (http://news.gc.ca/web/article-eng.do?m=/index&nid=529829), National
Research Council Canada, accessed May 5.
Bonacci, J.F. and Maalej, M. 2000. Externally Bonded FRP Fro Service-Life
Extension of RC Infrastructure. Journal of Infrastructure Systems, 6(1): 41-51.
ACI Committee 440. 2008. Guide for the design and construction of externally
bonded FRP systems for strengthening concrete structures. ACI 440.2R-08, American
Concrete Institute (ACI), Farmington Hills, Mich., 76 p.
Canadian Standards Association (CSA). 2006. Canadian Highway Bridge Design
Code. CAN/CSA S6-06, Mississauga, Ontario, Canada.
Canadian Standards Association (CSA). 2002. Design and Construction of Building
Components with Fibre-Reinforced Polymers. CAN/CSA S806-02, Mississauga,
Ontario, Canada.
Federation international du beton fib. 2001. Bulletin 14 - Externally Bonded FRP
reinforcement for RC structures. Group 9.3 FRP (fibre reinforced polymer)
reinforcement for concrete structures.
CIDAR. 2006. Design guideline for RC structures retrofitted with FRP and metal
plates: beams and slabs. Draft 3 - submitted to Standards Australia, The University of
Adelaide.
CNR-DT200. 2004. Guidelines for design, execution and control of strengthening
interventions by means of fibre reinforced composites. National Research Council,
Rome, Italy.
Holland, I.H. 1975. Adaptation in Natural and Artificial Systems. University of
Michigan Press, Ann Arbor.
Adhikary, B.B., Mutsuyoshi, H., and Ashraf, M. 2004. Shear Strengthening of
Reinforced Concrete Beams Using Fiber-Reinforced Polymer Sheets with Bonded
Anchorage. ACI Structural Journal, 101(5): 660-668.
Collins, M.P. 2001. Evaluation of shear design procedures for concrete structures. A
Report prepared for the CSA technical committee on reinforced concrete design.
Lima, J.L.T., and Barros, J.A.O. 2007. Design Models for Shear Strengthening of
Reinforced Concrete Beams with Externally Bonded FRP Composites a Statistical
Versus Reliability Approach. Proceedings of the FRPCS-8 Symposium, University of
Patras, Patras, Greece, 16-18 July, 10 p.

609

610

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

THE EFFECT OF HIGH TEMPERATURE ON THE PROPERTIES OF


FRP REINFORCEMENT EXPOSED TO MOISTURE ENVIRONMENT
FOR LONG PERIODS
D.Y. Moon1, H. Oh2, J. Sim3 and M. Shahid4
l

Full-time instructor, Dept. of Civil Eng., Kyungsung Univ., Busan, Korea


Assistant professor, Dept. of Civil Eng., Gyeongnam National Univ. of Sci. and Tech., Jinju, Korea
3
Professor, Dept. of Civil & Env. Eng., Hanyang Univ., Ansan, Korea
4
Graduate Student, Dept. of Civil & Env. Eng., Hanyang Univ., Ansan, Korea
2

ABSTRACT
When FRP reinforced concrete structures are subjected to fire, high temperature above the
glass transition temperature at the level of reinforcement can cause severe damages to
polymer resin of FRP composite reinforcements. At this temperature, the polymer resin
softens significantly and loses its mechanical properties. Consequently, it is expected that
durability as well as strength capacity of the FRP reinforcement could be affected by the
high temperature. However, most experimental studies on long-term durability of FRP
reinforcement focussed on change of properties under curing temperature, ranged from
200C to 800C . In addition, many studies addressed only the evaluation of tensile strength
reduction of the FRP reinforcements subjected to high temperature. However, there is a
lack of information related to long-term durability of FRP rebars subjected to high
temperature above the glass transition temperature.
In this study, FRP reinforcements exposed in simulated solutions of pH 12.3 and 7 for long
periods were subjected to 600C, 1000C, 1500C, and 3000C examined. In order to investigate
the effect of the high temperature, weight changes and inter laminar shear strengths are
measured and compared to those of control ones.
1. INTRODUCTION
In reinforced concrete structures, the tensile force resistance is accomplished by internal
reinforcement. The internal reinforcement should not lose its tensile capacity as well as
bond properties in severe fir event in order to enhance the structural safety of the structure.
Therefore, decision to bring back the original or equivalent capacity should be quickly
made based on the evaluation of fire damage and of structural capacity.

611

Since the early 2000s, FRP reinforcing bars have been widely studied and used in a
concrete structure as an alternative to ordinary steel reinforcing bar. It is still indicated that
the FRP reinforcement could be significantly damaged and loose mechanical properties
under high temperature, resulting in rupture of the element. This problem is initiated from
the lower resistance to high temperature of constitution materials of the composite
reinforcing bars. Polymer resin matrix, which are protecting fibres and transferring force to
the fibres, loses it role at the temperature above glass transition temperatures.
Intensive researches on the reduction in strength and durability of FRP reinforcing bars
under high temperature have been conducted [1-4]. Although those are mainly focused on
evaluation of mechanical and bond properties of the composite reinforcing bars when only
exposed to high temperature, these results may be not helpful to assess the FRP reinforced
concrete structures damaged by fire. Actually, the internal reinforcement suffers from
alkalinity of concrete and moisture environment before and after a fire event. Particularly, a
decision of keeping or replacing the reinforcing bars exists in severely damaged zone by
fire should be done with a caution and based on the data obtained from the composite
reinforcing bars exposed to similar conditions to those embedded in actual concrete
structures.
In this study, the bare samples of FRP reinforcement were preliminary conditioned in
simulated solutions of pH 12.3 and 7 for long periods. This work is to simulate the status of
the internal FRP reinforcement used for long period in the concrete structure. the
conditioned specimens were subjected to 60, 100, 150 and 300 to simulate the
fire event. The weight changes and inter laminar shear strengths were evaluated in terms of
ratio of weight change and strength reduction.
2. EXPERIMENTAL PROGRAM
2.1 FRP Reinforcements
Three types of FRP reinforcements were used in this study as shown in Fig. 1. All
reinforcements used are domestic products, which are designated as A, B, and C,
respectively. The fiber, surface type, and diameter of all FRP reinforcements are different as
listed in Table 1. The three reinforcements have similar diameter but surface type and
material of those are different. Therefore, it was expected that we could see the effect of
resin and fiber type from the results. For each variable, two same specimens were used to
get averaged results.
2.2 Exposure Conditions
Firstly, all samples were conditioned in solutions for specified periods and those were
exposed to high temperatures just before the tests. Each exposure conditions are as follows.
Fig. 2 and Fig. 3 show FRP reinforcements conditioned in solutions and in heating
chamber, respectively. Moreover, Table 2 summarizes exposure conditions for all
specimens. In Table 2, the results for the underlined conditions does not presented in this
paper.

612

(a) Reinforcement A

(b) Reinforcement B

(c) Reinforcement C
Fig. 1. FRP reinforcements used in Experimentation

ID.

Table 1. Summarization of the used FRP reinforcements


Fiber and resin type
Surface type
Diameter

Glass, Epoxy

Deformed

9.53mm

Hybrid(Glass+Carbon), Vinyl-ester

Braided

9.50mm

Carbon, Vinyl-ester

Roughened

9.60mm

2.2.1 Solutions
Two types of exposure conditions were used for aging the specimens as listed in Table 2.
The first conditioning is the immersion in distilled water and the other conditioning is the
immersion in CaOH2 (pH 12.3) solution. This program was planned to continue for 300
days. Results were obtained at the date of 15, 30, 60, 90, 150, and 300 days. Only the data
at 15, 30, 60, and 90 days is shown in this paper. Fig. 2 shows the samples conditioned in
each solution.
2.2.2 High temperatures
As summarized in Table 2, FRP reinforcements were exposed to four different range of
temperatures. One of those is the temperature (600C) below glass transition temperature of
the resin matrix. Second exposure is done in the temperature near glass transition
temperature. The other two temperatures above glass transition temperature, 150 and
300, were used to simulate the damaged internal reinforcement by fire. After the
conditioning of the samples in solutions for the specified periods, those were heated in
heating chamber for the specified the exposure temperature and time in Table 2.

613

Fig. 2. Conditioning in solutions

ID.

A,
B,
C

Fig. 3. Exposure to high temperature

Table 2. Summarization of s exposure conditions


Conditioning in solutions (days)
Exposure to high Temperature
CaOH2(pH 12.3)
Distilled water(pH 7)
Duration (hr)
Temp.()

0, 15, 30, 60, 90,


150, 300

0, 15, 30, 60, 90,


150, 300

Room Temp.

60
100
150
300

0.5

2.3 Change in Weight


The 5 cm long pieces of samples were prepared and the edges of each specimen were
sealed with epoxy resin to prevent liquid diffusion during the conditioning. Before
immersion, the weight of five specimens was measured and averaged for initial weight ( Wi
) in equation (1). The specimens were immersed in solutions and then subjected high
temperature. Thereafter weighed to obtain Wt in equation (1).

W%

Wt Wi
Wt

(1)

2.4 ISS tests


After the exposures to high temperatures, the specimens were tested immediately according
to ASTM D4475 standard to investigate the reduction in mechanical properties of the
specimens. The shear strength of the specimens was determined in accordance with
equation (2).

P
2
db
Where P = Maximum force during the test
db = Diameter of specimens
S 0.849

(2)

614

3. TEST RESULTS
3.1 Weight Change Ratio
Figures 4~ 6 show weight change ratio of the tested specimens. Although it is measured
with equipment with high precision and resolution, the obtained data were quite scattered.
In order to see change pattern, trend line was made for each result as shown in all the
figures. The curves for 0 in all figures indicates that the results of the specimens only
exposed to high temperature without conditioning in solution, which means only the effect
of high temperature.
It was found that weight decreased with the increasing temperature from the curves of the
specimens only exposed to high temperature. Moreover, the reduction was to be more
severe at the temperature above 100 which is near the glass transition temperature. The
reduction in weight for the all specimens may be due to mass loss of resin matrix at the
temperatures between 100~300 [5]. In addition, this trend was shown in specimens
immersed in solutions for 15days, 30days, 60days, and 90days. Nevertheless, the weight of
specimens immersed in solutions thereafter exposed to high temperature decreased more in
elevated temperature compared to specimens exposed to only high temperature. This
experimental program is not finished yet. However, these results obtained until now
demonstrated that the combined effects of solution and high temperature were significant
more than that of only high temperature. In addition, an evaluation based on the empirical
study only considering the effect of high temperature could underestimate the integrity of
FRP reinforcement in concrete structure damaged by elevated temperature. Unfortunately,
the difference in effects of Ca (OH)2 and distilled water on weight change of specimens was
not clear from the results obtained until now.
5

0
0
50

100

150

200

250

300

350

15
30
90

WeightChangeRatio(%)

WeightChangeRatio(%)

50

100

150

200

250

300

350

15
30

90
10

15
4

20

Temp.()

Temp.()

(a) Ca (OH) 2

(b) Distilled water


Fig. 4. Reinforcement A immersion in solution

3.2 Inter laminar Shear Strength (ISS) strength


Figures 7~9 show ISS strength results of the tested specimens. For all tested specimens, ISS
strength decreased with increasing temperature. Similar results to those obtained from
weight change ratio were shown from the comparisons of the trend lines for specimens only
exposed to high temperature with the trend lines for specimens conditioned in solutions and
615

exposed to high temperature. Furthermore, it is expected that the decrement of ISS for the
specimens immersed for more long periods, 150 days and 300days tend to be more severe
due to degradation of FRP reinforcement.
10

0
15
30
60

WeightChangeRatio(%)

0
2

15

0
0

50

100

150

200

250

300

WeightChangeRatio(%)

350

50

100

150

200

250

300

350

30
90

10

12

Temp.(C)HYAL

Temp.(C)HYDI

(a) Ca (OH)2
(b) Distilled water
Fig. 5. Reinforcement B immersion in solution
2

0.5

0
50

100

150

200

250

300

350

100

150

200

250

300

350

15

WeightChangeRatio(%)

WeightChangeRatio(%)

50

0.5

30
1

1.5

15
30

90

10

12

14
2.5
16

18

Temp.(C)CARAL

Temp.(C)CARDI

(b) Distilled water


(a) Ca (OH)2
Fig. 6. Reinforcement C immersion in solution
60

0day

55

15day

55

50

30day

50

0day
15day
30day
60day

60day
45

90day

40
35

ISS(MPa)

ISS(MPa)

60

45
40
35

30

30

25

25

20

90day

20

50

100

150

200

250

300

Temp.()

50

100

150

200

Temp.()

(a) Ca (OH)2
(b) Distilled water
Fig. 7. Reinforcement A immersion in solution

616

250

300

60
55

30day

40
35

ISS(MPa)

90day

15day
30day

50

60day

45

0day

55

15day

50

ISS(MPa)

60

0day

90day

40
35

30

30

25

25

20

60day

45

20

50

100

150

200

250

300

50

Temp.()

100

150

200

250

300

Temp.()

(a) Ca (OH)2
(b) Distilled water
Fig. 8. Reinforcement B immersion in solution
60

60
55
50

15day

55

30day

50

15day
30day
60day

60day

45

90day

40
35

ISS(MPa)

ISS(MPa)

0day

0day

45

90day

40
35

30

30

25

25
20

20
0

50

100

150

200

250

300

50

100

150

200

250

300

Temp.()

Temp.()

(b) Distilled water


(a) Ca (OH)2
Fig. 9. Reinforcement C immersion in solution
4. CONCLUSIONS
In this study, the effect of exposure only to high temperature and combined immersion in
solutions and high temperature on the weight change and inter laminar shear strength of
FRP reinforcements was investigated. This experimental program is still ongoing. The
findings until now are as follows:

The reductions in weight and inter laminar shear strength (ISS) were shown in all
specimens exposed to high temperature. Nevertheless, the rate of reduction
significantly affected by the degree of degradation of FRP reinforcement in concrete
structure based on the results obtained until now. From these results, we can found that
the used FRP reinforcement could be affected by high temperature more than newly
installed FRP reinforcement.

The effects of solution types used, fiber, and resin types of FRP reinforcement were not
clear until now.

617

For a making proper decision whether keeping or replacing of the damaged FRP
reinforcement in concrete structures by high temperature, it is required to study the
combined effects of temperatures above glass transition temperature and moisture
environments in concrete.

5. REFERENCES
1.
2.
3.
4.
5.

Masmoudi, R., Masmoudi, A., Ouezdou, M.B., and Daoud, A. 2011. Long-term bond
performance of GFRP bars in concrete under temperature ranging from 20 to 80.
Construction and Building Materials, 25(2): 486-493.
Kodur, V.K.R., and Dwaikat, M. 2008. A numerical model for prediction the fire
resistance of reinforced concrete beams. Cement & Concrete Composites, 30: 431-443.
Robert, M., and Benmokrane, B. 2010. Effect of aging on bond of GFRP bars
embedded in concrete. Cement & Concrete Composites, 32: 461-467.
Abbasi, A., and Hogg, P.J. 2005. Temperature and environmental effects on glass fibre
rebar: modulus, strength and interfacial bond strength with concrete. Composites Part
B, 36: 394-404.
Chin, J.W., Hughes, W.L., and Signor, A. 2001. Elevated temperature aging on Glass
Fiber Reinforced Vinyl ester and isophthalic polyester composites in Water, Salt water,
and concrete pore solution. American Society for Composites, 16th Technical
Conference, Blacksburg, VA, pp. 1-12.

618

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

FLEXURAL BEHAVIOR OF SIMPLE AND FIXED-END BEAMS


STRENGTHENED WITH FRP BARS IN NSM METHOD
M.K. Sharbatdar1, M. Mohamadian2 and S.M. Jaberi2
1
2

Assistant Professor, Faculty of Civil Engineering, Semnan University, Semnan, Iran.


MSC Structural Engineer, Semnan, Iran.

ABSTRACT
The corrosion of steel bars in reinforced concrete members and the resulting deterioration
of structures prompted research on fiber reinforced polymers (FRP) bars as potential
reinforcement for concrete members, for use in new construction. These FRP bars can be
used to increase flexural and shear capacity of existing concrete members at NSM (Near
Surface Mounted) method. A comprehensive experimental research program was
conducted to investigate the behavior of RC beams strengthened with FRP bars in order to
increase their flexural capacity.
Four large-scale reinforced concrete beams at two different groups were designed,
constructed, and tested under three point concentrated loading system. Two specimens had
simple support and two others had both fix-end support. Two specimens at each group had
same size and reinforcement characteristic. The paper presents the details and results of the
experimental programs. Moreover extra analytical work has done, the test result has been
used to calibrate the analytical models.
The results indicate that FRP bars can be used effectively in existing structures to increase
their flexural capacity and change their crack pattern. Photographs taken at selected stages
of loading illustrated the performance of each specimen. The force-displacement and forcestrain relationships of both groups were presented and compared to find efficiency of FRP
bars at NSM method. Results showed that simple beams had more increasing capacity
compare to fixed-end beams due to less difference of experimental ultimate flexural
capacity with expected capacity calculated based on Design code expressions.
1. INTRODUCTION
A large number of existing buildings and bridges overall the world are in need of
strengthening or retrofitting due to factors such as deterioration, construction or design
faults, additional load or functional change. Fiber Reinforced Polymer (FRP)
reinforcements have been recently used extensively as an alternative reinforcement material

619

to steel for new construction as well as for strengthening and repair of existing concrete
structures. Toimprove utilization of the FRP materials, near-surface mounted (NSM)
reinforcement was recently introduced as a promising technique for strengthening of
reinforced concrete members. Application of NSM FRP reinforcement does not requires
surface preparation work and after cutting the slit, requires minimal installation time and
cost compared to the externally bonded reinforcing (EBR) technique. After installation, the
NSM FRP reinforcements are protected against mechanical damage, wear, impact,
vandalism, and fire and aging effects. This aspect makes this technology particularly
suitable for the strengthening of negative moment regions of beams and slabs. This paper
focuses on research work on the structural aspects of NSM strengthening of concrete
beams.
All previous test results indicate that the NSM reinforcement improved the ultimate load
and the load at the yielding of steel reinforcement, as well as the post-cracking stiffness [7].
Based on the available experimental evidence reported in researches, the possible failure
modes of beams flexural-strengthened with NSM FRP reinforcement are different such as
Barepoxy interfacial de-bonding, - Concrete cover separation, Bar end cover separation,
Localized cover separation, Flexural crack-induced cover separation, Beam edge cover
separation, and
Epoxyconcrete interfacial de-bonding. The use of NSM FRP
reinforcement is also effective to improving shear capacity of RC beams. De Lorenzis and
Nanni [18] carried out eight tests on large size T-section beams, of which six had no
internal stirrups. CFRP ribbed round bars in epoxy-filled grooves were used as NSM shear
reinforcement. The NSM reinforcement produced a shear strength increase which is as high
as 106% in the absence of steel stirrups, and still significant in presence of a limited amount
of internal shear reinforcement. Barros and Dias [2] tested beams of different sizes with no
internal stirrups. Some of these beams were strengthened with NSM CFRP strips of
different inclinations, while the rest were strengthened with equivalent amounts of
externally bonded FRP shear reinforcement. The reported strength increases ranged from
22% to 77%, and were in all cases larger than those obtained with externally bonded FRP
and some of the beams are believed to have failed in bending.
2. EXPERIMENTAL PROGRAM
To assess the effectiveness of NSM strengthening techniques to increase the flexural
resistance of RC beams, two experimental groups was selected, a group of flexural simple
support RC beams and a group of two-ended fixed RC beams. To simulate a real
strengthening situation, flexural beams had a limited amount of longitudinal reinforcement,
respectively.
3. MATERIAL PROPERTIES
Prior to making of the beams, material characterization was carried out on concrete, steel
reinforcements and CFRP bars. The concrete average compression cylinder specimens in
simple two-end beams and two-end beams at the time of testing were 27 MPa and 46 MPa.
The steel bars used in the experimental program have diameter between 6-14 mm with
average tensile stress between 200-450 MPa. A new type of CFRP bar was proposed in this
study; the CFRP bars were manually made in laboratory and were hooked at end for
620

anchoring in concrete. The CFRP bars had a circle cross section and made of three
ingredient; i) CFRP sheet, ii) epoxy resin and iii) wooden bar. A high strength
unidirectional carbon sheet with 0.111 mm nominal fiber thickness, impregnated in situ
with epoxy resin by the wet lay-up technique. The nominal fiber content was 60% by
volume. The manufacturers values of tensile strength and elastic modulus of CFRP sheets
were equal to 3550 MPa and 235 GPa, respectively, yielding an ultimate tensile strain of
1.5%. The adhesives were epoxy XH-111 with 76.1 MPa flexural strength, as provided by
the supplier. Fabrication process of man-made CFRP bars is shown in Fig.1.
Mechanical properties of FRP bars were established through tension coupon test with the
procedure specified in CSA standard S806. Stress-strain relationship of FRP bars is shown
in Fig. 2. indicating that CFRP bars show linear elastic behavior.
2400
2100

