You are on page 1of 417

DEBONDING FAILURE IN CONCRETE MEMBERS

RETROFITTED WITH CARBON FIBRE REINFORCED


POLYMER COMPOSITES

by

HUY BINH PHAM (BE (Hons))

A thesis submitted in fulfilment of the requirements for the degree of Doctor


of Philosophy

DEPARTMENT OF CIVIL ENGINEERING


MONASH UNIVERSITY

September 2005
STATEMENT

STATEMENT

I hereby declare that this thesis contains no material accepted for any other
degree or diploma in any university. To the best of my knowledge, this thesis contains no
material previously written or published by another person, except where due
acknowledgment is made in the text.

___________________________
HUY BINH PHAM

iii
ACKNOWLEDGMENTS

ACKNOWLEDGMENTS

First of all, I would like to express my great appreciation to my supervisor,


Dr. Riadh Al-Mahaidi. His advice, encouragement and support thought out my research
have been invaluable. I would also like to thank Dr. Geoff Taplin for his help during my
first year of research.

I wish to thank the staff at the Monash University Civil Engineering Laboratories.
The completion of the experimental aspect would not be possible without the hard work
and assistance of Graeme Rundle, Kevin Nievaart, Jeffrey Doddrell, Glenn Davis, Peter
Dunbar, Roy Goswell, Don McCarthy and Alan Taylor. I would also like to thank
Andrew Sarkady and MBT (Australia) Pty. Limited for providing partial sponsoring for
the materials required for this research.

My gratitude also goes to Jennifer Manson for her great help during my whole
time at Monash University. I am also grateful to Associate Professor Jay Sanjayan for his
advice on the reliability study. Thanks also to my colleagues and friends for their support
and friendship.

Finally, I would like to thank my family for their constant encouragement through
out the course of my life.

v
SUMMARY

SUMMARY

The research described in this thesis is concerned with a strengthening approach


using bonded fibre reinforced polymer composites (FRP). This approach has shown
significant advantages compared to traditional methods, mainly due to the outstanding
mechanical properties of the composites. However, the retrofitted members are often
subjected to new, often brittle, failure modes due to separation or debonding of the
composites from the members. In this research, the short-term debonding behaviour of
shear-lap blocks and RC beams bonded with carbon fibre reinforced polymer composites
(CFRP) using the wet lay-up method is investigated.

Literature reviews and assessment studies of the existing strength models of the
two member types, shear-lap specimens and retrofitted beams, were conducted. It is
shown that while the shear-lap strength models are able to give reasonable predictions,
the debonding strength models for retrofitted beams generally produce inaccurate and
scattered results and some models are not based on the actual mechanism and therefore
lack grounding.

The experimental and numerical work on the behaviour of shear-lap specimens


under tensile loading established two failure modes, interfacial debond and shear-tension
failure. It was found that when the CFRP is relatively thin and the CFRP bond length is
reasonably long, the first mode is dominant. It was also demonstrated that a nonlinear
bond-slip relationship can be used to approximately describe the bond behaviour.

The experimental and numerical work on the testing of RC beams bonded with
CFRP on their soffit showed two main debonding modes, end debond and intermediate
span debond. It was found that the behaviour of retrofitted beams depends a great deal on
the CFRP stiffness, the CFRP bond length and the steel reinforcement amount. The
investigation also demonstrated the effectiveness of steel clamps and U-straps for
anchorage.

A relatively simple design method for predicting the capacity of a retrofitted


beam was developed. The method is based on the beam theory, the maximum FRP strain

vii
SUMMARY

and the maximum concrete shear stress along the tension rebar layer. The design method
was verified with the results from the experimental and numerical work carried out in this
research as well as a large database of tests available in the literature. It was found that
the method can predict the strength of a retrofitted beam with good accuracy.