Stress (Mpa)

1800
1500
1200
900
600
300
0
0.000

0.002

0.004

0.006

0.008

0.010

Strain (mm/mm)

Fig.1. Fabrication of FRP bars

Fig. 2.Stress-strain of FRP bars

4. DETAILS OF CONCRETE BEAMS


Geometry, reinforcement arrangements, loading and supporting conditions of the flexural
beams are represented in Fig. 3.a. The Simple two-end beams are containing of two
inverted T-section beams with overall length of 1800 mm, clear span of 1600 mm and
depth of 300 mm. The tensile steel reinforcement was composed of two 14 mm diameters
bars, while the top reinforcement consisted of two 10 mm diameter steel bars. Shear
reinforcement were made of closed double-legged stirrups with 10 mm diameter, uniformly
spaced at 100 mm. Shear reinforcement was selected to prevent shear failure prior to
bending failure for all beams. The beams were simply supported at the ends and were
loaded under two-point symmetrical bending.
The Fix two-end beams contained of two rectangular beams with overall length of 2420
mm, clear span of 1820 mm and depth of 200 mm. The tensile and compression steel
reinforcement was composed of three 10 mm diameters bars each ones. Shear
reinforcement were made of closed double-legged stirrups with 8 mm diameter, uniformly
spaced at 70 mm. Shear reinforcement was selected to prevent shear failure prior to
bending failure for beams. The beams were fixing supported at the ends with two columns
in the each side of beam, its detail shows in Fig 3.b. The fix two-end beams join to rigid
frame with 16 bolt in columns. Beams in this group were loaded under concentrated load.
5. STRENGTHENING SYSTEMS
In each series of beams, one beam was tested without strengthening as a control specimen,
BS1 and BF1 for simple and fix two-end beams respectively. The other beams,BS2 and
621

BF2 for simple and fix two-end beams respectively that were strengthened with using NSM
CFRP bars. Simple two-end beams was strengthened on each side of the centre line of the
specimen with 2 NSM CFRP Bar with total cross section of about 16 mm2 and fix two-end
beams was strengthened with 1 NSM CFRP Bar in each tensile part of top and bottom of
beams with cross section of about 8 mm2 for each bar. Other parameters such as type of
groove filling epoxy, dimension of grooves and space of the NSM reinforcement, were kept
constant. The beam BS2 was strengthened with embedment lengths of about 85 percent of
the beam length (1500mm) and beam BF2 was strengthened as shows in Fig 4.

a. Flexural simple two-end beams

b. Flexural fix two-end beams


Fig. 3. Geometrical details of test specimens

BS2 Specimen

BF2 Specimen

Fig.4. Locations of NSM CFRP Bars in specimens.


622

6. TEST SETUP AND INSTRUMENTATION


Prior to test of the beams, strain gauges were attached at critical positions on the steel
reinforcement and FRP bars. Strain gauges were also attached to all concrete beams at the
mid-span, to measure the concrete strain during loading. The beams were instrumented with
three LVDTs at mid-span on both sides of that in bottom side, to monitor the average of
deflection. The load was applied to each beam at a rate of 5 kN/step by means of the
hydraulic jacks. Displacements and strains were all recorded by an electronic data logger
system with a 1-Hz sampling rate.
7. TEST RESULTS AND DISCUSSION
After cracking and prior to yielding point of the tensile reinforcements, the flexural
behavior for all simple two-end strengthened beams was almost similar to un-strengthened
beam. This behavior indicates that using NSM FRP reinforcements did not contribution to
increase the stiffness in the elastic range. However, after yielding of the tensile
reinforcements, the flexural stiffness and strength of the NSM CFRP beams were
significantly improved compared to the control beam. After yielding of the steel
reinforcement, the increase in the applied load was observed until shear failure occurred
after de-bonding observed at one end of the beam at cut-off section of NSM CFRP bars, as
shown in Fig. 5. The beam BS2 with an embedment length of 1500 mm (about 85% of
beam length) had an 29 percent increase in the ultimate strength in comparison to the
control beam ,BS1.The formation of cracks in beam BS1, followed a typical crack pattern
of flexural members, so beam BS1 failed in due to yielding of tensile reinforcements and
crushing of concrete. As the external load increased, flexural cracks gradually formed
throughout constant moment region. Visual inspection after failure of beam indicated that
the separation of concrete cover occurred at the end of FRP bars. Finally, after de-bonding
of the CFRP bars, the load carrying capacity of beam decreased. Failure mode of beam can
be supposed as shear-flexural failure. Using NSM FRP reinforcement resulted in significant
reduction of the crack widths and increased quantity and propagation of new cracks in the
strengthened beams compared to control beam.
Two-end fixed beam specimen BF2 had a FRP bar at each tension area, the top of the end
with negative and the bottom of the middle with positive moments. First crack was
happened at the load 40 kN with corresponding deflection 1.39 mm. The tension steel bars
were yielded at the load 200 kN, deflection 13.4 mm. The final load capacity and deflection
were respectively 250 kN and 24.04 mm. The flexural failure mode was observed as shown
in Figure 6, and also most of cracks at tension zones were flexural. The FRP bar rupturing
and concrete crushing were observed at the end of test. The initial mounted bars were
calculated based on the design code nominal moment of the specimen. Since the unstrengthened specimen BF1 showed the much higher capacity compared to analytical
nominal capacity, it was seemed that the real mounted bars were less than the effective bar
numbers so the FRP bars stresses were increased very soon and they were ruptured and
could not increased the specimen BF2 capacity. Therefore both specimens have almost
same capacity with different ductility.

623

Fig. 5. Shear failure of beam BS2

Fig. 6. Flexural failure of beam BF2

B2

160

B3

140

Load (kN)

120
B1

100
80
60
40

a. simple two-end beams

b. fix two-end beams

20
0
0

10

15
20
Displacment (mm)

25

30

35

Fig. 7. Load-Mid-span Deflection behavior


Table 1. Experimental results of flexural beams
Increase
cr
U
Y
PY
PU
beam Pcr(kN)
in PU
Failure Mode
(mm) (kN) (mm) (kN) (mm)
(%)
BS1
30
1.7
118 12.5
124
33.0
Flexural Failure
SCC ;
BS2
32
1.4
130
8.5
160
20.5
29
Shear Failure
BF1
50
1.77 180 9.87 259.7 27.2
Flexural Failure
BF2
40
1.39 200 13.4
250 24.04
Flexural Failure
Pcr: Cracking Load ; cr : Midspan Deflection at Cracking Load
PY : Yielding Load ; Y : Midspan Deflection at Yielding Load
PU : Ultimate Failure Load; U : Midspan Deflection at Failure
Scc : Separation of Concrete Cover at the ends of the FRP bars
8. CONCLUSIONS
To appraise the effectiveness of a NSM strengthening technique for flexural strengthening,
concrete beams tests were carried out. The following conclusions can be drawn from the
outcomes of this experimental research:
Before yielding point of the tensile reinforcements, the load-deflection behavior of
beams was similar, so using NSM FRP reinforcements improve the stiffness of
specimen in plastic range.
Strengthened flexural beam, BS2, showed an increase in ultimate load about 29
percent, a decrease in both deflection and curvature about 61%, and significant
reduction of the crack widths and increased quantity and propagation of new cracks in
624

the strengthened beams over the control beam. Maximum strain of CFRP bars obtained
prior to de-bonding, was 1.03 and 1.08 percent, indicating proper utilization of the
tensile strength of the CFRP bars was achieved.
The strengthened fixed-end specimen ductility was decreased about 13% compared to
un-strengthened specimen with almost same capacity ductility. Initial suggested FRP
bars were based on design code nominal capacity, but the real mounted FRP bars were
less than required bars based on the real moment capacity and ruptured very son and no
significant increasing were not shown. More FRP bars based on the real capacity can
increased the un-strengthened specimen capacity.
It was verified that the use of the proposed man-made CFRP bars was feasible, easy to
apply, and effective in strengthening concrete beams. The proposed CFRP bars
increased the ultimate load-carrying capacity and improved overall behavior of
Strengthened beams.

9. REFERENCES
1.

ACI Committee 440. 2008. Guide for the design and construction of externally bonded
FRP systems for strengthening concrete structures. ACI 440.2R-08, American
Concrete Institute (ACI), Farmington Hill, MI, 76 p.
2. Barros, J.A.O and Dias, S. 2003. Shear strengthening of reinforced concrete beams
with laminate strips of CFRP. Cosenza (Italy), pp. 289294.
3. Asplund, S.O. 1949. Strengthening Bridge Slabs with Grouted Reinforcement. ACI
Structure Journal, 45(1): 397-406.
4. Blaschko, M., and Zilch, K. 1999. Rehabilitation of Concrete Structures with CFRP
Strips Glued into Slits. Proceedings of the 12th International Conference on Composite
Materials, Paris, France.
5. El-Hacha R. and Rizkalla, S.H. 2004. Near-surface-mounted fiber reinforced polymer
reinforcements for flexural strengthening of concrete structures. ACI Structural
Journal, 101(5): 717726.
6. Tang, W.C., Balendran, R.V., and Leung, H.Y. 2006. Flexural Strengthening of
Reinforced Lightweight Polystyrene Aggregate Concrete Beams with Near Surface
Mounted GFRP Bars. Building and Environment, 41: 1381-1393.
7. De Lorenzis, L. and Teng, G.J. 2007. Near Surface Mounted FRP Reinforcement: An
Emerging Technique for Strengthening Structures. Composites Part B: Engineering,
38: 119-143.
8. Carolin, A., Hordin, H., and Taljsten, B. 2001. Concrete Beams Strengthened with
Near Surface Mounted Reinforcement of CFRP. Proceeding of the International
Conference on FRP Composite in Civil Engineering, CICE, 2001, Hong Kong, China,
12-15 December.
9. De Lorenzis, L. and Nanni, A. 2002. Bond between NSM fiber-reinforced polymer
rods and concrete in structural strengthening. ACI Structural Journal, 99(2): 12332.
10. El-Hacha, R. and Rizkalla, S.H. 2004. Near-surface-mounted fiber reinforced polymer
reinforcements for flexural strengthening of concrete structures. ACI Structural
Journal, 101(5): 717726.
11. Hassan, T. and Rizkalla, S. 2004. Bond mechanism of near-surface-mounted fiber
reinforced polymer bars for flexural strengthening of concrete structures. ACI
Structural Journal, 101(6): 830839.
625

12. Arduini, M., Gottardo, R., and De Riva, F. 2001. FRP Rods for Flexural Reinforcement
of Existing Beams: Experimental Research and Applications. Proceeding of the
International Conference on FRP Composite in Civil Engineering, CICE, 2001, Hong
Kong, China, 12-15 December, 2: 1051-1058.
13. De Lorenzis, L. 2002. Strengthening of RC structures with near surface mounted FRP
rods. PhD Thesis, Department of Innovation Engineering, University of Lecce, Italy.
14. Parretti, R. and Nanni, A. 2004. Strengthening of RC members using near surface
mounted FRP composites: design overview. Advances in Structural Engineering An
International Journal, 7(6): 469-483.
15. De Lorenzis, L. and Nanni, A. 2001. Shear Strengthening of Reinforced Concrete
Beams with Near-Surface Mounted FRP Rods. ACI Structural Journal, 98(1): 60-68.
16. Barros, J.A.O and Dias, S. 2003. Shear strengthening of reinforced concrete beams
with laminate strips of CFRP. Cosenza (Italy), pp. 289294.
17. Nanni, A., Di Ludovico M., and Parretti, R. 2004. Shear strengthening of a PC bridge
with NSM CFRP rectangular bars. Advances in Structural Engineering, 7(4).
18. De Lorenzis, L., Nanni, A., La Tegola, A. 2000. Flexural and shear strengthening of
reinforced concrete structures with near surface mounted FRP rods. In: Proceedings
ACMBS III, Ottawa, Canada.
19. ACI Committee 318. 2005. Building code requirements for structural concrete. (ACI
318-05) and commentary (ACI 318R-05). American Concrete Institute (ACI),
Farmington Hills, MI.
20. Sharbatdar, M.K., 2003. Concrete Columns and Beams Reinforced with FRP Bars and
Grids under monotonic and Reversed Cyclic Loading. PhD Thesis, Department of
Civil Engineering, University of Ottawa, Canada.
21. Mohammadian, M. 2010. The Investigation of Flexural Behavior of RC Fixed-End
Beams StrengthendWith FRP Bars at NSM Method. MS Thesis, Department of Civil
Engineering, University of Takestan, Iran.
22. Jaberi, S.M. 2009. Strengthening of Reinforced Concrete Beams with Near Sarface
Mounted CFRP Bars. MSc Thesis, Department of Civil Engineering, University of
Semnan, Iran.

626

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

FATIGUE PERFORMANCE OF RC BEAMS STRENGTHENED WITH


PRESTRESSED NSM-CFRP STRIPS
F. Oudah1 and R. El-Hacha2
1
2

MSc student, Dept. of Civil Engrg., Univ. of Calgary, Canada


Associate Professor, Dept. of Civil Engrg., Univ. of Calgary, Canada

ABSTRACT
This paper evaluates the fatigue performance of Reinforced Concrete (RC) beams
strengthened using prestressed and non-prestresssed Near Surface Mounted (NSM) Carbon
Fibre Reinforced Polymer (CFRP) strips. Three 5m long RC beams were tested under
fatigue conditions representing in-service loading; control un-strengthened beam, beam
strengthened using non-prestressed NSM CFRP strips, and beam strengthened using NSM
CFRP strips prestressed to 20% of its ultimate tensile strength. Upper and lower load
fatigue limits were chosen such that to induce a stress range of 125 MPa in the tension
reinforcement, with upper stress limit of 70% of the steel yielding stress, during the initial
static cycle. NSM strengthening using prestressed and non-prestressed reinforcement was
found to enhance the serviceability concerns, reduced deflection and enhanced stiffness,
under cyclic loading. Both strengthened beams experienced debonding at initial fatigue
cycling manifested in reduction in CFRP strains followed by a stabilized increase.
However, prestressing the CFRP strips enhanced the bonding properties, reduced the initial
stress reduction, and resulted in overall stress increase at the completion of the fatigue
loading as compared with the non-prestressed beam.
1. INTRODUCTION
The high strength-to-weight ratio, ease of installation, and long term durability make Fibre
Reinforced Polymer (FRP) materials favourite choice for strengthening and repairing
damaged structures. Lots of researches were conducted to investigate the performance of
Reinforced Concrete (RC) structures strengthened with FRP in shear and flexure. However,
limited researches were devoted to study its long-term in-service performance. According
to BS 5400: Part 10, the code specifies a bridge design life of 120 years and 2 million
standard fatigue vehicles per year on the slow lane of a three-lane motorway (Cadei, 2003).
This high number of expected load cycles applied on a typical bridge girder initiates the
need to study the durability of FRP strengthened RC girders in terms of serviceability and
damage accumulation.

627

Near Surface Mounted (NSM) technique was developed to overcome the drawbacks
associated with externally bonded FRP sheets and plates such as debonding. FRP strips or
rods are embedded inside pre-cut grooves into the concrete cover filled with epoxy.
Embedding the FRP inside the groove protects it from external environmental conditions
and avoids possible mechanical damage.
NSM strengthening demonstrates a very controlled damage accumulation rate as compared
with other strengthening techniques when submitted to cyclic loading (Quattlebaum et al.,
2005). Test results indicate that the composite action is maintained during the fatigue and
post-fatigue loading. However, stiffness degradation was observed within the elastic
portion of the load-deflection curves due to the cyclic loading (Youst et al., 2007). The
strengthening efficiency is further enhanced by means of prestressing the FRP
reinforcement. Prestressing the NSM-FRP reinforcement addresses the serviceability
concerns in terms of reduced deflection, reduced crack width, and higher yielding load.
Test results on RC bridge girders strengthened with prestressed CFRP-NSM strips indicate
very limited degradation throughout the fatigue loading as compared with externally
bonded FRP strips (Rosenboom and Rizkalla, 2006; Badawi, 2007).
2. EXPERIMENTAL PROGRAM
The experimental program consisted of testing five 5.0 m long RC beams. One control unstrengthened beam (B0-00), one beam strengthened with non-prestressed NSM-CFRP
strips(B1-00), and three beams strengthened with CFRP strips prestressed to 20%, 40% and
60% of its ultimate tensile strength (B1-20, B1-40 and B1-60). However, only the results of
beams B0-00, B1-00 and B1-20 are reported in this paper. Each beam was strengthened
with two 216 strips bonded to each other using epoxy. Beams details and dimensions are
shown in Figure 1. The amount of CFRP reinforcement was chosen to increase the ultimate
flexural capacity of the strengthened beams by 54%. An innovative prestressing system,
developed at the University of Calgary, was used to prestress the CFRP reinforcement by
means of reacting against the beam itself (Gaafar, 2007).
The average 28 days concrete compressive strength of three 100200 mm cylinders is
40.41 MPa. CFRP ASLAN 500 Tape was used as strengthening material. The tensile
strength of the CFRP strips is 2068 MPa while the modulus of elasticity is 124 GPa. Steel
bars 15M and 10M were used for tension and compression reinforcement, respectively. The
mechanical properties of the steel reinforcement is as follows; modulus of elasticity of 200
GPa and yielding strength of 440 MPa. Two strain gauges (SG) were attached to the tensile
steel reinforcements and two SG were attached to the compression steel at the mid-span of
the beam. Strains at top fibre, top steel level, and bottom steel level were monitored using
Linear Strain Conversion (LSC) attached on the web at both sides of the beam. The CFRP
strips were instrumented with SGs distanced at 500, 1250, 1850 mm from the center of the
beam on both sides. Two SGs were attached at the mid-span of the CFRP strips while one
SG was attached at each other location.
Load was applied via a servo-hydraulic actuator reacting against a steel frame anchored to
the laboratory strong floor. The fatigue life of a RC beam depends on the stress range in the
steel reinforcement. Therefore, to make a valid comparison, it is reasonable to apply the
628

same stress limits on both the strengthened and un-strengthened beams. Consequently, the
maximum and minimum load limits would differ due to the application of the CFRP
reinforcement and due to prestressing, being the highest for B1-20 and the lowest for B000. The stress range adopted in this study is based on the Canadian Highway Bridge
Design Code (CSA S6-06) which specifies a maximum stress range of 125 MPa while the
maximum stress limit was chosen to be 0.7fy, where fy is the steel yield stress.