viii
TABLE OF CONTENTS

TABLE OF CONTENTS
I

STATEMENT III

ACKNOWLEDGMENTS V

SUMMARY VII

TABLE OF CONTENTS IX

LIST OF FIGURES XV

LIST OF TABLES XXXI

NOTATION AND ABBREVIATION XXXIII

CHAPTER 1 - INTRODUCTION 1

1.1 General background 1


1.2 Background to FRP strengthening 2
1.3 Research problem and aims 4
1.4 Thesis outline 5

PART I: INITIAL INVESTIGATION

CHAPTER 2 - LITERATURE REVIEW 7

2.1 Introduction 7
2.2 Shear-lap specimens under tensile testing 7
2.2.1 Experimental work 7
2.2.2 Prediction formulae for bond strength 13
2.3 Retrofitted beams under bending 20
2.3.1 Experimental investigations 20
2.3.2 Existing theoretical studies for flexural failure 31
2.3.3 Existing theoretical studies for intermediate span debond 32
2.3.4 Existing theoretical studies for end debond 43
2.3.5 Design guidelines 56

ix
TABLE OF CONTENTS

2.4 Finite element modelling 58


2.4.1 Review of concrete properties and modelling methods 59
2.4.2 Finite element modelling of shear-lap tests 64
2.4.3 Finite element modelling of beam tests 66
2.5 Summary of findings 76

CHAPTER 3 - ASSESSMENT OF PREDICTION MODELS 79

3.1 Introduction 79
3.2 Assessment of analytical studies of shear-lap specimens 79
3.2.1 Database 79
3.2.2 Validation results 81
3.3 Assessment of analytical studies of beam specimens 85
3.3.1 Introduction 85
3.3.2 Existing database and new database 86
3.3.3 Validation of beam theory 89
3.3.4 Validation of existing models to predict intermediate span
debond 90
3.3.5 Validation of existing models to predict end debond 93
3.4 Validation of design guidelines 101
3.5 Summary of findings 101

PART II: EXPERIMENTAL WORK

CHAPTER 4 - EXPERIMENTAL INVESTIGATION OF SHEAR-LAP


SPECIMENS UNDER TENSILE TESTING 103

4.1 Introduction 103


4.2 Aim and design methodology 103
4.3 Experimental set-up 104
4.3.1 Specimen details 104
4.3.2 Materials 105
4.3.3 Specimen preparation and instrumentation 107
4.3.4 Testing set-up and procedure 109
4.4 Experimental results 111
4.4.1 Failure modes and ultimate capacity 111

x
TABLE OF CONTENTS

4.4.2 Strain, stress and slip distributions 114


4.4.3 Load-slip curves and parametric study 126
4.5 Summary of findings 131

CHAPTER 5 - EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS


UNDER BENDING 133

5.1 Introduction 133


5.2 Preliminary experimental program 134
5.2.1 Beam Set-up 134
5.2.2 Applicability of existing theoretical models 138
5.3 Design methodology of the main experimental programs 140
5.4 Testing of beams in four-point bending 140
5.4.1 Beam details 140
5.4.2 Materials 143
5.4.3 Specimen preparation and instrumentation 144
5.4.4 Test set-up and procedure 146
5.4.5 Experimental results 149
5.5 Testing of beams in three-point bending 182
5.5.1 Beam details 182
5.5.2 Materials 185
5.5.3 Load set-up, specimen preparation and instrumentation 185
5.5.4 Experimental results 187
5.6 Summary of findings 204

PART III: NUMERICAL AND THEORETICAL MODELLING

CHAPTER 6 - SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS


207

6.1 Introduction and scope 207


6.2 Finite element idealisation 207
6.3 Material models 209
6.3.1 Concrete 209
6.3.2 CFRP composite and adhesive 212
6.4 Solution procedure 213

xi
TABLE OF CONTENTS

6.5 Verification of finite element model 214


6.5.1 Comparison of crack patterns 214
6.5.2 Comparison of peak loads and behavioural trends 219
6.5.3 Comparison of strain distributions and bond-slip curves 221
6.6 Local strain and stress distributions 232
6.7 Sensitivity study 236
6.7.1 Sensitivity to mesh size 236
6.7.2 Sensitivity to mesh type 239
6.7.3 Sensitivity to concrete tensile strength and fracture energy 241
6.7.4 Sensitivity to adhesive stiffness 244
6.8 Bond-slip based model 245
6.8.1 Finite element idealisation 245
6.8.2 Material models 247
6.8.3 Results 248
6.9 Summary of findings 249