(b) Cross-section (mm)


(c) Groove (mm)
Fig.1. Beam details and test set-up.
The testing protocol consisted of the application of three types of loadings: Quasi-Static
loading, fatigue loading, and post-fatigue monotonic loading. Only test results obtained
from the first two types of loading are included in this paper. Three Quasi-Static cycles,
refereed as QS hereafter, were conducted at certain intervals to evaluate the stiffness
degradation; before fatigue testing (QS1), after 2 million fatigue cycles (QS2), and after 3
million fatigue cycles (QS3). In the first cycle of QS1, The beams were loaded until the
strain in tension steel reached a stress equivalent to 0.7fy and then unloaded until the strain
in the tension reached a stress equivalent to 0.42fy . The loads at which the beams reached
the predetermined lower and upper strain values were set as the maximum and minimum
fatigue limits for the other QS cycles and for 3 million fatigue loading cycles. The vehicle
load frequency on a typical RC bridge can be as high as 1.7Hz (Masoud et al., 2005) or
1.65Hz (Lin, 2006). Thus, fatigue loading was applied at a frequency of 2 Hz chosen to
simulate the in-service vehicle load frequency. It should be noted that the testing frequency

629

was adjusted several times during testing B0-00 in order to examine the effect of increased
frequency on stiffness degradation. The frequency was increased from 2.0 to 3.0 after 2.0
million cycles, and then increased from 3.0 to 4.0 after 2.5 million cycles; finally it was
reduced back to 2.0Hz for the remaining test cycles. Further discussion is presented in
section 3.2. After the completion of the cyclic loading; the beams were tested up to failure
in displacement-control mode. However, post-fatigue test results are not included in this
paper.
3. RESULTS AND DISCUSSION
3.1 Quasi-Static Cycles (QS)
Figure 2 presents the first static cycle of the load-deflection curves in series QS1 and QS3
of the three tested beams. As can be seen, B0-00 had the lowest initial stiffness followed by
B1-00 and B1-20, respectively. It is observed that the stiffness was enhanced upon
unloading the specimen after which it started to degrade with respect to the increase in
number of cycles. Table 1 presents the maximum load, load range, cracking load, stiffness,
and permanent deflection calculated for the first cycles of QS1 and QS3 series. B1-20
experienced the highest cracking load, as compared with B0-00 and B1-00, and initial
upward deflection of 0.49 mm due to prestressing. Stiffness shown in Table 1is calculated
as the average slope of the load deflection curves of the second and third cycles in QS1 and
QS3 while the permanent deflection is based on beams deflection upon unloading. B1-00
and B1-20 experienced the same stiffness degradation rate, 18%, even though they
exhibited different initial stiffness. Same trend was observed for the change in permanent
deflection values. Strengthening using non-prestressed and prestressed NSM-CFRP
reinforcement resulted in higher maximum load limits as compared with the control beam.
The increase in maximum load limit for B1-00, as compared with B0-00, is due to the
application of the CFRP reinforcement which resulted also in delaying the onset of steel
yielding. Furthermore, prestressing the NSM reinforcement, B1-20, increased the yielding
load hence, increased the upper fatigue load limit.

Fig. 2. Load-deflection curves of the first static cycle in QS1 and QS3.
630

Table 1. QS test results


Specim
en

B0-00
B1-00
B1-20

QS
QS1
QS3
QS1
QS3
QS1
QS3

Max
.
Loa
d
(kN)

Loa
d
rang
e
(kN)

44.1

26.7

55.6

32.1

62.5

31.1

Crackin
Stiffness
g load
(kN/mm)
(kN)
11.34
NA
14.06
NA
17.66
NA

4.667
3.917
5.159
4.219
5.223
4.277

Change
w.r.t.
QS1
(%)
-16.10
-18.21
-18.10

Change
w.r.t.
QS1
(%)

Permanent
Deflection
(mm)
5.44
8.66
5.30
8.52
5.01
8.16

59.2
60.6
63.1

3.2 Fatigue Loading


Table 2 presents the strains at mid-span of bottom steel, top fibre, CFRP strip, mid-span
deflection, and stiffness for the first and last cycle of the fatigue loading.
Table 2. Fatigue test results at the maximum load limit
Bottom steel
Top concrete
Deflection
strain
fibre strain
Cycle
@
Beam
@
@
(million) mid- Change
Change
Change
midmidspan
(%)
(%)
(%)
span
span
(mm)
1629
913
B0- 0.000001 14.00
38.21
59.55
34.69
00
3
19.35
2599
1230
1803
952
B1- 0.000001 16.42
28.01
19.19
18.15
00
3
21.02
2149
1125
1745
916
B1- 0.000001 17.03
35.82
13.35
96.29
20
3
23.13
1978
1798
Note: this is the change between first cycle and 3 million cycles

CFRP strip
strain
@
midspan
NA
NA
2722
2614
6138
6436

Change
(%)
NA
-3.96
4.85

Damage accumulation is manifested in the increase of deflection and stiffness degradation


with respect to the increase in number of cycles. The increase in mid-span deflection during
the fatigue loading, shown in Table 2, is attributed to the cyclic concrete creep phenomenon
and to the bond degradation between the tension reinforcement and the surrounding
concrete (Balaguru and Shah, 1982). Concrete cyclic creep strain depends on the concrete
mean cycling stress and on the number of cycles. Thus, the noticeable increase in deflection
during the early cycles is believed to be caused mainly by the rapid bond degradation
during the early life of the RC member. Figure 3 compares the increase of deflection at the
upper fatigue load, normalized to the initial deflection at the first QS1 cycle, of B0-00, B100, and B1-20. It was observed that the deflection increased dramatically at the initial
cycles, within the first 500 cycles, followed by a stabilized increase pattern. B0-00 and B1631

20 have almost the same normalized deflection increase while B1-00 experienced lower
normalized deflection at the end of the 3 million cycles. The lower normalized increased
deflection of B1-00 is attributed to fact that the stabilized deflection increase pattern was
achieved earlier than B0-00 and B1-20.
B0-00

1.40

B1-00

0.95

Normalized Stiffness

Normalized mid-span deflection

1.00

1.30

1.20
B0-00
B1-00
B1-20

1.10

B1-20

0.90
0.85
0.80
0.75
0.70

1.00
0.0

0.3

0.6

0.9

1.2

1.5

1.8

2.1

2.4

2.7

3.0

Number of cycles (million)

0.0

0.3

0.6

0.9

1.2

1.5

1.8

2.1

2.4

2.7

3.0

Number of cycles (million)

Fig. 3. Normalized mid-span deflection.

Fig. 4. Normalized stiffness degradation.

Figure 4 shows the variation in stiffness with respect to the increase in number of cycles. It
is shown that stiffness degradation is more apparent in the first 500 cycles after which no
major reduction in stiffness was experienced. However, stiffness decreased between 2.0-2.5
million cycles followed by more reduction between 2.5-2.7 million cycles. This reduction
in stiffness is due to the increased loading frequencies during these ranges of cycles. The
spikes observed in Figures 4 and 5 are attributed to the rest periods encountered before and
after each QS cycle, as well as to the scatter nature of data obtained from fatigue loading.
Under the application of fatigue loading, the mid-span strain in the CFRP strip for B1-00
decreased with respect to the increase in number of cycles indicating the occurrence of
bond degradation(Table 2) . However, the strain in the CFRP strips had increased for the
case of B1-20 indicating the efficiency of the prestressing system in enhancing the bond
performance. The mid-span CFRP strain variation, normalized to the strain value at first
cycle, is shown in Figure 5. It is observed that CFRP strain values of B1-00 and B1-20
decreased at initial cycling, within 500 cycles, after which they increased in stabilized
manner. The initial strain reduction indicates bond degradation at the level of NSM-CFRP
reinforcement for both strengthened beams. However, prestressing the CFRP strips resulted
in enhanced bonding properties presented in smaller reduction of CFRP strain at initial
cycles, as compared with B1-00, followed by stabilized increase pattern above unity.

632

1.10

1.01

B1-00
B1-20

Normalized mid-span CFRP strain

1.00
0.99
0.98

1.05

0.015

0.03

1.00

0.95

0.90

B1-00
B1-20

0.85
0.0

0.3

0.6

0.9

1.2

1.5

1.8

2.1

2.4

2.7

3.0

Number of cycles (million)

Fig. 5. Normalized mid-span CFRP strain.


Figure 6 shows the strain profile at first static cycle (which is supposed to be same as at the
first fatigue cycle, however because the data was not stabilized at the beginning of the
fatigue test due to the nature of the application of the sinusoidal cyclic loading), 1 million
cycle, 2 million cycle, and 3 million cycle along the length of the NSM CFRP strips of B100 and B1-20. For B1-00, the strain decreased at the center of the beam while increased at
other locations along the CFRP strips. This observation indicates that NSM debonding was
localized at the center of the beam since no signs of debonding were observed in other
locations. However, CFRP strain values at the center of B1-20 increased within the first 1
million cycles after which no major increase was observed.
The location of neutral axis increased by 2.8% and 2.2% at the end of fatigue loading, as
compared with first cycle, for B1-00 and B1-20, respectively. It should be noted that CFRP
strain in Figure 8(a) is the strain due to only the applied loading, calculated as the
prestressing strain subtracted from the total strain. The concrete top strain after 3 million
cycles increased by 18.1% and 96.3%, as compared with the first cycle for B1-00 and B120 respectively. The remarkable increase in concrete top fibre strain of B1-20, points to the
potential concrete fatigue failure.
8000

3200

Initial prestressing

2800

1 Million cycle

7000

1st Cycle

2400

2 Million cycle
3 Million cycle

6000

2 Million cycle

2000
1600
1200

1 Million cycle
3 Million cycle

5000
4000
3000

800

2000

400

1000

20%Prestress

Strain of FRP ()

Strain of CFRP strip ()

1st Cycle

0
0

1000

2000

3000

4000

5000

1000

2000

3000

Distance from support (mm)

Distance from support (mm)

(a) B1-00

(b) B1-20

Fig. 6. Strain profile along the length of the NSM CFRP strips.
633

4000

5000

TopFibre

1 Million cycle
3 Million cycle

-150

1st Cycle
1 Million Cycle

-100

2 Million cycle

Section depth (mm)

Section depth (mm)

TopSteel

-50

1st Cycle

-100

TopFibre

TopSteel

-50

-200
-250
-300

2 Million Cycle
3 Million Cycle

-150
-200
-250
-300

BottomSteel

-350

BottomSteel

-350
CFRP

-400
-1200 -800 -400

400

CFRP

-400
-2000 -1500 -1000 -500

800 1200 1600 2000 2400 2800

Strain ()

500 1000 1500 2000 2500 3000 3500

Strain ()

(a) B1-00
(b) B1-20
Fig. 7. Strain distribution across the depth of the beam at mid-span.
4. CONCLUSIONS
The performance of reinforced concrete beams strengthened with non-prestressed and
prestressed NSM CFRP strips subjected to fatigue loading was investigated in this research.
The following conclusions are drawn:

At maximum load limit, mid-span deflection increased by 28.0% and 35.8%, as


compared with the initial cycle, for B1-00 and B1-20, respectively. At the completion
of the fatigue loading, the deflection increase with respect to B0-00 was 8.6% and
19.5% for B1-00 and B1-20, respectively. .
Comparing the quasi-static performance of B1-00 and B1-20, both beams experienced
almost the same stiffness degradation rate, 18%, and increased permanent deflection,
62%, as compared with the first cycle.
FRP Stain at mid-span of B1-00 decreased by 4% at the completion of the fatigue
loading as compared with the initial cycle while increased by 5% for B1-20. This
indicates the effectiveness of prestressing in enhancing the bonding properties of RC
strengthened beams.
NSM strengthening using prestressed FRP reinforcement was found to enhance the
serviceability concerns, reduced deflection and enhanced stiffness, under cyclic loading
when compared to identical reinforced concrete beams.

5. ACKNOWELDEGEMENTS
The authors would like to express their gratitude to Lafarge Canada for providing the
concrete, Sika for providing the epoxy, and the technical staff at University of Calgary for
their assistance in the laboratory.
6. REFERENCES
1.

Cadei, J.M.C. 2003. Fatigue of FRP Composites in Civil Engineering Applications In


B. Harris (Ed.), Fatigue in Composites: Science and Technology of the Fatigue

634

Response of fibre-reinforced plastics. Abington Hall, Abington: Woodhead Publishing


Limited, pp. 658-685.
2. Quattlebaum, J.B., Harries, K.A., and Petrou, M.F. 2005. Comparison of Three
Flexural Retrofit Systems under Monotonic and Fatigue Loads. Journal of Bridge
Engineering, 10(6): 731-740.
3. Yost, J.R., Gross, S.P., and Deitch, M.J. 2007. Fatigue Behavior of Concrete Beams
Strengthened in Flexure with Near Surface Mounted CFRP. Proceedings of the 8th
International Symposium on Fiber Reinforced Polymer Reinforcement for Reinforced
Concrete Structures (FRPRCS8), University of Patras, Greece, 16-18 July, pp. 1-8.
4. Rosenboom, O., and Rizkalla, S. 2006. Behavior of Prestressed Concrete Strengthened
with Various CFRP Systems Subjected to Fatigue Loading. Journal of Composites for
Construction, 10(6): 492-502.
5. Badawi, M., and Soudki, K. 2009. Flexural Strengthening of RC Beams with
Prestressed NSM CFRP Rods - experimental and analytical investigation. Construction
and Building Materials, 23(10): 3292-3300.
6. Gaafar, M.A. 2007. Strengthening Reinforced Concrete Beams with Prestressed Near
Surface Mounted Fibre Reinforced Polymers. MSc Thesis, University of Calgary,
Calgary, Alberta, Canada, 205 p.
7. Canadian Standard Association (CSA). 2006. Canadian Highway Bridge Design Code
(CHBDC). CAN/CSA S6-06, Toronto.
8. Masoud, S., Soudki, K., and Topper, T. 2005. Postrepair Fatigue Performance of FRPRepaired Corroded RC Beams: Experimental and Analytical Investigation. Journal of
Composites for Construction, ASCE, 9(5): 441-449.
9. Lin, J.H. 2006. Response of a Bridge to a Moving Vehicle Load. Canadian Journal of
Civil Engineering, 33 (1): 49-57.
10. Balaguru, P., and Shah, S.P. 1982. A Method of Predicting Crack Widths and
Deflections for Fatigue Loading. SP 75-7, ACI Special Publications, Fatigue of
Concrete Structures, September 2, pp. 153-175.

635

636

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

CONCRETE DEMOLITION OF BRIDGE BARRIERS AND DECK


SLABS REINFORCED WITH GFRP BARS
E.A. Ahmed 1, C. Dulude 2, S. Goulet 3 and B. Benmokrane 4
1

Post-Doctoral Fellow, Dept. of Civil Engrg., University of Sherbrooke, Quebec Canada/ Lecturer,
Minoufiya University, Shebin El-Kom, Egypt.
2
Master Student, Dept. of Civil Engineering, University of Sherbrooke, Qc, Canada
3
Structures Division, Ministry of Transport of Quebec, Quebec City, QC, Canada
4
Canada & NSERC Research Chairs Professor, Dept. of Civil Engineering, University of
Sherbrooke, Quebec, Canada

ABSTRACT:
An extended investigation is being conducted at the University of Sherbrooke in
collaboration with the Ministry of Transportation of Quebec (MTQ) to evaluate the
performance of the bridge barriers types MTQ 210 and MTQ 311 reinforced with GFRP
bars under static and dynamic (impact) loads. This investigation extended to evaluate the
required time and damage that may occur to the GFRP and steel bars due to the demolition
of a concrete section of the barrier or the deck slab using the jackhammer technique. This
paper presents comparative demolition tests conducted on bridge barriers types MTQ 210
and MTQ 311 reinforced with steel and GFRP bars. The study included the demolition of
GFRP and steel-reinforced sections of the barriers or the curbs and the deck slabs using
jackhammer technique. Besides, the necessary time to complete the demolition as well as
the damage the steel and GFRP bars suffered during the demolition were evaluated and
compared. The comparative study revealed that the Jackhammer demolition technique can
be used for the GFRP-reinforced concrete barriers and deck slabs.
1.

INTRODUCTION

The anticipated service life of many steel-reinforced concrete structures is shortened due to
corrosion of steel reinforcement and related deteriorations problems. Reinforced concrete
bridge barriers and decks are among those structural elements that suffer from corrosion
and related deteriorations. This problem is exacerbated when large amounts of de-icing
salts are used on bridges during the winter season. Whereas there is a multitude of methods
to protect the steel reinforcement from corrosion, such as epoxy coating or galvanizing, the
use of non-corrodible fibre-reinforced polymer (FRP) materials can completely eliminate it.
The validity of the GFRP technology in reinforced concrete bridges has been demonstrated
through a number of field implementations conducted in North America (Benmokrane et al.
637

2004; 2006; 2007a; 2007b). In addition, recent investigation concerning the durability of
the GFRP-reinforced concrete (RC) structures showed that no degradation occurred in
GFRP RC structures under real field conditions for five to eight years in aggressive
environments, including exposure to wet-dry and freeze-thaw cycles, and de-icing salts or
salt water (Mufti et al. 2007). Furthermore, in Canada, the Canadian Highway Bridge
Design Code, CHBDC, Section 16 (CSA 2006) provides a step forward towards the
transition from research to commercial projects based on cost-benefit considerations.
However, there was an inquiry about the possibility of demolition and repair of GFRPreinforced sections that may face accidental damage. The issue of providing guidelines and
recommendations to repair, rehabilitate, or extend an existing concrete structure reinforced
with FRP internal reinforcement has been raised by officials from the Ministry of
Transportation of Quebec (MTQ), the Ministry of Transportation of Ontario (MTO) and
Manitoba Infrastructures & Transportation (MIT). From the owners perspective, every
bridge or overpass will require some kind of repair due to various kinds of damage such as
freeze-thaw damage, surface scaling, sulphate attack, alkali-aggregate reactions, vehicle
impact damage, substructure movement, non-functional bearings and deck joints, excessive
cracking, poor design details, poor quality construction, inadequate maintenance etc. They
are also concerned about the need for future structural modifications to accommodate
changes in level of service such as bridge widening or strengthening.
As the repair and/or rehabilitation of reinforced concrete elements often starts with
demolitions of concrete, previous studies were conducted to demolish the concrete of
GFRP-reinforced slabs using hydro-demolition and jack hammer techniques (Deitz et al.
2000; El-Salakawy et al. 2009). Deitz et al. (2000) reported that the hydro-demolition is
not a viable technique in the repair of bridge deck slabs reinforced with GFRP bars.
Besides, the water jet may remove the sand-coating and the surface resin layer which
impact the bond performance of the bars and the durability of the bare fibres due to the
alkali attack. On the other hand, El-Salakawy et al. (2009) concluded that the use of jack
hammers seems to be the most suitable and practical technique for the partial demolition of
concrete elements reinforced with GFRP bars. To further investigate this technique and
provide guidelines for repair of GFRP-reinforced concrete bridge deck slabs and barriers,
Ministry of Transportation of Ontario (MTO), Ministry of Transportation of Quebec
(MTQ), University of Sherbrooke, and Queens University established a task group for
preparing the Guidelines for Repair of Damaged Concrete Bridge Elements Reinforced
with FRP bars. Comparative demolition tests were conducted on the MTQ bridge barriers
reinforced with steel and GFRP bars using the jackhammer technique and the main
objectives of this investigation were:
To evaluate the necessary time to demolish a GFRP-reinforced section of the bridge
barriers Types MTQ 210 and MTQ 311 using jackhammer technique and to
compare it with that of a steel-reinforced one;
To evaluate the damage that the vertical GFRP reinforcement may suffer due to the
demolition of the concrete using jackhammer technique and to compare it with that
of steel bars.
This paper presents the demolition tests that were conducted on the MTQ 210 and MTQ
311 bridge barriers reinforced with steel and GFRP bars. The barriers used herein were
638

parts of an extended project that is being conducted at the University of Sherbrooke in


collaboration with the MTQ to evaluate the behaviour of bridge barrier types MTQ 210 and
MTQ 311 reinforced with GFRP bars under static and pendulum impact loads
(Benmokrane et al. 2009; Ahmed el al. 2011).
2.

TEST PROTOTYPES

Four bridge barriers types MTQ 210 and MTQ 311 were employed in the current study.
Two of them were totally reinforced with GFRP bars (210-GFRP and 311-GFRP) and the
other two were totally reinforced with steel (210-Steel and 311-Steel). The geometry and
reinforcement details of these prototypes are shown in Figures 1 to 3.
The horizontal deck slabs measured 11.0 m long 1.5 m wide 0.225 m deep. The curb of
the MTQ 210 prototypes measured 11 m long 0.45 m wide at the base and 0.415 m wide
at the top 0.28 m high over the slab surface. The barrier of the MTQ 311prototype
measured 11 m long 0.46 m wide at the base and 0.275 m wide at the top 0.88 m high
over the slab surface. The reinforcement details of both MTQ 210 and MTQ 311 prototypes
reinforced with steel and GFRP bars are shown in Figures 2 and 3, respectively. The
reinforcement of the steel and GFRP-reinforced barrier prototypes was provided by the
MTQ which is the most commonly used reinforcement for bridges. From Figures 2 and 3 it
can be noticed that the barrier walls and curbs are reinforced with the same amount of steel
and GFRP reinforcement (steel 20M @ 200 mm or GFRP No. 6 @ 200 mm for MTQ 311
and steel 15M @ 220 mm or GFRP No. 5 @ 220 mm for the MTQ 210 prototypes).
However, the reinforcement ratio of the slab of the GFRP-reinforced prototype was more
than that of the steel-reinforced prototype. The transverse bar spacing in the top mat was 75
mm and 120 mm for GFRP and steel-reinforced prototypes, respectively. More information
about the test prototypes can be found elsewhere (Ahmed et al. 2011). Those four barriers
were subjected to pendulum impact load (Ahmed et al. 2011) which caused a degree of
damage in the barriers (deformation and cracking). Thereafter, they were considered herein
for the current study.
C.L.
11000
2000

2000

Post

Post

2000

2000

Post

2000

500

Post

Post

Post

1500

845

1500

500

)
LE
O 0
(H 1 0

(a)
MTQ
210

1000
225

C.L.
11000
1200

1200

Post

2400

Post

1900

Post

1500

Post

845

1500

2400

)
LE
O 0
(H 10

(b)
MTQ
311

1900

1000
225

Fig. 1. Geometry of the MTQ 210 and MTQ 311 barrier prototypes: (a) 210; (b) 311

639

STEEL 15M (5)


Chamfer
15 x 15 (TYPE)
Concrete Cover 75 mm

450
415

35

STEEL 15M @ 220


TYPE G1

STEEL 15M @ 120


75

225

38 149 38

STEEL 15M @ 180

GFRP No. 15 (5)


Chamfer
15 x 15 (TYPE)
Concrete Cover 75 mm

450
415

35

GFRP No. 15 @ 220


TYPE G1 or G2

GFRP No. 20 @ 75
GFRP No. 20 @ 125
40

225

38 149 38

GFRP No. 20 @ 185

(a) 210-Steel
(b) 210-GFRP
Fig. 2. Reinforcement details of the MTQ 210 barrier prototypes

340
GFRP No. 15 (12)

225

38 149 38

75

STEEL 15M @ 180

GFRPNo. 15 @ 200
(TYPE D1)
GFRP No. 20 @ 75
GFRP No. 20 @ 125
40

GFRP No. 20 @ 185

225

STEEL 15M (10)

STEEL 15M @ 120

GFRP No. 20 @ 200 (TYPE D2)

38 149 38

340

STEEL 20M @ 200 (TYPE D2)

STEEL 15M @ 200


(TYPE D1)

GFRP No. 20 @ 200 (TYPE P1)

Concrete Cover 75

STEEL 20M @ 200 (TYPE P1)

Concrete Cover 75

(a) 311-Steel
(b) 311-GFRP
Fig. 3. Reinforcement details of the MTQ 311 barrier prototypes
3.

DEMOLITION TESTS

The demolition of both barrier types was conducted according to the MTQ activities (3131FRP & 3071-FRP (2010). The demolition was completed using pneumatic jackhammers.
For comparing the necessary time to complete the demolition, 1.0 m long of the 210 curbs
or 311 barrier walls was selected. However, for the slabs, an area of 1.50.55 m was
selected.
The equipments used in the demolition tests of the MTQ 210 and MTQ 311 barriers
included pneumatic jackhammers of 7 and 15 kg weight and a concrete saw. Firstly, a notch
of 20 mm deep was made in the concrete using the concrete saw at the boundaries of the
area that had to be demolished. Then a jackhammer of 15 kg weight was used to demolish
the concrete cover for the curb of MTQ 210, the barrier wall of MTQ 311 and the slabs.
Thereafter, a jackhammer of 7 kg weight was used to demolish the remaining of the

640

concrete elements and the areas close to the reinforcement to minimize the damage that
might happen to the reinforcing bars. The demolition tests had been conducted by workers
of a firm that has work experience in the demolition of steel-reinforced structural elements.
The workers were advised to be cautious during the demolishing of the concrete reinforced
with GFRP bars and to avoid the direct hit of the GFRP bars. Besides, they were advised
also not to try bending the GFRP bars during the demolition. Thus, the time to demolish
the GFRP prototypes was expected to be higher than that of steel reinforced one.
For the curbs and the barrier walls reinforced with GFRP bars, the longitudinal bars were
cut to facilitate the demolition. Those bars have to be replaced during the repair of those
barriers using epoxy based adhesives. However, the repair was not a part of this study.
Figures 4 to 6 show the demolition of the MTQ 210 curbs. While Figures 7 to 9 show the
demolition of the MTQ 311 barrier walls.

DEMOLITION

SAW CUT 20 mm (TYP.)

DEMOLITION
STIRRUPS KEPT

LONG. REBAR CUT


AND TO REPLACE (5)
1000

Fig. 4. Schematic for the demolition test of the MTQ 210 barriers

Fig. 5. Demolition of the concrete curb of the MTQ 210 barriers

641

(a) 210-GFRP
(b) 210-Steel
Fig. 6. MTQ 210 barriers after demolition of concrete
DEMOLITION

LONG. REBAR CUT


AND TO REPLACE (12)

VERTICAL REINF.
KEPT

1000

Fig. 7. Schematic for the demolition test of the MTQ 311 barriers

Fig. 8. Demolition of the concrete barrier wall of the MTQ 311 barriers

(b) 311-GFRP
(b) 311-Steel
Fig. 9. MTQ 311 barriers after demolition of concrete
642

On the other hand, the demolition of the steel and GFRP-reinforced slabs is shown in
Figures 10 to 12. In the slabs of both barriers reinforced with steel and GFRP bars, the
concrete was removed to the extent of 25 mm under the top mat.
The necessary time to complete the demolition of each element (curb, barrier, and slab) in
both cases of steel and GFRP reinforcement was evaluated and listed in Table 1. From this
table, it can be noticed that removing the horizontal reinforcement of the 210-GFRP
reduced the required time to complete the demolition by 40%. The barrier wall 311-GFRP
needed 36% more time than that of barrier wall 311-Steel. Besides, the longer time for the
demolition of the GFRP-reinforced slabs than that of steel-reinforced one was due to the
closer reinforcement spacing (75 mm in comparison with 120 mm).

Fig. 10. Demolition of a slab reinforced with GFRP bars

Fig. 11. GFRP reinforcing bars of the slab after concrete demolition

Fig. 12. Steel reinforcing bars of the slab after concrete demolition

643

Table 1. Time elapsed to complete the concrete demolition of each element


311-GFRP 210-GFRP GFRP-Slab 311-Steel 210-Steel
Time
225 min
105 min
225 min
165 min
185 min

Steel-Slab
175 min

After the demolition of curbs, barrier walls, and slabs reinforced with GFRP and steel bars,
the damage of the steel and GFRP bars due to the pneumatic jackhammer was evaluated.
The degree of damage was categorized into three categories: Light (noted as 1), Medium
(noted as 2), and Severe (noted as 3). Light damage indicated that the coating of the bar and
very small layer of the bar surface was affected. Medium damage indicated that the bar was
affected to some extent which was about 20 to 25% of the bar diameter. Severe damage
indicated that the cross-section of the bar was greatly reduced and the effect was too deep
into the bar section. The steel bars, however, did not show any signs of severe damage.
Tables 2 and 3 show the photos of the classified damage and the corresponding damage
category for the GFRP and steel bars, respectively. It is worth mentioning that the direct hit
by the jack hammer was the reason for the GFRP bar damage. However, removing the
surrounding concrete without hitting the GFRP bars directly using the jack hammer did not
yield any damage.
Table 2. Classified damage of GFRP bars
Damage Degree

Photos

Light damage of
GFRP straight bars
(Noted as 1)
Light damage
GFRP bent bars
(Noted as 1)

of

Medium damage of
GFRP straight bars
(Noted as 2)
Medium damage of
GFRP bent bars
(Noted as 2)
Severe damage of
GFRP straight bars
(Noted as 3)

644

Damage Degree

Table 3. Classified damage of steel bars


Photos

Light damage steel


bars
(Noted as 1)
Medium
damage
steel bars
(Noted as 2)

According to this classification, the 210 curbs did not show any significant damage for steel
or GFRP stirrups. This was refereed to the large spacing between the stirrups which was
220 mm. For the vertical bars of the 311 barrier walls, it was noticed that few damage
locations were recorded for steel and GFRP bars. The 311-Steel barrier showed 5 locations
classified as light damage whereas the 311-GFRP showed only two locations with one
classified as light and the other one classified as medium. Figure 13 shows the classified
damage for the vertical bars of both of 311-Steel and 311-GFRP barrier walls. On the other
hand, a lot of damage points were observed in the horizontal reinforcement of the steelreinforced barrier wall (311-Steel) as shows in Figure 14. The degrees of damage at these
locations were classified as light or medium. However, as mentioned earlier, the horizontal
GFRP bars of the 311-GFRP barrier wall were cut and removed during the demolition as it
was decided in the demolition procedure.
The higher degree of damage that occurred to the longitudinal steel reinforcement of the
311-Steel barrier wall was referred to the location of these bars perpendicular to the
demolition direction which was parallel to the vertical bars. Thus, the horizontal bars were
more susceptible to the direct hit from the pneumatic jackhammer during the demolition.
The damage of the steel and GFRP bars in slabs was also evaluated and marked similarly as
shown in Figures 15 and 16. The damage occurred to the transverse steel and GFRP bars
was similar except a few points with severe damage occurred to the GFRP bars. This is
refferred to the closer reinforcement in case of GFRP bars than that in case of steel bars (75
mm compared to 120 mm). Besides, the longituidal GFRP bars showed higher damage due
to their location perpendicular to the demoltion direction.

645

INTACT CONCRETE

LONG. BARS CUT

INTACT CONCRETE

INTACT CONCRETE

LONG. BARS CUT

INTACT CONCRETE

(a) 311-Steel

(b) 311-GFRP

Fig. 13. Classified damage to the D1, D2 and P1 bars of the 311 barriers (1, 2 and 3
corresponding to light, medium and severe damage, respectively)
INTACT CONCRETE

INTACT CONCRETE

Fig. 14. Classified damage to the longitudinal steel bars of the 311 barriers

Fig. 15. Classified damage for the steel bars of the steel-reinforced slab (Longitudinal and
transverse bars)
646

(Transverse bars)
(Longitudinal bars)
Fig. 16. Classified damage for the GFRP bars of the GFRP-reinforced slab
4.

CONCLUSIONS

Based on the results and the discussion presented above, the following conclusions can be
drawn:
Generally, the GFRP-reinforced barriers and slabs can be demolished using the
pneumatic jackhammer technique similar to the steel reinforced ones. With the same
reinforcing bar spacing, the damage that GFRP bars suffered was very close to that
steel bars showed except at few locations.
The direct hit of the GFRP bars by the jack hammer was the main reason for the
damage that the GFRP bars showed. The demolition of the surrounding concrete
without hitting the GFRP bars would not yield such damage.
The GFRP stirrups in the 210 barriers have suffered slight damage following the
demolition of the concrete. Cutting the longitudinal reinforcement contributed to
reducing the time of demolition to approximately 60% compared to that of the steel
reinforced one.
The vertical bars (D1, D2 and P1) of the 311-GFRP barrier wall suffered slight
damage due to the demolition of the concrete. The time required for the demolition of
the GFRP-reinforced barrier wall was longer than that of the counterpart reinforced
with steel (36% more).
The GFRP-reinforced slab suffered more damage than the steel reinforced one. The
closer spacing of transverse reinforcement made the demolition of the concrete more
difficult compared to the barrier walls or curbs. The time required for the demolition
of the GFRP-reinforced slab was 1.30 times that of the steel-reinforced one.
The Jackhammers of 15 kg weight may be used without problems to demolish the
concrete cover. Moreover, when the spacing between the reinforcing bars allows, the
same hammer could be used to demolish the concrete between the bars. However, the
jackhammer of 7 kg should be used to demolish the concrete close to GFRP bars.
The results of the demolition tests conducted herein contributed to the preparation of
the first draft of Recommended Best Practice for Repair of Damaged Bridge Barrier
Walls, Curbs, and Slabs Reinforced with FRP Bars by the MTO/MTQ Task group
Guidelines for Repair of Damaged Concrete Bridge Elements Reinforced with FRP
bars.

647

5.

ACKNOWLEDGEMENTS

The authors acknowledge the financial support received from the Ministry of Tranportation
of Quebec. The authors also acknowledge Les Coffrages Carmel Inc. (Sherbrooke,
Quebec), the demolition contractor.
6.

REFERENCES

7.

Ahmed, E., Dulude, C., and Benmokrane, B. 2011. Static and Dynamic Testing of 210
and 311 Barriers Reinforced with GFRP Bars. Final Technical Report Submitted to the
Ministry of Transportation of Quebec (MTQ), February, 129 p.
Benmokrane, B., Dulude, C., Ahmed, E., and El-Gamal, S. 2009. Static and Dynamic
Testing of 210 and 311 Bridge Barriers Reinforced with GFRP Bars. Progress Report
No. 1 (Static Testing), Submitted to the Ministry of Transportation of Quebec (MTQ),
Structure Division, September, 120 p. (in French).
Benmokrane, B., El-Salakawy, E., Desgagn, G., and Lackey, T. 2004. FRP Bars for
Bridges. Concrete International, 26(8): 8490.
Benmokrane, B., El-Salakawy, E., El-Gamal, S., and Goulet, S. 2007a. Construction
and Testing of an Innovative Concrete Bridge Deck Totally Reinforced with Glass FRP
Bars: Val-Alain Bridge on Highway 20 East. Journal of Bridge Engineering, 12(5):
632645.
Benmokrane, B., El-Salakawy, E., El-Ragaby, A., and El-Gamal, S. 2007b.
Performance Evaluation of Innovative Concrete Bridge Deck Slabs Reinforced with
Fibre-Reinforced Polymer Bars. Canadian Journal of Civil Engineering, 34(3): 298
310.
Benmokrane, B., El-Salakawy, E., El-Ragaby, A., and Lackey, T. 2006. Designing and
Testing of Concrete Bridge Decks Reinforced with Glass FRP Bars. Journal of Bridge
Engineering, 11(2): 217-229.
Canadian Standard Association (CSA). 2006. Canadian Highway Bridge Design Code.
CAN/CSA S6-06, Rexdale, Ontario, Canada, 788 p.
Deitz, D., Harik, I.E., and Gesund, H. 2000. GFRP Reinforced Concrete Bridge Decks.
Research Report KTC-00-9, Kentucky Transportation Center, College of Engineering,
University of Kentucky, Kentucky, USA, 185 p.
El-Salakawy, E., Mufti, A., El-Ragaby, A., Fllis, G., Saltzberg, W. 2009. Repair of
Concrete Structures Reinforced with GFRP Bars. Proceedings of the 9th Fibre
Reinforced Polymers in Reinforced Concrete Structures conference (FRPRCS-9),
Sydney, Australia, 13-15 July, 4 p.
Ministre des Transports du Qubec. 2010. Rparation de glissire ou de chasse-roue
en bton arm de matriaux composites. Activit PRF-3071, Fvrier, 10 p.
Ministre des Transports du Qubec. 2010. Rparation de dalle sur poutres avec
armature en matriaux composites. Activit 3131-PRF, Fvrier, 13 p.
Mufti, A., Banthia, N., Benmokrane, B., Boulfiza, M., and Newhook, J. 2007.
Durability of GFRP Composite Rods. Concrete International, 29(2): 3742.

8.

9.
10.

11.

12.
13.
14.
15.

16.
17.
18.