CHAPTER 7 - SMEARED CRACK MODELLING OF RETROFITTED BEAMS


251

7.1 Introduction 251


7.2 Modelling of beams under four-point bending 251
7.2.1 Finite element idealisation 251
7.2.2 Material models 255
7.2.3 Solution procedure 257
7.2.4 Verification 258
7.2.5 Sensitivity study 274
7.2.6 Parametric study 278
7.3 Modelling of beams under three-point bending 281
7.3.1 Finite element idealisation and material models 281
7.3.2 Verification 284
7.4 Summary of findings 289

CHAPTER 8 - DISCRETE/SMEARED CRACK MODELLING OF SHEAR-LAP


SPECIMENS AND RETROFITTED BEAMS 291

8.1 Introduction and scope 291

xii
TABLE OF CONTENTS

8.2 Brief review on concrete crack models available in MERLIN 292


8.2.1 Localised failure of concrete 292
8.2.2 Distributed failure of concrete 296
8.3 Modelling of shear-lap specimens under testing 296
8.3.1 Finite element idealisation 296
8.3.2 Material models 298
8.3.3 Solution procedure 300
8.3.4 Verification of finite element model 301
8.3.5 Sensitivity study 307
8.4 Modelling of retrofitted beams under four-point bending 310
8.4.1 Finite element idealisation 310
8.4.2 Material models 312
8.4.3 Solution procedure 313
8.4.4 Verification of finite element model 314
8.4.5 Sensitivity study 325
8.5 Summary of findings 327

CHAPTER 9 - DEBOND STRENGTH MODELS AND RELIABLITY ANALYSIS


FOR RETROFITTED BEAMS 329

9.1 Introduction and scope 329


9.2 Analysis of retrofitted sections 329
9.2.1 Implementation of beam theory 329
9.2.2 Average strain in steel and FRP reinforcement 332
9.2.3 Verification and recommendation 333
9.3 Debond strength models for retrofitted beams 337
9.3.1 Model derivation 337
9.3.2 Verification against experiments 341
9.3.3 Parametric study 343
9.4 Reliability study 345
9.4.1 Introduction 345
9.4.2 Scope and procedure 347
9.4.3 Selection of variables and their variability 348
9.4.4 Monte Carlo simulations and results 355
9.4.5 Reliability 355

xiii
TABLE OF CONTENTS

9.5 Design recommendation 358

CHAPTER 10 - CONCLUSIONS 361

10.1 Concluding remarks 361


10.2 Recommendations for future work 364

REFERENCES 367

APPENDICES A-1 to D-38

LIST OF PUBLICATIONS

xiv
LIST OF FIGURES

LIST OF FIGURES

Figure 2.1 Different set-ups for shear-lap tests: a) Double pull-pull test; b) Single pull-
push test; c) Bending test .............................................................................................8
Figure 2.2 Failure modes observed in shear-lap tests.........................................................9
Figure 2.3 FRP-concrete bond specimen (a) top view (b) strain distribution along FRP
and (c) shear stress distribution along FRP (Lee, 2003)............................................10
Figure 2.4 Measured bond-slip curves for a test series by Nakaba et al. (2001) ..............12
Figure 2.5 Bond-slip curves plotted by Savoia et al. (2003) ............................................13
Figure 2.6 Deformation and stresses of a bonded joint (Yuan et al., 2001) .....................16
Figure 2.7 Bond-slip relations: (a) Linear ascending; (b) bilinear relation; and (c) linear
descending .................................................................................................................16
Figure 2.8 A bond-slip curve following Popovics’ equation............................................19
Figure 2.9 Flexural strengthening using FRP or steel plates ............................................24
Figure 2.10 Flexural strengthening using FRP fabrics/sheets ..........................................24
Figure 2.11 Observed failure modes.................................................................................25
Figure 2.12 Shear strengthening using steel plates, FRP plates or FRP sheets ................26
Figure 2.13 FRP efficiency ratio (max. reported strain / strain at rupture) versus relative
stiffness (Bonacci and Maalej, 2001) ........................................................................30
Figure 2.14 Flexural crack debond (Sebastian, 2001) ......................................................32
Figure 2.15 Similarity between the shear-lap test and the situation near a flexural crack34
Figure 2.16 Forces acting on an element of a composite beam (Niu and Wu, 2001).......35
Figure 2.17 Debonding mechanism due to flexural cracks (Niu and Wu, 2001) .............35
Figure 2.18 Niedermeier’s methodology. .........................................................................37
Figure 2.19 Maximum possible increase in tensile stress between two subsequent cracks
(fib Bulletin 14, 2001)................................................................................................38
Figure 2.20 Displacements near the tip of a shear crack ..................................................39
Figure 2.21 Differential crack mouth opening displacements causing peeling (Neubauer
and Rostasy, 1999).....................................................................................................40
Figure 2.22 A peel test (Karbhari et al., 1997) .................................................................41
Figure 2.23 Crack sliding model (Mohamed Ali, 2000)...................................................42
Figure 2.24 Predicted shear and normal stress distributions near the plate end (Roberts,
1989) ..........................................................................................................................44