648

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

EFFECT OF FREEZE/THAW CYCLES ON THE COMPRESSION


STRENGTH OF CONCRETE-FILLED-FRP TUBE CYLINDERS
H. El-Zefzafy1, H.M. Mohamed2 and R. Masmoudi3
1

Ph.D. Candidate, University of Sherbrooke, Sherbrooke, QC, Canada


NSERC-Postdoctoral Fellow, University of Sherbrooke, Sherbrooke, QC, Canada
3
Professor , University of Sherbrooke, Sherbrooke, QC, Canada
2

ABSTRACT
Previous studies have demonstrated the high performance of the concrete-filled fiberreinforced polymer (FRP) tubes (CFFT) technique as a stay-in-place formwork and
confining system. However, the environmental effects such as freeze-thaw cycles and deicing salt solutions may affect materials properties, and therefore the structural response of
this technique. This paper investigates experimentally the durability short and long term
behaviour of 24 CFFT and 24 plain concrete cylinders (152 305 mm). Test variables
include confined effect of using glass-FRP tubes and effects of different freeze-thaw
exposure. Normal strength CFFT cylinders were exposed to 100 and 300 freeze-thaw
cycles both while saturation in (salt and/or fresh) water and air. Then, pure axial
compression tests were conducted to evaluate the change of mechanical properties of the
test specimens. The test results revealed that FRP tubes successfully protected the concrete
core of the test specimens against the environmental condition. The short-term freeze-thaw
cycling had almost no effect on the compressive strength of the CFFT cylinders.
1. INTRODUCTION
The application of composite materials has been propagated by the deterioration of the
existed conventional concrete, steel, and timber structures. The glass fiber-reinforced
polymers tubes can play an important role in replacing transverse steel by providing
ductility and strength for conventional columns. The use of FRP tubes offers several
advantages confinement, protective jackets, providing shear or/and flexural reinforcement
and permanent formwork for the concrete structures. Although a number of studies have
been conducted on the durability performance of FRP-wrapped concrete cylinders, there is
currently a lack of understanding the behaviour of CFTT under the same conditions. Both
earlier and more recent studies have revealed that the freeze-thaw cycles can significantly
reduce the effects of confinement due to materials degradation. Karbhari (1994) reported a
reduction of more than 30% in the ductility of GFRP-confined concrete as an effect of
freeze-thaw cycles combined with moisture. Belarbi and Bae (2006) conducted uni-axial
compressive failure tests on RC columns strengthened with CFRP and GFRP sheets. The
649

study aimed to evaluate the change of mechanical properties due to effects of a combined
environmental cycle (freezethaw cycles, high-temperature cycles, high-humidity cycles,
saline solutions, and ultraviolet (UV) radiations). The saline solution was found to decrease
significantly the failure load and ductility of the GFRP test specimens, whereas CFRPcolumns exhibited slight decrease in failure load. Touanji et al. (2007) highlighted their
observation that the concrete confined with FRP and PVC-FRP hybrid composite
significantly protected the concrete when subjected to harsh environmental condition
(freeze-thaw and wet-dry). Micelle and Myers (2008) conducted experimental axial tests on
exposed GFRP-confined columns to environmental cycles condition (freeze-thaw, moisture
and high- temperature cycles) and immersion in NaCl solution. The test results indicated
that more than 40% loss in the ductility and moderate decrease in the ultimate strength were
observed. Also cylinders immersed in NaCl solution lost about 30% of their axial strain. A
recent study by El-Hasha et al. (2010) investigated the axial behaviour of thirty-six CFRPconfined concrete cylinders subjected to freeze-thaw cycles after exposure to heating and
cooling cycles (23 to 45C). The test results indicated that no significant difference was
found in strength between the wrapped cylinders which were subjected to heating and
cooling and the specimens that were kept at room temperature. In addition, freezingthawing exposure as well as fresh and salt water immersion had a slightly negative effect
on the strength of both, unwrapped and wrapped cylinders.
This paper presents the axial capacities of CFFT cylinders exposed to short and long term
freeze-thaw cycles. The freeze-thaw exposure was conducted in three sets, two of them
were meanwhile submerged in (salt and/or fresh) water and the other set was in air. This
study is focused only on the behaviour of the unreinforced CFFT specimens.
2. EXPERIMENTAL WORK
2.1

Materials

2.1.1

GFRP Tubes

FRP tubes with an internal diameter of 152 mm were used in the experimental program.
The FRP tubes were fabricated using the filament winding technique; E-glass fiber and
Epoxy resin were utilized for manufacturing these tubes. The split-disk and coupon tensile
tests were performed on five specimens according to the ASTM D-2290-08 and ASTM D
638-08 standard to determine the strength and Youngs modulus in the axial and transverse
directions. Table 1 shows the details, dimensions and the mechanical properties the used
FRP tubes.
Table 1. Dimension and mechanical properties of FRP tubes
Hoop
Axial
Diameter Thickness Stacking Youngs Ultimate
Youngs Ultimate
Ultimate
Ultimate
(mm)
(mm) sequence modulus strength
modulus strength
Strain
Strain
(MPa) (MPa)
(MPa)
(MPa)
152
2.65
[60]3
10385
348
0.0388 12808
60.10 0.009328

650

2.1.2

Concrete

All specimens were constructed from the same batch of normal strength ready mix
concrete. Water reducing admixture with super plasticizer was used to increase the
workability of the concrete mix. Ten plain concrete cylinders (152305 mm) were tested at
28-days under axial load. The average concrete compressive strength for ten cylinders was
found 330.4 MPa.
2.2 Specimens Details and Preparation
Twenty-four (152305 mm) plain concrete and 24 CFFT cylinders were prepared and
tested in this study, see Figure 1. Half of the specimens were exposed to 100 freeze-thaw
(F/T) cycles and the reset were exposed to 300 freeze-thaw (F/T) cycles. The specimens
were divided into four groups; each group consists of 6 specimens: 6 CFFT cylinders and 6
control plain concrete cylinders (PCC). The first group was used as reference and kept at
room temperature for a period equivalent to the 100 and 300 freeze-thaw (F/T) cycles. The
other groups were exposed to 100 and 300 freeze/thaw cycles inside an environmental
chamber under three different conditions. Different 3 CFFT and 3 PCC specimens were
kept in dry air, submerged in fresh and salt water inside the chamber. The specimens
submerged in salt water were to simulate the environment of de-icing conditions of the
infrastructures. In this study the ends of conditioned cylinders were not exposed by
protecting them with epoxy. This is attributed to the fact that, in field applications, columns
ends are connected in the superstructure. The objective of doing so was to permit the
diffusion of the saturated F/T exposure only permitted through the FRP tube. In addition,
some specimens were cast with thermocouples, which help determine the appropriate
thermal cycle to be used for controlling the environmental chamber.

a. Formwork and FRP


b. casting specimens
tubes
Fig. 1. Fabrication and casting the control and CFFT cylinders
2.3 Freeze-Thaw Exposure
The dry freeze-thaw specimens were placed in the freeze-thaw chambers. Whereas, those
for saturated freeze-thaw were left in fresh water bath and placed in the same
environmental chamber using the isolated wooden tanks. According to the different types
651

of environmental conditions (freezethaw cycles in dry, fresh water) existed inside the
same environmental chamber, it was hard to reach the same temperature inside the
saturated and non-saturated specimens. Therefore, a temperature acquisition system was
used to monitor the temperature during cycling (see Figure 2). The freezing cycles
consisted of lowering the temperature in the middle of the saturated specimens from 4.4 to 17.8C in a period of 16.5 h. The thawing cycles consisted of raising the temperature in the
middle of the saturated specimens from -17.8 to 4.4C. Those freeze-thaw hours were
sufficient to vary the temperature of non-saturated specimens between (+28C to -28C) as
shown in Figure 3. Thus, the specimens underwent one freeze-thaw cycles per 27 hours,
rather than in accelerated shorter cycles. This procedure which followed to practice this
type of exposure was considered to simulate the winter effect.

a. Specimen in wooden tanks inside the


b Monitoring the chamber during F/T
chamber
cycles
Fig. 2. Fabrication and conditioning the test specimens
2.4 Test Procedure for Compression Tests
All specimens were brought to room temperature before being tested. Uniaxial compression
tests were conducted until failure. Before testing, two axial and two hoop electrical strain
gauges were mounted at the mid height, 180 degree apart, along the hoop direction on the
external surface of the specimens. Strain gauges of 6 mm length were used to monitor the
strain distribution of the GFRP tubes. 30 mm strain gauges were bonded on the surface of
the plain cylinders. The axial displacement for each cylinder was measured by two linear
variable displacement transducers (LVDTs) 180 degrees apart along the hoop direction of
the specimen. All specimens were prepared before the test by a thin layer of the high
strength sulfur capping on the top and bottom surfaces to insure the uniform stress
distribution during the test. The specimens were tested using a 6,000 kN capacity FORNEY
machine, where the CFFT cylinder were setup vertically at the center of loading plates of
the machine. The FORNEY machine, strain gauges and LVDTs were connected by a 20
channels data acquisition system and the data were recorded every second during the test.
The loading rate range was 2.0 to 2.50 kN/s during the test by manually controlling the

652

loading rate of the hydraulic pump. Figure 4 illustrates the experimental test setup used in
this study.

Fig. 3. Temperature variation inside the specimens during two complete F/T cycles

(b) Overview of test setup


(a) Instrumentations
Fig. 4. The instrumentations and test setup
3. RESULT AND DISSECTIONS
3.1 Effect of Confinement: Virgin Specimens
Table 2 presents the test results of PCC kept in room temperature and environmental
chamber under different conditions. Table 3 showed the axial stress, axial and hoop strain
values of the test specimens. Confining using the GFRP tubes showed a large increase in
the compressive strength, axial and hoop strain in comparison with unconfined specimens,
as a result of the confinement in the radial direction. The average ratio of confined to
unconfined concrete strength (fcc/fc) was 2.2. The average confined strength of the
unconditioned CFFT cylinders was 71.6 and 71.5 MPa for specimens tested at time of 100
and 300 F/T cycles, respectively, and , representing 33% increases over the unconfined
strength of plain concrete cylinders. The maximum confined concrete strength for confined
specimen using a GFRP tube reached up to 75 MPa. Table 3 confirms the fact that
improvements in ductility was achieved, as measured by the increase in level of axial and
653

hoop failure strain, on average 0.03 mm/mm as compared to that of PCC specimens 0.003
mm/mm.
3.2 Effects of Freeze- Thaw Cycles
It is important that we review results here to explain the effects of freezing and thawing on
the confined concrete. During the freeze cycle water in the pores turns into ice and
increases in volume by approximately 9%. Thus, stress is induced on the surrounding
concrete. Similar pressure will be generated if water is present in spaces between the FRP
tubes and the confined concrete core. However, during the thaw cycles this induced stress is
released. If the stress induced during the freeze cycles is higher than the tensile strength of
the concrete, with increasing the freeze thaw cycles, damage such as cracks in the concrete
occurs. Thus, as long as the F/T cycles are repeated more damage will happen due to water
penetration into the cracks. The unconditioned control PCC specimens were tested after the
total period 0f the 100 and 300 cycles and it gave an average strength of 33.4 and 32.7
MPa, respectively. In comparison with the unconfined conditioned control cylinders, a
reduction in (fc) by 10.3%, 7% and 28% of the unconfined strength after exposure to 100
F/T cycles in air, fresh water and salt water, respectively, were observed. The
corresponding values after exposure of 300 F/T cycles were 13.5, 10 and 36.4%.
Table 2. Test results of PCC exposed to 100 and 300 freeze-thaw cycles
Condition

100 Freeze-Thaw Cycles

300 Freeze-Thaw Cycles

Room

(fc) Average
strength* (MPa)
33.4

(fc) Reduction after


exposure (%)
--

(fc) Average
strength* (MPa)
32.7

(fc) Reduction after


exposure (%)
--

F/T Dry

30.2

-10.3

28.3

-13.5

-7.0

29..5

-10

-28

20.8

-36.4

F/T Fresh
31.4
water
F/T Salt
24.6
water
* Test results for three cylinders

It should be noted at this point that the reduction occurred although air entrainment
concrete was used in this study to reduce the freeze-thaw effect as much as possible. In
comparison with the reference CFFT specimens kept at room temperature, the CFFT
specimens tested after exposure to the 100 dry F/T cycles showed insignificant increase in
the average compressive strength. The CFFT specimens conditioned to 300 dray F/T cycles
presented a reduction 1.6%. On the other hand, the CFFT cylinders exposed to saturated
100 and 300 F/T cycles in fresh water exhibited a slight decrease in the average
compressive strength by 1.4 and 1.17%, respectively, as compared with the CFFT
specimens kept in room temperature. The corresponding values for specimens conditioned
in salt water were, respectively, 2.65 and 4.8%. The test results indicated that the combined
environmental cycles such as freeze-thaw cycles in fresh water or dray air did not show a
significant effect on the CFFT specimens. This was probably because the adverse effect,
such as plasticization of matrix and micro-cracking at matrix-fibre interface, induced by the
low temperature during the freeze-thaw cycles, compromised the positive effect, such as
matrix hardening effect. On the other hand, the freeze-thaw cycles affect the compressive
654

behaviour of PCC and RC specimens. The possible reason for the decrease could be
attributed in large part to the cracks, spalling, damage and deterioration of the concrete
cover.
Table 3. Test results of CFFT specimens exposed to 100 and 300 freeze-thaw cycles
100 Freeze-Thaw Cycles
Axial strain*
Hoop strain*
(mm/mm)
(mm/mm)
0.041
0.025
1
75
0.032
0.030
0.031
0.032
Room
2
73
0.044
0.020
0.029
3
70
0.035
0.022
Average
72.66
0.038
0.026
SD
2.51
0.004
0.004
0.026
0.034
1
77
0.026
0.022
0.023
0.034
F/T Dry
2
73
0.017
0.029
0.035
0.066
3
70
0.055
0.012
Average
73.33
0.030
0.032
SD
3.51
0.013
0.018
0.049
0.050
1
72
0.032
0.040
0.029
0.030
F/T Fresh
2
74
0.026
0.037
water
0.034
0.026
3
76
0.029
0.035
Average
71.6
0.033
0.036
SD
1.15
0.008
0.008
0.023
0.025
1
66
0.031
0.035
0.022
0.025
F/T Salt
2
71
0.032
0.042
water
0.027
0.030
3
72
0.037
0.032
Average
69.61
0.029
0.031
SD
3.22
0.006
0.007
* strain readings for two electrical strain gauges

Condition

Specimen
No.

Stress
(MPa)

300 Freeze-Thaw Cycles


Stress
Axial
hoop
(MPa)
strain
strain
0.032
0.038
73
0.022
0.033
0.033
0.037
71
0.031
0.027
0.028
0.031
71
0.024
0.024
71.5
0.028
0.031
1.15
0.004
0.005
0.016
0.013
72
0.017
0.020
0.039
0.023
69
0.032
0.027
0.021
0.034
70
0.034
0.021
70.33
0.026
0.023
1.52
0.009
0.007
0.008
0.010
71
0.008
0.011
0.184
0.016
75
0.007
0.001
0.022
0.05
74
0.010
0.029
70.66
0.039
0.019
2.1
0.007
0.017
0.028
0.028
67.67
0.023
0.018
0.033
0.027
70.55
0.029
0.026
0.030
0.021
65.48
0.018
0.019
68
0.026
0.023
2.54
0.005
0.004

3.3 Failure Modes


From carful observation during the freeze-thaw cycles, it should be noted that the
unconfined cylinders exposed to freeze/thaw conditions started to showed visual signs of
deterioration in the concrete. Several cracks were observed after 40 F/T cycles. By the end
of the 300 cycles prior to testing, all the specimens were visually inspected for evidence of
degradation in of micro-cracks. More signs of deterioration in the saturated plain concrete
cylinders after F/T exposure were reported, rather than specimens exposed to F/T cycles in
air. Figure 5.a shows the damage and deterioration of the concrete cover of PCC. In all
cases, careful observation showed rupture of the fibre in the hoop direction at the ultimate
655

hoop stress, resulting from the dilation of the concrete. The fracture of the GFRP tubes
occurred within the total height of the cylinders started from top or bottom and extended to
the opposite direction. The rupture of the tube was very clear for the specimens exposed to
F/T in salt water. The shape of the failure was a zigzag pattern normal to the direction of
fibres in the hoop direction. Popping sounds heard during the early-to-middle stages of
loading were referred to the dilation, micro-cracking of concrete and offset of the aggregate
and the ultimate failure was very explosive due to rupture of the fibres. The room
temperature specimens failed in less destroyed behaviour between the four conditions.

a. PCC
b. dry air
c. fresh water
Fig. 5. Different failure modes of conditioned RC and CFFT specimens
4. SUMMARY AND CONCLUSIONS
The study presented herein intended to examine experimentally the short and long term
effects of freeze-thaw cycling on the compressive behaviour of concrete-filled GFRP tubes
cylinders. In summary, it was found that the confinement using a GFRP tube appear to
provide excellent protection against freeze-thaw cycles. Based on the test results so far, the
following conclusions are drawn:
Regardless of the type of the freeze-thaw cycles, the 100 freeze-thaw cycles have almost
no affect on the average ultimate stress. However, insignificant decrease (2.6 and 2.7%)
in the confined strength of the specimens exposed to freeze-thaw in salt and fresh water,
respectively, was observed.
The CFFT specimens exposed to long term freeze-thaw cycles (300 cycles) showed a
reduction on average 8.6% in the axial capacity.
The PCC exposed to 300 freeze-thaw cycles in dray air, fresh water and salt water
showed a significant reduction in the axial compressive capacity, on average 13.5, 10
and 36.4%, respectively.
Confinement using GFRP tubes was found to enhance the compressive strength to 2.15
times that of plain concrete. Also, a significant increase in ductility, in term of average
axial and hoop strains, was achieved.
5. ACKNOWLEDGEMENTS
The research reported in this paper was partially sponsored by the Natural Sciences and
Engineering Research Council of Canada (NSERC). The authors also acknowledge the
contribution of the Canadian Foundation for Innovation (CFI) for the infrastructure used to
conduct testing. Special thanks to the FRE Composites Inc, QC, Canada, for providing the
FRP tubes.
656

6. REFERENCES
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.

ASTM. 2008a. Standard test method for apparent hoop tensile strength of plastic or
reinforced plastic pipe by split disk method. D 2290-08, West Conshohocken, Pa.
ASTM. 2008b. Standard test method for tensile properties of plastics. D638-08, West
Conshohocken, Pa.
Belarbi, A., Bae, S.W. 2007. An experimental study on the effect of environmental
exposure and corrosion on RC columns with FRP composite jackets. Composites: Part
B, 38: 674684.
Callery, K., Green, M.F., and Archibald, J.F. 2000. Environmental effects on the
behavior of wrapped concrete cylinders. Proceedings of the 3rd international conference
on Advanced Composite Materials in Bridges and Structure, Canada, pp. 759-766.
EL-Hacha, R., Green, M.F., and Wight, G.R. 2010. Effect of severe environmental
exposures on CFRP wrapped concrete columns. Journal of Composites for
Construction, ASCE, 14(1): 83-93.
Fam, Z.A., and Rizkalla, S. 2002. Flexural behavior of concrete-filled fiber reinforced
polymer circular tubes. Journal of Composites for Construction, ASCE, 6(2): 123132.
Micelli, F. and Myers, J.J. 2008. Durability of FRP confined concrete. Construction
Materials, 6(4).
Mirmiran, A. and Shahawy, M. 1997. Behavior of concrete columns confined by fiber
composites. Journal of Structural Engineering, 123(5): 583590.
Mohamed, H., and Masmoudi, R. 2010. Axial load capacity of reinforced concrete-filled
FRP tubes columns: experimental versus theoretical predictions. Journal of Composites
for Construction, ASCE, 14(2): 231-243.
Mohamed, H., and Masmoudi, R. 2008. Compressive behaviour of reinforced concrete
filled FRP tube. ACI Special Publications (SP), SP-257-6, pp. 91-109.
Saenz1, N., and Pantelides, C.P. 2006. Short and medium term durability evaluation of
FRP-confined circular concrete. Journal of Composites for Construction, 10(3): 242253.
Teng, M., Sotelino, E.D. and Chen, W. 2003. Performance evaluation of reinforced
concrete bridge columns wrapped with fiber reinforced polymer. Journal of
Composites for Construction, ASCE, 7(2): 83-92.
Toutanji, H., Zhao, L. and Isaacs, G. 2007. Durability studies on concrete columns
confined with advanced fiber composites. International Journal of materials and
product technology, 28: 8-28.

657

658

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

EFFECTS OF HARSH LABORATORY AND FIELD


ENVIRONMENTAL CONDITIONS ON THE DURABILITY OF
CONCRETE WRAPPED GFRP BARS
T.H. Almusallam1, Y.A. Al-Salloum1, S. Alsayed1, S.E. El-Gamal2 and M. Aqel3
1

Professor, Dept. of Civil Eng., King Saud University, P.O. Box 800, Riyadh 11421, Saudi Arabia
Assistant Prof., Dept. of Civil Eng., King Saud Univ., P.O. Box 800, Riyadh 11421, Saudi Arabia
3
Research Assistant, Dept. of Civil Eng., King Saud Univ., Riyadh 11421, Saudi Arabia
2

ABSTRACT
This paper presents the test results of an experimental study to investigate the durability of
newly developed Glass Fiber Reinforced Polymer (GFRP) bars. A total of 90 GFRP bars
were embedded in concrete prisms (40 40 450 mm) and exposed to nine environmental
conditions for different periods (6, 12, and 18 months). The environments included
exposure to sea water, ordinary tap water and alkaline solution at two temperatures (room
and 50C). The environments also included two typical field conditions of the Kingdom of
Saudi Arabia (Gulf area and Riyadh area). The performance of the GFRP bars has been
assessed by conducting tensile strength test on the bars extracted out of the concrete
specimens after exposure to the environmental conditions. After 18 months of exposure, the
test results showed that the tap water at 50C followed by alkaline solution at 50C had the
maximum effect on the tensile strength of the GFRP bars. The two field conditions did not
show any effect on the tensile properties of the tested bars.
1. INTRODUCTION
In addition to their non-corrodible nature, FRP bars exhibit several properties, such as high
tensile strength, that make them suitable for use as reinforcement in structures subjected to
aggressive environments, such as marine structures, bridges, and parking garages. Among
different types of FRPs, glass FRP (GFRP) bars have drawn more attention in civil
engineering applications due to their low cost com-pared to other types of FRPs [1].
Durability of GFRP bars, however, is not a straightforward subject; it tends to be more
complex than corrosion of steel reinforcement because the degradation of the material
could depend on both re-sin and fibers and on their interface bond behavior. Therefore,
accelerated testing and evaluation pro-grams are needed to evaluate the expected service
performance of concrete members reinforced with these bars. In addition, calibration of the
accelerated test results with different natural weathering data of in-service structures is
needed to establish safe service life of a structure [2].

659

There have been several durability studies that were carried out in USA, Canada, Japan and
some other countries. These studies concluded that most common types of GFRP bars
appeared to have some deterioration problems when subjected to harsh environments [3].
Many of these studies have been carried out on the old generations of FRP bars. The
manufacturers of GFRP are now claiming that they have produced new types of GFRP bars
that have greater resistance to alkaline and to other environmental conditions. Therefore,
before prescribing the materials to practitioners, there is an essential need to evaluate the
long-term performance of the new GFRP bars when subjected to different environmental
and loading conditions.
2. OBJECTIVES
This study aims to investigate the durability of the new generations of GFRP bars under
accelerated laboratory environmental conditions as well as actual field conditions. The
laboratory environments include exposure to ordinary tap water, sea water, and alkaline
solution at different temperatures. The field conditions include the hot weather field
conditions of the Middle East and the Arabian Gulf area in particular. This is represented in
this study by the hot-dry (Central Province) and hot-humid (Eastern Province),
environments of Saudi Arabia.
3. EXPERIMENTAL PROGRAM
The experimental program of this study includes two phases. The first phase investigates
the short-term mechanical properties of the available types of GFRP bars in the market to
select the best type to be used in the second phase. The second phase investigates the
durability of the selected type of the GFRP bars.
3.1 Phase 1: Short-Term Mechanical Properties
The properties of composite materials are dependent on the individual component
properties, the manufacturing technique and the quality control of the production process.
Any variations in the characteristics of these three items will produce composite materials
with variable short-term mechanical properties. Durability study cannot be conducted using
materials with intolerable variations in short-term mechanical properties. Such variation
will make the results of the durability study meaningless. There-fore, before commencing
the durability tests, tensile tests were carried out to identify the short-term mechanical
properties of the available GFRP bars.
Four different types of GFRP bars have been procured from three different suppliers (Table
1). All the GFRP bars were subjected to the screening test to determine their properties.
GFRP bars which scored the highest stable results were used for the durability study.
Tensile test was used to identify the suitable bars. Information collected from this test
includes: tensile strength, tensile modulus of elasticity, and strain at failure. Ten GFRP bar
specimens of each type have been used in the screening tensile tests. The tensile tests were
carried out according to the ACI 440.3R-04 B2 test method [4].

660

Table 1. Types of GFRP bars used in screening test


Supplier Country
Type
Diameter (mm)
Fiber type/resin type
USA
Type I
10
E-Glass/Vinylester
Germany
Type II
12
E-Glass/Vinylester
Type
III
16
E-Glass/Polyester
New Zealand
Type IV
16
E-Glass/Vinylester
For all tested bars, the stress-strain relationships were linear up to failure as shown in Fig.
1. The failure was a brittle fracture with delamination and/or shearing of the matrix. The
average, standard deviation (SD), and coefficient of variation (COV) of the tensile test
results for all tested specimens are given in Table 2.
1400
1200

Stress (MPa)

1000
800
600
400

Type-I
Type-II
Type III
Type IV

200
0
0.000

0.005

0.010

0.015

0.020

0.025

0.030

Strain (mm/mm)

Fig. 1. Typical stress-strain relatioships of the tested GFRP bars


Table 2. Screening tests (tensile test) results of GFRP bars.
Peak Stress
Peak Strain
Modulus of
FRP Samples
[MPa]
[mm/mm]
Elasticity [GPa]
Average
432
0.0104
41.9
Type I
S.D.
59.2
0.0016
1.56
COV (%)
13.68
15.76
3.71
Average
1478
0.0245
60.4
Type II
S.D.
29.9
0.0005
1.69
COV (%)
2.03
2.15
2.79
Average
838
0.01
59.9
Type III
S.D.
48.9
0.0008
2.45
COV (%)
5.84
5.91
4.08
Average
611
0.0097
63.3
Type IV
S.D.
66.8
0.0016
4.44
COV (%)
10.93
15.93
7.01

661

It can be noticed from Fig. 1 and Table 2 that Type II of the tested GFRP bars had the best
engineering characteristics in comparison with other GFRP bars considered in the screening
tests. Lower variations in the results of GFRP bars Type II were also observed in
comparison with other types. Based on the screening test results, GFRP bars Type II (EGlass/Vinylester) were selected to be used in the durability phase of this investigation.
3.2 Phase 2: Durability of GFRP Bars
This second phase of the study investigates the effect of different environmental conditions
(control, laboratory, and field) on the tensile properties of the selected GFRP bars. To be
closer to real field conditions where the bars are embedded in concrete, the GFRP bars used
in this study were embedded in concrete prisms before aging in environmental conditions.
For the test specimens, concrete was first cast and cured under normal conditions then the
specimens were transferred into the different environmental conditions until the day of
testing.
3.1.1 Test Specimens
The durability study presented in this paper focuses on the sole effect of different
environmental conditions on test specimens without applying any stress on the GFRP bars.
Fig. 2 shows a schematic drawing and a photo of the test specimens. For these specimens,
GFRP bars centrally embedded in concrete prisms (40 40 450 mm) were used. Both
ends of the GFRP bars were protected against environmental conditions using a plastic
tape. These ends were used later for the anchor-age of the bars before testing. The
specimens were cast in wooden molds. Before casting, all gaps in the molds were filled
with silicon and brushed with oil. After demolding, specimens were cured using wet
burlaps for a period of 28 days.

40

100mm

450mm

.Fig 2. Test specimens before and after casting

662

100mm

3.1.2 Environmental Conditions


The specimens have been subjected to ten environmental conditions for 6 and 12 months.
The environments included expo-sure to ordinary tap water, sea water, and alkaline solution
at various temperatures. Sea water was brought from the Arabian Gulf- Eastern Province of
Saudi Arabia. Alkaline solution was prepared using Calcium Hydroxide, Potassium
Hydroxide and Sodium Hydroxide (1.185 g of Ca(OH)2 + 9.0 g of NaOH + 42.0 g of
KOH per 10 liters of water). The environments also included two typical field conditions of
the Kingdom of Saudi Arabia (Riyadh area and Jubail area). The environmental conditions
are summarized as follows:
(a) Unconditioned Laboratory Specimens (Control Specimens):
Specimens of this group, LE, have been exposed to controlled laboratory environment
(temperature 23 2C). Similar to the specimens in other conditions, after 6 and 12 months
of exposure, bars were extracted out of the specimens and tested in tension. Test results of
this group were used as reference results for specimens under all other conditions.
(b) Laboratory Specimens under Different Environ-mental Conditions:
The specimens of this group have been exposed to different environmental conditions (at
normal room temperature or at 50C) by either immersion into different types of liquids or
exposure to some specified conditions. Similar to the unconditioned specimens, the
conditioned specimens were tested in tension after being immersed in the different
solutions for 6 and 12 months. The effects of each environmental condition on the tensile
strength of the bars were determined by comparing the test results of this group with those
of the counterpart control specimens, LE specimens. The environmental conditions
considered in this group of testing are as follows:
(i) Immersion in tap water at ambient and 50C;
(ii) Immersion in sea water at ambient and 50C;
(iii) Wet/dry cycles in sea water at 50C;
(iv) Immersion in alkaline solution at 50C; and
(v) Hot-dry condition at 50C.
(c) Field Conditions:
Two field conditions were considered for this category of specimens. A set of specimens,
RF, was exposed to Riyadh hot-dry field conditions (representing hot-dry arid land of the
Middle East). A second set, JF, was exposed to the Eastern coast of the Kingdom of Saudi
Arabia (Jubail City) which represents the hot-humid environment of the Middle East. It
should be mentioned that the monthly average temperature is almost similar for both field
conditions. It ranges between 9 and 44C. However, the annual average relative humidity is
about 26 and 52% in Riyadh and Gulf areas, respectively [5].
3.1.3 Tensile Test
Similar to the specimens in Phase 1, all bars were tested in tension. Three specimens from
each environment were tested to determine the effects of different environmental conditions

663

on the tensile properties of the bars. Each specimen was instrumented with a Linear
Variable Differential Transformer (LVDT) to capture the elongation during testing.
The test was carried out using an INSTRON testing machine and the load was increased
until failure. The rate of loading ranged between 250 and 300 MPa/min. The applied load
and bar elongation were recorded during the test using a data acquisition system monitored
by a computer.
4. TESTS RESULTS AND DISCUSSION
4.1 Tensile Strength
Fig. 3 shows the tensile strength retention of the specimens as a function of exposure time
and temperature. The tensile strength of the conditioned specimen is divided by that of the
unconditioned specimen in the controlled lap environment, LE, and the corresponding
value is denoted as the retention ratio in percent.
For the specimens in the hot-dry laboratory environment, HD50, and the two field
conditions: Riyadh field condition, RF, and Gulf (Jubail) field condition, JF, almost no
reduction in the tensile strength (1 to 2%) was recorded after 18 months. This indicates that
the used laboratory environments were too harsh compared to the real field conditions.
For the test specimens in TWR, a slight gradual reduction in the tensile strength was
recorded after 6, 12, and 18 months of exposure. The residual strengths were about 95, 94,
and 91%, respectively. The maximum reduction in the residual tensile strength was
recorded for the specimens in tap water at 50oC, TW50. After 6 months of exposure, the
strength loss was 17%. After 12 and 18 months of exposure, this strength loss increased to
23 and 31%, respectively. This indicates that increasing the temperature to 50oC increased
the diffusion rate of water and harmful ions into the bars which resulted in a faster
degradation in the glass fibers leading to that decrease in the tensile strength with time.
For the specimens in seawater at room temperature, SWR, the reduction in the tensile
strength increased gradually with time. It was 6, 8, and 14%, respectively, after 6, 12, and
18 months of exposure which is comparable to the results in TWR. Increasing the
temperature to 50oC, (SW50), resulted in additional reduction in the tensile strength after 6
months of exposure. This reduction, however, did not increase with time. After 6, 12, and
18 months, the strength loss was about 14, 12, and 11%, respectively. For the specimens in
dry/wet seawater at 50oC, SW50DW, the recorded tensile strengths were greater than those
obtained in the SW50 specimens. This indicates that the dry/wet condition has less harmful
effect on the tensile strength of the tested GFRP bars which is in agreement with the results
obtained by Chen et al. [5]. This could be related to the absence of the seawater at the dry
stage which results in less diffusion of the solution into the bars compared to the SW50
condition. For both SW50 and SW50DW conditions, it can be noticed that the tensile
strengths were almost stable with time. The tensile strengths after 12 and 18 months of
exposure were almost similar to those obtained after 6 months. This may be attributed to
the formation of a very thin layer of salt on the concrete surface, especially at higher

664

temperature, which decreases the diffusion rate of the solution into the bars. This was not
recorded in the TW50 specimens.

98
100
99

99
97
98

18Months

92
88
80

93
93
91

86
88
89

94
92
86

12Months

100
101
98

80

83
77
69

95
94
91

100

100
100
100

RetentionofTensileStrength(%)

6Months
120

HD50

RF

JF

60
40
20
0
LE(control) TWR

TW50

SWR

SW50 SW50DW ALK50


Environment

Fig. 3. Retention of tensile strength for all test specimens


For the specimens in alkaline solution at 50oC, ALK50, a significant reduction in the tensile
strength was recorded with time. After 6, 12, and 18 months of exposure, the decrease was
8, 12, and 20%, respectively. This reduction in the tensile strength was less than that
obtained in the TW50 specimens. This indicates that, for the GFRP bars embedded in
concrete, the tap water was more aggressive than the alkaline solution. This could be
attributed to the concrete around the GFRP bars. It is well known that the internal concrete
environment has high alkalinity with pH between 10.5-13, depending on the design mixture
of the concrete and type of cement used. This alkaline environment damages the glass fiber
through loss in toughness and strength. In addition, calcium, sodium and potassium ions
found in concrete pore solution are highly aggressive towards glass fibers [6].
4.2. Youngs Modulus and Strain at Failure
The flexural modulus of all conditioned specimens was not significantly affected compared
to the control specimens. For all environments, a slight decrease ranging between 1 and
12% was observed. This behavior is due to the fact that flexural modulus of glass fiber
reinforced polymers is only affected by temperatures approaching polymer glass transition
temperature.
The reductions in the tensile strains at failure were almost similar those recorded in the
tensile strength. For the specimens in TWR, HD50, RF, and JF, almost no decrease in the
strain at failure was recorded after 18 months of exposure. The specimens in the seawater
solution at room temperature and at 50oC show a decrease of about 4 and 6%, respectively,
after the 18 months of exposure. The reduction was about 10% for the specimens in the wet
and dry seawater exposure at 50oC. The specimens in the alkaline solution at 50oC show a
decrease of about 12% after 18 months of exposures. Similar to the degradation in the
tensile strength, the maximum reduction in strains at failure was recorded for the specimens

665

in tap water at 50oC. After 6, 12, and 18 months of exposure, the reduction in strains at
failure was 10, 18, and 22%, respectively.
5. CONCLUSIONS
This study is a part of an ongoing durability research program on fiber-reinforced polymer
(FRP) reinforcing bars for concrete structures. The test results show that the durability of
tested concrete-wrapped GFRP bars exposed to actual filed conditions was not affected
after 18 months of exposure. Accelerated laboratory environments showed some effects of
the tensile strength of the bars. The TW50 environment followed by the ALK50 had the
maximum effect on the tensile properties of the tested concrete-wrapped GFRP bars.
However, it is reasonable to assume that these environments are too harsh compared to the
two real field conditions which did not cause any degradation in the tensile properties of the
tested GFRP bars after one year and half.
6. ACKNOWLEDGMENT
The authors would like to acknowledge the Center of Excellence for Research in
Engineering Materials (CEREM), College of Engineering, King Saud University, for its
support to this research project. The authors are also thankful to the technicians of the
Department of Civil Engineering, King Saud University, for their help in fabricating and
testing of the specimens.
7. REFERENCES
1.
2.
3.

4.
5.
6.

El-Gamal, S.E., El-Salakawy, E.F. and Benmokrane, B. 2007. Influence of


Reinforcement on the Behavior of Concrete Bridge Deck Slabs Reinforced with FRP
Bars. Journal of Composites for Constructions, ASCE, 11(5): 449-458.
Ceroni, F., Cosenza, E., Gaetano, M., and Pecce, M. 2006. Durability issues of FRP
rebars in reinforced concrete members. Cement & Concrete Composites, 28: 857868.
Almusallam, T., Al-Salloum, Y., Alsayed, S., and Alhozaimy, A. 2002. Durability of
GFRP Rebars in Concrete Beams Under Sustained Loads at Severe Environments.
Proceedings of the 6th International Symposium on FRP Reinforcement for Concrete
Structures (FRPRCS-6), Edited by Kiang Hwee Tan, 2002, Singapore, pp. 823-832.
ACI Committee 440. 2004. Guide Test Methods for Fiber-Reinforced Polymers (FRPs)
for Reinforcing or Strengthening Concrete Structures. ACI 440.3R-04, American
Concrete Institute (ACI), Farmington Hills, Michigan, 40 p.
Chen, Y. Davalos, J., Ray, I., and Kim, H. 2007. Accelerated aging tests for
evaluations of durability performance of FRP reinforcing bars for concrete structures.
Composite Structures, 78: 101111.
Correia, J., Cabral-Fonseca, S., Branco, F., Ferreira, J., Eusbio, M., and Rodrigues, M.
2005. Durability of Glass Fibre Reinforced Polyester (GFRP) Pultruded Profiles Used
in Civil Engineering Applications. Proceedings of the Third International ConferenceComposites in Construction, Lyon, France, 1113 July.