xv
LIST OF FIGURES

Figure 2.25 Stresses acting on a FRP laminate element near the cut-off point (Malek et
al, 1998) .....................................................................................................................46
Figure 2.26 Definition of bond development length for retrofitted beams by Nguyen et al.
(2001) .........................................................................................................................49
Figure 2.27 Stresses acting at the cut-off point of FRP (Saadatmanesh and Malek, 1998)
....................................................................................................................................49
Figure 2.28 Teeth scenario (Zhang and Raoof, 1995) ......................................................54
Figure 2.29 Shear-lap scenario for end debond ................................................................56
Figure 2.30 Finite element mesh adopted (a) and comparison of strain distributions
reported (b) by Maeda et al. (1997) ...........................................................................64
Figure 2.31 A typical FE mesh adopted by Chen et al. (2001).........................................65
Figure 2.32 The FE mesh (above) and the crack patterns for two example specimens with
bond lengths of 50 mm (below left) and 130 mm (below right) as reported in Lee
(2003) .........................................................................................................................66
Figure 2.33 Boundary conditions and mesh discretization by Camata et al. (2003) ........66
Figure 2.34 A FE mesh adopted (a) and the stress contours near the plate end reported (b)
by Teng et al. (2002) ..................................................................................................68
Figure 2.35 FE discretization model (a), loading curve (b) and FRP stress distributions
(c) as reported by Niu and Wu (2001) .......................................................................69
Figure 2.36 A crack pattern at failure reported by Ross et al. (1999)...............................70
Figure 2.37 A typical two-dimensional model of retrofitted beams adopted by Rahimi
and Hutchinson (2001)...............................................................................................71
Figure 2.38 Node and element displacements for RC beam model with slip in plate by
Aprile et al. (2001) .....................................................................................................73
Figure 2.39 Comparison of load-deflection curves for Zarnic et al.’s beams by Wong and
Vecchio (2003)...........................................................................................................73
Figure 2.40 The three-dimensional model of FRP plated beams adopted by Wong et al.
(2001) .........................................................................................................................74
Figure 2.41 Initial FE mesh (left) and crack pattern at a high load level (right) reported
by Yang et al. (2003)..................................................................................................75
Figure 3.1 Validation results for Van Gemert’s model (a) and ........................................81
Figure 3.2 Validation results for Maeda et al.’s model (a) and Izumo et al.’ model (b) ..82
Figure 3.3 Validation results for Niedermeier’s model (a), Neubauer and Rostasy’s
model (b) and Chen and Teng’s model (c) ................................................................83