666

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

DURABILITY OF CONCRETE-FILLED FRP TUBES SUBJECTED TO


SALT SOLUTION
M. Robert1, A. Fam2 and B. Benmokrane3
1

Postdoctoral Fellow, Dept. of Civil Eng., Queens University, Kingston, Canada


Canada Research Chair Professor, Dept. of Civil Eng., Queens University, Kingston, Canada
3
Canada & NSERC Research Chairs Professor, Department of Civil Engineering, University of
Sherbrooke, Quebec, Canada
2

ABSTRACT
This paper presents mechanical, microstructural and physical characterization of concretefilled FRP tubes (CFFTs) used as structural elements such as pier columns and marine
piles. GFRP tubes filled with normal concrete were exposed to salt solution at 23, 40, and
50C to accelerate the effect of the environment and of the salt solution. The measured
apparent hoop tensile strength of the tube before and after exposure were considered as a
measure of the durability performance of the specimens and were used for long-term
properties prediction based on the Arrhenius theory. In addition, Fourier Transform
Infrared Spectroscopy (FTIR), Differential Scanning Calorimetry (DSC) and Scanning
Electron Microscopy (SEM) were used to characterize the aging effect on the CFFT. The
main objective of this study is to evaluate the durability and to predict the long-term
behaviour of the GFRP tubes used in the CFFT system. The test results showed that the
GFRP tubes investigated in this study present a very good stability in the environment
simulating the field service conditions. The maximum observed reduction in hoop tensile
strength under the rather aggressive environment was 20%.
1. INTRODUCTION
The concrete-filled FRP tubes (CFFTs) system has been used successfully for different
structural applications [1]. The CFFTs can be used as pier columns and girders for bridges
[2] and also as fender piles in marine structures [3]. CFFTs provide several advantages such
as confinement, permanent formwork, provide shear and/or flexural reinforcement, and
protect the concrete from aggressive environments. Under some special conditions, such as
in high alkalinity of concrete used to fill the tubes or in saline environments of seawater or
deicing salts, the long-term performance of the CFFT remains unknown. Strength of the
glass fibres and resin matrix, the two constituents of the GFRP materials, can decrease
when subjected to wet alkaline or saline environments. Thus, an adequate durability study
on the GFRP material should be performed in order for CFFT systems to become widely
accepted in the construction industry. CFFT systems in most cases consist of a filament
667

wound GFRP tube which acts as a stay-in-place form for a concrete core. From a structural
point of view, the presence of the GFRP tube eliminates the need for internal
reinforcement. Moreover, the GFRP tube prevents ingress of moisture and ions into the
concrete which lead to an enhanced durability of the piles [4]. However, the diffusion of
free hydroxyl ions (OH-), NaCl or H2O molecules through the matrices of the CFFT can
lead to the deterioration of GFRP material [5]. Some studies have been conducted on the
evaluation of durability of FRP-confined concrete [6-8], but the data on durability of
filament wound FRP tubes used for CFFT technique is very limited. In this paper,
evaluation of environmental durability and prediction of long-term behavior of the tubes
used in the CFFT system was studied.
2. EXPERIMENTAL PROGRAM
2.1 Material
The GFRP tubes used to fabricate the CFFT used in this study were manufactured by
Ameron International. The tubes were made of continuous E glass fibres impregnated in an
epoxy resin using the filament winding process. The mass fraction of glass is 66.4 % and
was determined by thermogravimetric analysis according to ASTM E 1131 standard. Their
relative density according to ASTM D 792 standard is 1.57. The used tube has a nominal
outer diameter of 169 mm and a wall thickness of 3.5 mm, comprised of nine layers in
alternating angles of 9 and -86 with respect to the longitudinal axis. The 9 layers add up
to 37 percent of the total wall thickness. All specimens were cut into 25.4 mm wide disks as
specified by the ASTM D 2290 standard, from a long CFFT member. The tubular
specimens were divided into two series; 1) the unconditioned reference samples; and 2)
conditioned samples (60 specimens). The concrete mixture used in filling the tube led to a
concrete pH of 12.07 measured by the extraction of the interstitial solution after aging. The
concrete was cast in the tube before the aging of the CFFT disks.
2.2 Test Plan
The samples were immersed in salt solution of 3% NaCl. The pH of the solution inside the
CFFT is a result of the absorption of water by concrete, thus allowing the liberation of the
alkaline ions of the concrete directly in the FRP environment. The solution level was kept
constant throughout the study to avoid a pH increase which could be due to a solution level
decrease and a significant increase of the concentration of the alkaline ions in the solution.
The specimens were completely immersed at three different temperatures (23, 40 and 50oC)
and were removed from the solution after four different periods of time (90, 180, 270 and
365 days).
2.3 Hoop Tensile Tests
The concrete core was removed from all the GFRP disks and hoop tensile strength of the
tubes was determined by the split disk method, for both reference and aged samples. The
tests were performed according to procedure A of ASTM D 2290 test method except that
the test specimens were full section specimens instead of reduced section specimens. Five
specimens were tested after each conditioning to evaluate the hoop tensile strength
668

retention after different durations of aging. The test fixture used to perform the test involves
two half-disk shape components, which were connected to the upper and lower gripping
arms of the test fixture, with pins. Split-disk test specimens were located between the two
half disk shape components and the upper and lower connecting arms. The test fixture was
therefore designed to minimize the effect of this bending moment.
2.4 Arrhenius Relation
Equation 1 expresses the Arrhenius relation, in terms of the degradation rate [9].
Ea
k A exp

RT

(1)

where k = degradation rate (1/time); A = constant relative to the material and degradation
process; Ea = activation energy of the reaction; R = universal gas constant; and T =
temperature in Kelvin. The primary assumption of this model is that only one dominant
degradation mechanism of the material operates during the reaction and that this
mechanism will not change with time and temperature during the exposure [10]. Only the
rate of degradation will be accelerated with the temperature increase. Equation 1 can be
transformed into:
1 1
E
exp a
k A
RT
1 E 1
ln a ln A
k R T

(2)
(3)

From equation 2, the degradation rate k can be expressed as the inverse of time needed for a
material property to reach a given value [10]. From equation 3 one can further observe that
the logarithm of time needed for a material property to reach a given value is a linear
function of 1/T with the slope of Ea /R [10]. Ea and A can be easily calculated by using the
slope of the regression and the point of intersection between the regression and the Y axis
respectively.
2.5 Differential Scanning Calorimetry (DSC)
Twelve-milligram to 15-milligram specimens from both unconditioned and aged samples
were sealed in aluminum pans and analyzed in a TA Instruments DSC Q10 calorimeter
equipped with a refrigerated cooling system. Specimens were heated from 25C to 195C at
a rate of 5oC/min. Glass transition temperature was determined for both specimens in
accordance with ASTM E 1356 standard
2.6 Fourier Transform Infrared Spectroscopy (FTIR)
Fourier transform infrared spectroscopy (FTIR) spectra were recorded using a Nicolet
Magna-550 spectrometer equipped with an attenuated total reflectance device. Fifty scans
669

were routinely acquired with an optical retardation of 0.25 cm to yield a resolution of 4


cm1.
2.7 Scanning Electronic Microscopy (SEM)
SEM observations and image analysis were performed to observe the microstructure of
specimens before and after aging. Samples observed in the SEM were the unconditioned
specimens and specimens aged during 12 months in salt solution at 50oC, which is the
harshest aging environment. These observations were conducted to see the potential
degradation of polymer matrix, glass fibres or interfaces, if any.
3. EXPERIMENTAL RESULTS
3.1 Hoop Tensile Strength Retention
The hoop tensile test of unconditioned and aged samples showed an approximately linear
behavior up to failure. Specimens failed through the rupture of fibers. The failure was
accompanied by the delamination of fibers and resin, as shown in Figure 1b. Erdiller [11]
also observed similar tensile failure modes of filament wound tubes.
Reference
Aged

b)
Fig. 1 . Hoop tensile test specimens : a) ready for tests, and b) mode of failure
Table 1 shows the experimental results obtained during the hoop tensile tests concerning
the hoop tensile strength of aged specimens tested after immersion. Figure 2 shows the
retention of the hoop tensile strength of aged specimens according to the duration of
immersion of filled tubes at various temperatures. As shown in Table 1, the hoop tensile
strength for unconditioned CFFT tube specimen was equal to 588 12.5 MPa. Note that
the hoop tensile strength of CFFT tube specimens was reduced to 464 6.6 MPa after 365
days exposure to water at 50C. Figure 2 shows a slight decrease of the ultimate hoop
tensile strength with immersion duration. Furthermore, it is clear that the temperature of
immersion affects the loss of resistance. It can be seen that for duration of immersion of 12
months, the loss of resistance is equal to 20, 17 and 10 % at 50o, 40o and 23oC, respectively.
This phenomenon is due to the increasing of the diffusion rate of the solution and to the
acceleration of the chemicals reaction of degradation with the temperature of immersion.

670

This leads to a larger absorption rate of the solution for the same time of immersion and
accelerated reaction of degradation. The absorption of solution can lead to a degradation of
the fibres, the matrix and fibre/matrix interface, leading to a loss of the ultimate hoop
tensile strength.
Table 1. Experimental Hoop Tensile Strength of Reference and Conditioned Specimens
Time of
Temperature Mean Hoop Tensile
COV
immersion (days)
(oC)
Strength (MPa)
(%)
0
23
588
3.2
23
569
4.4
90
40
537
5.2
50
508
7.3
23
541
6.1
180
40
501
2.4
50
482
4.4
23
528
2.5
270
40
492
3.0
50
468
6.4
23
526
4.2
365
40
490
2.0
50
464
3.2

Fig. 2. Hoop tensile strength retention of conditioned CFFT aged in salt solution at 23o, 40o,
and 50C
3.2 Effect on Polymer Matrix
The change in the Tg of CFFT after aging in salt solution was also investigated. No
significant change of the Tg was measured after the aging in salt solution during 365 days at
50oC. In fact the Tg of reference specimen was 127oC and the Tg of aged specimens was
129oC. These results clearly show that even after 365 days in salt solution at 50oC, the
polymer matrix is not affected.
A FTIR analysis of unconditioned and CFFT immersed in salt solution during 365 days at
50oC has been conducted. The most interesting region of the FTIR spectra is located
671

between 3300 cm-1 and 3600 cm-1, which corresponds to the stretching mode of the
hydroxyl groups of the epoxy resin. When a hydrolysis reaction occurs, new hydroxyl
groups are formed and the corresponding infrared band increases. Changes in the peak
intensity were quantified by determining the ratio of the OH peak to the carbon-hydrogen
stretching peak of the resin, which is not affected by the conditioning. The experimental
ratio of the OH peak to the carbon-hydrogen stretching peak of the resin for the aged
samples was 0.67 compared to 0.70 for unconditioned samples. So, the hydroxyl peak does
not show any significant changes. This indicates that the hydrolysis did not significantly
occur in these environmental conditions.
3.3 Microstructural Observations
3.3.1 External Surface
SEM observations of external surface of CFFT were performed to investigate the surface
deterioration of the polymer matrix after conditioning in salt solution. Figure 3 presents
micrographs of the surface of reference and aged dowels.

a)
b)
Fig. 3. Micrographs of the tube external surface for: a) reference tube specimen; b) tube
specimen aged in salt solution during 365 days at 50oC
It can be observed that the aging in salt solution leads to the creation of cracks or microcracks in the epoxy matrix. These cracks could be the result of thermal degradation or
chemical degradation by the salt solution. However, the presence of cracks or microcracks
can lead to an increase of the moisture absorption at saturation. In fact, the moisture
absorption at saturation has increased from 0.45% before aging to 0.61% after aging in salt
solution. This 36% increase of the moisture uptake at saturation could lead to the
plasticization of the polymer matrix and can explain the slight decrease of hoop tensile
strength.
3.3.2 Microstructural Effects
Figure 4 presents micrographs for both reference specimens and specimens aged in salt
solution during 365 days at 50oC. The visual and microstructural observations showed no
significant damage on CFFT after 365 days of immersion in the salt solution at the highest
temperature (50oC). Observations of the fibre/matrix interface and of the microstructure, in

672

general, demonstrate that the conditionings of CFFT in salt solution do not affect the
microstructural properties of the CFFT (Figure 4b).

a)
b)
Fig. 4. Micrographs at the fibre/matrix interface before mechanical tests for: a) reference
CFFT, b) CFFT aged in salt solution during 365 days at 50oC
3.4 Service Life Prediction
Following the procedure proposed by Bank et al. [12], the natural logarithm of time to
reach a set of levels of normalized performances versus 1/T, expressed as the inverse of
absolute temperature (1000/K), was used to predict the service life at the Mean Annual
Temperature (6oC) in Montral, Qubec, Canada. From the Arrhenius plot, the service life
time necessary to reach the established hoop tensile strength retention levels (PR) can be
extrapolated for any temperature. Consequently, predictions are made for hoop tensile
strength retention as a function of time for an immersion at 6C and the general relation
between the PR and the predicted service life at the average temperature of 6oC are drawn
(Figure 5). The predicted time to reach the determined strength property retention level
(PR) for GFRP tubes filled with moist concrete and immersed in salt solution at an
isotherm temperature of 6oC is approximately 3 and 48 years for a PR of 90 and 70%,
respectively. Also, the predicted service life of GFRP tubes filled with moist concrete and
immersed in salt solution at an isotherm temperature of 6oC to reach a PR less than 50%
can be estimated to be infinite. These predictions show that the GFRP tubes are durable
face to the concrete environment and to salt environment, which is supposed to be well
simulated by the immersion of concrete-filled GFRP tubes in salt solution.

Fig. 5. General relation between the PR and the predicted service life at a mean annual
temperature of 6oC (Montreal, Quebec)
673

4. CONCLUSIONS
Based on the results of this study the following conclusions may be drawn on the tested
product only:
1- Even at high temperature (50oC), where the environment is quite aggressive, the
reduction in hoop tensile strength is minor. For example, increasing the temperature of
the solution from 40 to 50oC during 365 days results in a decrease of the hoop tensile
strength retention by 17 to 20% of the original hoop tensile strength.
2- No significant microstructural changes were observed after 365 days immersion of
CFFT in salt solution at 50oC. The interface between the resin and the fibres doesnt
seem to be affected by the moisture absorption and high temperatures.
3- No changes of the glass transition temperature occur as observed by differential
scanning calorimetry. FTIR did not show any significant changes of the polymer
chemical structure, i.e. degradation. On the other hand, microstructural observations of
the external surface of the aged CFFT showed the presence of cracks or microcracks
which lead to an increase of the moisture uptake at saturation.
4- According to the long-term predictions, the hoop tensile strength retention of CFFT
immersed in salt solution will decreased by 10 and 30% after 3 and 48 years,
respectively. It was shown that the service life time required to reach tensile strength
retention of less than 50% should be infinite.
7. ACKNOWLEDGEMENTS
This research was supported by the Natural Science and Engineering Research Council of
Canada (NSERC Postdoctoral Fellowship Program).
8. REFERENCES
1.
2.
3.
4.
5.
6.
7.

Fam, A. and Rizkalla, S.H. 2001. Confinement model for axially loaded concrete
confined by circular fiber-reinforced polymer tubes. ACI Structure Journal, 98(4):
451461.
Son, J.-K., and Fam, A. 2008. Finite element modeling of hollow and concrete-filled
fiber composite tubes in flexure: Model development, verification and investigation of
tube parameters. Engineering Structures, 30: 26562666.
Karbhari, V.M. 2004. Fiber reinforced composite bridge systems transition from the
laboratory to the field. Composite Structures, 66: 516.
Rizkalla, S. 2009. FRP for Sustainable Precast Concrete Structures. Proceedings of the
US-Japan Workshop on Life Cycle Assesment of Sustainable Infrastructure Materials,
Sapporo, Japan.
Robert, M., Cousin, P., and Benmokrane, B. 2009. Durability of GFRP Bars Embedded
in Moist Concrete. Journal of Composites for Construction, ASCE, 13(2): 66-73.
Soudki, K.A. and Green, M. F. 1997. Freeze-thaw response of CFRP wrapped
concrete. Concrete International, 19(8): 6467.
Toutanji, H.A. and Balaguru, P. 1999. Effects of freezethaw exposure on performance
of concrete columns strengthened with advanced composites. ACI Material Journal,
96(5): 605610.

674

8.

Karbhari, V.M., Rivera, J., and Dutta, P.K. 2000. Effect of short-term freezethaw
cycling on composite confined concrete. Journal of Composites for Construction,
ASCE, 4(4): 191197.
9. Nelson, W. 1990. Accelerated testingStatistical models, test plans, and data analyses.
Wiley, New York.
10. Chen, Y., Davalos, J.F., Ray, I. 2006. Durability Prediction for GFRP Bars Using
Short-Term Data of Accelerated Aging Tests. Journal of Composites for Construction,
ASCE, 10(4): 279-286.
11. Erdiller, E.S. 2004. Experimental Investigation for Mechanical Properties of Filament
Wound Composites Tubes. MSc Thesis in Mechanical Engineering, Middle East
Technical University, Ankara, Turkey, 149 p.
12. Bank, L.C, Gentry, T.R, Thompson, B.P, and Russel, J.S. 2003. A Model Specification
for Composites for Civil Engineering Structures. Construction and Building Materials,
17(6-7): 405-437.

675

676

CDCC-11

DURABILITY AND SUSTAINABILITY OF FIBRE REINFORCED POLYMER


(FRP) COMPOSITES FOR CONSTRUCTION AND REHABILITATION
DURABILIT ET CYCLE DE VIE DES MATRIAUX COMPOSITES
EN POLYMRES RENFORCS DE FIBRES (PRF) EN
CONSTRUCTION ET RHABILITATION

FREEZE-THAW CYCLING EFFECT ON THE MECHANICAL


PROPERTIES OF FRP-FILAMENT-WOUND TUBES
H. El-Zefzafy1, H.M. Mohamed2 and R. Masmoudi3
1

Ph.D. Candidate, University of Sherbrooke, Sherbrooke, QC, Canada


NSERC-Postdoctoral Fellow, University of Sherbrooke, Sherbrooke, QC, Canada
3
Professor , University of Sherbrooke, Sherbrooke, QC, Canada
2

ABSTRACT
The advantages of fiber-reinforced polymer (FRP) composite material have attracted
structural engineers as alternative materials of the traditional wood, steel and concrete.
Filament wound FRP tubes are one of the most promising products can be used in different
applications in civil engineering industry. In this study, an experimental work was
performed to investigate the short and long term effect of freeze-thaw cycles on the
mechanical behavior in axial direction of the filament wound glass FRP tubes. The tubes
had an internal diameter equal to 152 mm, thickness 2.65 mm, and the fiber orientations
were mainly in the hoop direction (60). Test parameters in this investigation include the
effect of number of freeze-thaw cycles (100 and 300) and the type of freeze-thaw cycles
exposure (dry air, fresh and salt water). A total of seventy coupon specimens were cut from
the FRP tubes, sixty were conditioned to 100 and 300 freeze-thaw cycles, while the rest
were kept in the room temperature. Then, pure axial tension and compression tests were
conducted in order to evaluate the change of mechanical properties of the test specimens
due to the freeze-thaw exposure. The test results indicated that the axial tensile and
compressive strength behavior of the FRP tubes were increased after short and long term
freeze-thaw cycles considered in this study in the three different conditions.
1. INTRODUCTION
Recently, the use of fiber-reinforced polymers (FRP) products in civil engineering
applications has come about as a result of the many desirable characteristics that are
superior to those of conventional materials such as steel, concrete, and wood. One of these
products is filament wound FRP tubes which can be used in application of overhead power
lines and telecommunications structures. In addition, the FRP tubes have been introduced in
the civil engineering community as structurally integrated stay-in-place forms for concrete
members, such as beams, columns, bridge piers, piles and fender piles as an innovative
solution to the corrosion problem. In such integrated systems, the FRP tubes may act as a
permanent form, often as a protective jacket for concrete, and especially as external
reinforcement in the primary and secondary directions such as for confinement (Moahmed
677

and Masmoudi 2008; 2010; Moahmed 2010; Fam and Rizkalla


Shahawy 1997).

2002; Mirmiran and

However, there are interests to the overall durability of FRP composite materials, especially
as related to their capacity for sustained load performance under harsh environmental
conditions. In aggressive environments, FRP systems are subjected to moisture, salts,
alkalises, ultraviolet radiations and freeze-thaw cycles, which it may cause degradation to
the resin and fibres, hence limiting the strength. Research work related to the durability
effect on the mechanical properties of the FRP tubes is, however, very limited to date. This
paper presents the mechanical behavior in the axial direction of the filament wound FRP
tubes conditioned to short and long term freeze-thaw cycles in three different conditions.
2. EXPERIMENTAL WORK
2.1

Filament-Wound Glass FRP Tubes

In this study, coupon specimens were cut from glass-fiber reinforced polymer (GFRP)
tubes and conditioned to different exposure and then tested under axial tension and
compression. The tubes were manufactured using continuous filament winding process
adopted by FRE Composites, St-Andre-dArgenteuil, Quebec, Canada. E-glass fiber and
Epoxy resin were utilized for manufacturing these tubes. The glass fiber volume fraction as
provided by the manufacture was 68% 3%. The material properties for both the fiber and
the resin, as given by the manufacture, are presented in Table 1. The GFRP tubes consist of
four (60) layers oriented in the hoop direction with respect to the longitudinal axis of the
tubes, the total thickness is 2.65 mm, with an internal diameter equal to 152 mm. The
winding angles of the tubes were optimized for below underground pipe applications.

Linear mass
km )
Tensile
modulus(MPa)
Poissons ratio
2.2

Table 1. Mechanical and physical properties of fibres and resin


Glass fibers
Epoxy resin
(g /
2000
Density (kg / m3 )
1200
Tensile
modulus(MPa)
Poissons ratio

80 000
0.25

3380
0.4

Specimens Details and Preparation

Two different experimental tests were carried out to measure the mechanical properties of
the FRP tubes in the axial direction after and before freeze-thaw cycles. Seventy coupon
specimens were prepared and tested in this study. Thirty specimens were exposed to 100
freeze-thaw (F/T) cycle, and other thirty specimens were maintained up to 300 F/T cycles.
The rest ten specimens were kept in the room temperature. The conditioned coupon
specimens were kept inside an environmental chamber under three different conditions.
Each five coupons of the tension and compression specimens were kept in dry air,
submerged in fresh and salt water inside the chamber. The specimens submerged in salt
water were to simulate the environment of de-icing conditions of the infrastructures. The

678

following section presents an experimental program that was conducted to determine the
mechanical properties of the FRP tubes in axial direction after 100 and 300 F/T cycles.
2.2.1

Freeze-Thaw Exposure

The dry freeze-thaw specimens were placed in the freeze-thaw chambers. Whereas, those
for saturated freeze-thaw were left in fresh and salt water bath and placed in the same
environmental chamber using the isolated wooden tanks, see Figure 1. According to the
different types of environmental conditions (freezethaw cycles in dry, fresh and salt water)
existed inside the same environmental chamber, it was hard to reach the same temperature
inside the saturated and non-saturated specimens. Therefore, a temperature acquisition
system was used to monitor the temperature during cycling. The freezing cycles consisted
of lowering the temperature in the middle of the water bath from 4.4 to -17.8C in a period
of 16.5 h. The thawing cycles consisted of raising the temperature in the middle of the
saturated specimens from -17.8 to 4.4C. Those freeze-thaw hours were sufficient to vary
the temperature of non-saturated specimens between (+28C to -28C) as shown in Figure
2. Thus, the specimens underwent one freeze-thaw cycles per 27 hours, rather than in
accelerated shorter cycles.

a. Specimen in wooden tanks inside


b Monitoring the chamber
the chamber
during F/T cycles
Fig. 1. Conditioning the test specimens

Fig. 2. Temperature variation inside the specimens during two complete F/T cycles
2.2.2

Axial Tension Test

Coupon specimens were cut from the FRP tubes followed to ASTM D638-1, Tensile
properties of Plastics. The tensile coupons dimensions were prepared according the

679

specification of the standard to provide an adequate gripping area at each end. The width of
specimen at the grip length was more than that of the free length. Figure 3.a shows the
typical dimensions of coupon specimens for different types of FRP tubes. Tests were
performed in the MTS universal testing machine under load control. The rate of loading
was approximately equal to 3 to 4 kN/min. High pressure hydraulic wedge grips were used
to hold the specimens in positions. The grip surfaces were deeply serrated with pattern
similar to those of a coarse single-cut file, serrations about 2 mm apart and about 1.5 mm
deep. The specimens were instrumented with an extensometer at the middle of the
specimens. Figure 3.b shows the test setup and instrumentation.
2.2.3

Axial Compression Test

19

13

Uniaxial compression tests followed to the ASTM D695-02 were performed in the MTS
universal testing machine under displacement control. The compression coupons
dimensions were prepared according to the specification of the standard. Figure 4.a shows
the typical dimensions of coupon specimens for different types of FRP tubes. Special steel
fixtures were prepared to prevent the buckling of specimens during the test, see Figure 4.b.
To reduce the friction between the sample and the steel fixtures, Teflon sheet was inserted
between them during the test. The assembly (fixtures and sample) was placed between the
jaws of the machine. The specimens were instrumented with an extensometer at the middle
of the specimens. The rate of displacement was approximately equal to 0.5 mm/min. Test
setup is shown in Figure 4.c.

57
165 mm

24

19

(a) Specimen dimension


(b) Tests setup
Fig. 3. Test procedures for coupon tension test

50
255 m m

(a) Specimen dimension


(b) Test fixture used
Fig. 4. Test procedures for coupon compression test

680

(c) Tests setup

3. RESULT AND DISSECTIONS


3.1

Room Temperature Specimens

Failure of coupon specimens under axial tension load started with matrix cracking and was
followed by fiber rupture in the longitudinal direction. Sound snapping could be heard with
increasing the load, which attributed to cracking of the resin. Failure was always sudden,
with a burst rupture of fibers almost always at one end of the coupons. Figure 5.a presents
the dominant failure mechanisms observed for coupon tension test specimens. The average
stress-strain relationships for five specimens are presented in Figure 5.b, at the first stage of
loading the curve was linear up to 80% of the peak load. Beyond this level the curve was
nonlinear up to failure. Small load drops accompanied by the change in the stiffness were
observed. This resulted from the earlier rupture of fibers and matrix cracking. The ultimate
average tensile stress for five specimens was 60.13 MPa, while the average axial strain at
the peak stress was 0.0069. From the measured axial strains and stresses, the elastic
modulus in the axial direction of the different specimens was determined. The value of the
average elastic modulus was 12434 MPa.
Compression coupon specimens of different specimens failed in identical manner. The
failure occurred at one end of the narrow width of the specimen in the axial direction. The
axial compression force produced shear stresses in the axial direction of the specimens.
However, interlock or overlap was observed at the failure area between the fracture lines of
the two faces. On the other hand, it was observed slightly local buckling for thin specimens
after failure. Figure 6.a shows the dominant failure mechanisms observed for coupon
compression test specimens. The average stress-strain responses of coupons in compression
are shown in Figure 6.b for the five types of FRP tubes. The response is generally linear in
the first stage of loading up to 60% of the peak load. After that shows nonlinear responses
until reaching the peak load. The stress-strain curve presents plastic plateau after reaching
the peak stress. The ultimate average compressive stress for five specimens was 90. 3 MPa,
while the average axial strain at the peak stress was 0.019.

a. Dominant failure mechanisms

b. Stress-strain curve for the coupon tensile test

Fig. 5. Failure mode and stress-strain curve for coupon tensile test

681

a. Dominant failure mechanisms

b. Stress-strain curve for the coupon compression


test
Fig. 6. Failure mode and stress-strain curve for coupon compression test

3.2

Effects of Freeze-Thaw Cycles

Little research has been conducted on the isolated performance of fibre reinforced polymers
when exposed to cold climates. Most studies that have considered thermal cycling of
carbon and glass FRPs those fiber properties such as tensile strength and stiffness are not
significantly affected by the environmental exposures (Dutta 1989).
In this section the test results of all specimens were compared to the reference coupon
specimens kept at room temperature. Table 2 presents the test results of coupon tensile
specimens kept in room temperature and environmental chamber under different conditions
for 100 and 300 cycles. The results are presented in terms of average stress and strain for
five specimens with the standard deviation (SD) for all groups. A significant increase 25,
44 and 32% in the tensile strength of coupon specimens that conditioned to 100 F/T cycles
at dry air and submerged in fresh and salt water, respectively, was observed. The
corresponding values after 300 cycles showed more increase in the tensile strength,
respectively, 28, 71 and 47%. However, an adverse effect was observed for the average
strain for all specimens. The average strain values were decreased by 24, 16 and 25% after
100 cycles at dry air and submerged in fresh and salt water, respectively. The
corresponding reduction values after 300 cycles were 23, 26 and 47%, respectively.
Table 3 presents the test results of coupon compression specimens kept in room
temperature and environmental chamber under different conditions for 100 and 300 cycles.
The test result indicated that an insignificant increase in the axial compressive capacity 1.5,
9 and 4% were observed for specimens kept in dry air and submerged in fresh and salt
water, respectively, for 100 cycles. The specimens conditioned to 300 cycles showed
similar behaviour, where the corresponding values were 5, 24 and 10%, respectively. On
the other hand, the average strain values showed a reduction 30 and 37% for specimens
kept in dry air for 100 and 300 cycles, respectively. The corresponding values for
specimens kept at fresh water were 6% reduction and 23% an increase. The specimens kept
submerged in salt water to 100 cycles showed an increase 21% in the axial strain, while
specimens conditioned to 300 cycles presented a decrease 10%.

682

The test results indicated that the combined environmental cycles such as short and long
term freeze-thaw cycles in dry air, fresh and salt water did not show a significant adverse
effect on the mechanical axial behavior of the FRP tubes. This was probably because the
adverse effect, such as plasticization of matrix and micro-cracking at matrix-fiber interface,
induced by the low temperature during the freeze-thaw cycles, compromised the positive
effect, such as matrix hardening effect.
Table 2. Test results of coupon tensile specimens

Group (1)
22.5C
Group (2)
Thaw in dry
air
Group (3)
Thaw in
fresh water
Group (4)
Thaw in salt
water

Average
SD
Average

100 F/T cycles


Stress
Strain
(MPa)
(mm/mm)
60.13
0.0069
8.4
0.0008
75.4
0.0066

300 F/T cycles


Stress
Strain
(MPa)
(mm/mm)
60.13
0.0069
8.4
0.0008
77.15
0.0067

SD

17.5

0.0045

5.91

0.006

Difference %*

+25

-24

+28

-23

Average

86.47

0.0073

103.19

0.0064

SD

10.91

0.0029

8.46

0.001

Difference %*

+44

-16

+71

-26

Average

79.72

0.0065

88.50

0.0046

SD

14.02

0.0022

14.54

0.0008

Difference %*

+32

-25

+47

-47

*as compared to the average results of Group No. 1


Table 3. Test results of coupon compression specimens
100 F/T cycles
Stress
Strain
(MPa)
(mm/mm)
Average
-91.30
-0.0190
Group (1)
22.5C
SD
2.56
0.003
Average
-92.72
-0.0119
Group (2)
SD
3.91
0.0072
Thaw in dry
air
Difference %*
+1.5
-37
Average
-99.54
-0.0179
Group (3)
SD
13.45
0.0103
Thaw in
fresh water
Difference %*
+9.0
-6.0
Average
-94.88
-0.0230
Group (4)
SD
10.35
0.0180
Thaw in salt
water
Difference %*
+4
+21
*as compared to the average results of Group No. 1

300 F/T cycles


Stress
Strain
(MPa)
(mm/mm)
-91.30
-0.0190
2.56
0.003
-95.31
-0.0134
5.59
0.0023
+5.0
-30
-113.38
-0.0234
7.71
0.0110
+24
+23
-100.05
-0.0171
7.21
0.0011
+10
-10

4. CONCLUSIONS
The study presented herein intended to examine experimentally the short and long term
effects of freeze-thaw cycling on the axial tensile and compressive behavior of filament
wound FRP tubes. Seventy coupon specimens were conditioned in three different
conditions, dry air, fresh and salt water. In summary, regardless of the type of the short and
683

long term freeze-thaw cycles (100 and 300 freeze-thaw cycles), the freeze-thaw cycles have
increased the average axial tensile and compressive strength of the FRP tubes. The possible
reason for the increase axial strength could be attributed to the matrix-hardening effect
resulting from the extremely low temperature during the freeze-thaw cycles.
5. ACKNOWLEDGEMENTS
The research reported in this paper was partially sponsored by the Natural Sciences and
Engineering Research Council of Canada (NSERC). The authors also acknowledge the
contribution of the Canadian Foundation for Innovation (CFI) for the infrastructure used to
conduct testing. Special thanks to the FRE Composites Inc, QC, Canada, for providing the
FRP tubes.
6. REFERENCES
1.
2.
3.
4.
5.
6.
7.
8.

ASTM. 2008a. Standard test method for apparent hoop tensile strength of plastic or
reinforced plastic pipe by split disk method. D 2290-08, West Conshohocken, Pa.
ASTM. 2008b. Standard test method for tensile properties of plastics. D638-08, West
Conshohocken, Pa.
Dutta, P.K. 1989. Fiber composite materials in Arctic environments. Proceedings of the
sessions related to structural materials at Structures Congress, ASCE, San Francisco,
California, USA, pp. 216-225.
Fam, Z.A., and Rizkalla, S. 2002. Flexural behavior of concrete-filled fiber reinforced
polymer circular tubes. Journal of Composites for Construction, ASCE, 6(2): 2332.
Mohamed, H.M. 2010. Axial and flexural behavior of reinforced concrete-filled fiber
reinforced polymer tubes: experimental and theoretical studies. Ph.D. thesis. University
of Sherbrooke, Sherbrooke, QC, Canada.
Mohamed, H. and Masmoudi, R. 2010. Axial load capacity of reinforced concrete-filled
FRP tubes columns: experimental versus theoretical predictions. Journal of Composites
for Construction, ASCE, 14(2): 231-243.
Mohamed, H. and Masmoudi, R. 2008. Compressive behaviour of reinforced concrete
filled FRP tube. ACI Special Publications (SP), SP-257-6, pp. 91-109.
Mirmiran, A., and Shahawy, M. 1997. Behavior of concrete columns confined by fiber
composites. Journal of Structural Engineering, 123(5): 583590.

684

AUTHOR INDEX / INDEX DES AUTEURS

Coelho, M. ........................................ 253


Crawford, K. ...................................... 167
Cree, D. ............................................. 151

A
Abdelrahman, K. ............................... 275
Ahmed, E.A. ............................. 207, 637
Aidoo, J. ............................................ 365
Alam, M.S. ....................................... 603
Al-Bayati, N. .................................... 337
Al-Hammoud, R. .............................. 375
Ali, O. ............................................... 501
Ali, S.M. ........................................... 547
Al-Mahaidi, R. .................................. 109
Almusallam, T.H. ............................. 659
Al-Salloum, Y.A. .............................. 659
Alsayed, S. ........................................ 659
Alves, J.R. ........................................... 71
Aqel, M. ............................................ 659
Asprone, D. ....................................... 245

D
Dai, J.G. .............................................. 37
De Castro, J. ...................................... 357
Demers, M. .......................................... 91
Dolan, C.W. ...................................... 101
Donchev, T. ....................................... 159
Drouin, B. ......................................... 345
Kulikowski, A. .................................. 563
Dulude, C. ......................................... 637
Durante, M. ....................................... 245
E
El-Gamal, S.E. .................................. 659
El-Hacha, R. ................. 55, 275, 575, 627
El-Ragaby, A.A.................................... 71
El Refai, A. ....................................... 213
El-Salakawy, E..................................... 71
El-Sayed, M. ..................................... 337
El-Zefzafy, H. ..................... 513,649,677

B
Bakis, C.E. ................................... 63, 401
Barboura, S. ...................................... 285
Barros, J. ........................................... 253
Basbagill, J.P. ................................... 547
Baverel,O. ......................................... 285
Bellakehal, H. ................................... 593
Bnichou, N. ...................................... 151
Benmokrane, B. .. 27,81,139,309,637,667
Benzarti, K. ....................................... 443
Bigaud, D. ......................................... 501
Bijlaard, F.S.K. ................................. 555
Bisby, L.A. ........................................ 151
Bouazza, A. ...................................... 109
Bouhicha, M. ............................ 453, 593
Boulfiza, M. ............................... 263, 461

F
Fam, A. ...................... 139, 237, 425, 667
Ferracuti, B. ...................................... 327
Ferrier, E. ........................... 129, 175, 501
Flood, J. ............................................. 401
Freddi, F. ........................................... 443
Fyfe, E.R. .......................................... 409
G
GangaRao, H.V.S............................... 485
Gao, W.Y. ........................................... 37
Gentry, T.R. ....................................... 221
Goulet, S. .......................................... 637
Green, M.F. ........................................ 151

C
Carloni, C. ........................................ 195
Caron, J.-F. ................................ 203, 285
Carrigan, L. ....................................... 109
Caspary, A.D ............................... 11, 185
Ceroni, F. .......................................... 327
Chataigner, S. ................................... 443
Chen, D. ............................................ 575
Chowdhury, E.U. ............................... 151

H
Hamelin, P. ................................ 129, 175
Hamilton, H.R. .................................. 101
Hanna, M. ......................................... 425
Harries, K.A. ................................... 1,365
685

Hassan, A. ......................................... 337


Hillman, J. ........................................ 469

Mufti, A.A............................................ 27
Myers, J.J. ............................................ 19

I
Ibrahim, H.......................................... 391

N
Neale, K.W........................................... 91
Newhook, J. ........................................ 27
Nikravan, N. ...................................... 337
Nishizaki, I. .......................................... 91
Nuncy, N. ............................................ 19

J
Jabri, S.M. ......................................... 619
Jacques, E. ........................................ 477
Jaipuriar,A. ................................. 63, 401
Jeong, Y. ..................................... 63, 401
Jiang, X. ............................................ 555
Jimenez, T.T. .................................... 539
Johnson, D.T.C. ................................ 185
Juette, B. ..................................... 47, 417

O
Oh, H. ................................................. 611
Oudah, F. ........................................... 627
Ow, M.C. .......................................... 585
P
Panigrahi, S. ...................................... 461
Parvin, A. .................................. 523, 563
Pearson, M.C.K. ................................. 159
Pecce, M. ........................................... 327
Percival, J.C. ..................................... 539
Prota, A. ............................................ 245

K
Kamal, A.S.M. .................................. 461
Keller, T. .................................... 229, 357
Koay, Y.C. ........................................ 433
Kolstein, H. ....................................... 555
Kotelnikova-Weiler, N. ............ 203, 285
L
Labossire, P........................................ 91
Latour, G. .......................................... 345
Lepech, M.D. ..................................... 547
Li, H. ................................................. 493
Limbachiya, M. ................................ 159
Lloyd, A. ........................................... 299
Lopez, M.M. ............................... 63, 401
Louie, J. ............................................ 337

Q
Quayyum, S. ..................................... 531
Quek, J. ............................................. 585
Quiertant, M. ..................................... 443

M
Mahfouz, I. ....................................... 391
Marshall, C, ...................................... 603
Mashrik, M.A. .................................... 55
Masmoudi, R. .... 453, 513, 593, 649, 677
Mathieson, H. ................................... 237
Mazzotti, C. .............................. 119, 195
McCullagh, M. .................................. 409
Mohamadian, M. ............................... 619
Mohamed, H.M. ....... 345, 513, 649, 677
Mohammed, T.A. ............................. 523
Monnell, J. ........................................ 365
Montaigu, M. ...................................... 81
Moon, D.Y. ........................................ 611
Moussa, O. ........................................ 357

S
Saatcioglu, M. ........................... 299, 477
Sarfaraz, R. ........................................ 229
Savoia, M. ......................... 119, 195, 327
Sayed-Ahmed, M. ............................. 337
Sena-Cruz, J. ..................................... 253
Sennah, K. ................................. 337, 417
Sentry, M. ......................................... 109
Shahrooz, B.M ....................................... 1
Sharbatadar, M.K. ............................. 619
Sheikh, S.A. ................................. 11, 185
Si Larbi, A.................................. 129, 175
Silva, L. ............................................. 253
Sim, J. ................................................ 611
Sizemore, J. ....................................... 365
Soudki, K. .................................. 213, 375

R
Reddy, M.P.K. .................................. 485
Robert, M. ............................ 81, 139, 667
Rteil, A. ...................................... 531, 603

686

Stafford, M. ........................................ 11
Subramaniam, K.V. .......................... 195
Svecova, D. ........................................ 319

W
Wang, C. ........................................... 493
Watson, R. ......................................... 409
Weber, A. ............................. 47, 383, 417
West, J. .............................................. 213
Witt, C. ................................. 47, 383, 417
Woltman, G. ...................................... 425

T
Tadros, G. ............................................ 27
Tanner, J.E. ....................................... 101
Teng, J.G. ........................................... 37
Tikka, T.K. ....................................... 299
Tomiyama, T. ..................................... 91
Tomlinson, D.G. ............................... 425
Topper, T. .......................................... 375

X
Xian, G. .............................................. 493
Xiao, B. ............................................. 493
Z
Zaidi, A. ..................................... 453, 593

V
Vassilopoulos, A.P. .................. 229, 357
Vijay, P.V. ......................................... 485
Vogel, H.M. ................................ 27, 319

687

You might also like