xvi
LIST OF FIGURES

Figure 3.4 Validation results for Kanakubo et al.’s model (a) and Ulaga and Vogel’s
model (b)....................................................................................................................84
Figure 3.5 Validation of the beam theory .........................................................................90
Figure 3.6 Validation results for Arya and Farmer’s model (a) and Shehata et al.’s model
(b)...............................................................................................................................91
Figure 3.7 Validation results for Maruyama and Ueda’s model (a) and Teng et al.’s
model (b)....................................................................................................................92
Figure 3.8 Validation results for Blaschko (1997) model (a) and Mohamed Ali (2000)
model (b)....................................................................................................................92
Figure 3.9 Validation results for Niedermeier’s model ...................................................93
Figure 3.10 Validation results for Shehata et al.’s model.................................................94
Figure 3.11 Validation results for Nguyen et al.’s model (a) and Fanning and Kelly’s
model (b)....................................................................................................................94
Figure 3.12 Validation results for Saadatmanesh and Malek’s model (a), Tumialan et
al.’s model (b) and El-Mihilmy and Tedesco’s model (c) .........................................95
Figure 3.13 Validation results for Oehlers’ model (a) and Smith and Teng’s model (b) 96
Figure 3.14 Validation results for Jansze’s model (a) and Ahmed and Van Gemert’s
model (b)....................................................................................................................97
Figure 3.15 Validation results for Ziraba et al.’s model ..................................................97
Figure 3.16 Validation results for Raoof and Hassanen’s model – lower limit (a) and
Chaallal et al.’s model (b)..........................................................................................98
Figure 3.17 Validation results for Neubauer and Rostasy’s model (a) and Neubauer and
Rostasy’s model with Arya and Farmer’s limit (b) ...................................................99
Figure 3.18 Validation of the methods recommended in ACI 440.................................101
Figure 4.1 Specimen dimensions and test set-up............................................................104
Figure 4.2 MBrace FRP system components and main tools used for installation ........106
Figure 4.3 Locations of strain gauges on CFRP shear-lap specimens............................109
Figure 4.4 Loading and gripping method (a) and a specimen before testing (b)............110
Figure 4.5 Failure modes ................................................................................................112
Figure 4.6 Typical crack surfaces in interfacial debond failure .....................................113
Figure 4.7 A typical crack surface in shear-tension failure ...........................................113
Figure 4.8 Layers of the composite ................................................................................115
Figure 4.9 Strain distributions along bond length in specimen T1a (Lf = 60 mm).........116
Figure 4.10 Strain distributions along bond length in specimen T2a (Lf = 80 mm).......116

xvii
LIST OF FIGURES

Figure 4.11 Strain distributions along bond length in specimen T3a (Lf = 100 mm).....117
Figure 4.12 Strain distributions along bond length in specimen T4a (Lf = 140 mm).....117
Figure 4.13 Strain distributions along bond length in specimen T5a (Lf = 180 mm).....117
Figure 4.14 Strain distributions along bond length in specimen T6a (Lf = 220 mm).....118
Figure 4.15 Strain distributions along bond length in specimen T7a (H = 5 mm) .........118
Figure 4.16 Strain distributions along bond length in specimen T8a (H = 70 mm) .......119
Figure 4.17 Strain distributions along bond length in specimen T12a (bf = 50 mm) .....119
Figure 4.18 Strain distributions along bond length in specimen T13a (bf = 70 mm) .....120
Figure 4.19 Strain distributions along bond length in specimen T9a (Lf = 100 mm).....120
Figure 4.20 Strain distributions along bond length in specimen T10a (Lf = 140 mm)...121
Figure 4.21 Strain distributions along bond length in specimen T11a (Lf = 180 mm)...121
Figure 4.22 Average bond stress distributions along bond length in specimen T4a ......122
Figure 4.23 Average bond stress distributions along bond length in specimen T5a ......122
Figure 4.24 Average bond stress distributions along bond length in specimen T6a ......123
Figure 4.25 Bond stress versus slip development in specimen T4a (bf = 100 mm) .......124
Figure 4.26 Bond stress versus slip development in specimen T5a (bf = 100 mm) .......124
Figure 4.27 Bond stress versus slip development in specimen T6a (bf = 100 mm) .......125
Figure 4.28 Bond stress versus slip development in specimen T8a (bf = 100 mm) .......125
Figure 4.29 Fitted bond-slip relationships for specimens with two layers of CFRP of 100
mm width .................................................................................................................125
Figure 4.30 Bond stress versus slip development in specimen T12a (bf = 50 mm) .......126
Figure 4.31 Bond stress versus slip development in specimen T13a (bf = 70 mm) .......126
Figure 4.32 Load versus slip curves for specimens with two layers of CFRP and variable
bond lengths .............................................................................................................127
Figure 4.33 Load versus slip curves for specimens with six layers of CFRP and variable
bond lengths (subjected to alignment errors)...........................................................127
Figure 4.34 Effect of bond length on bond capacity.......................................................128
Figure 4.35 Load-slip curves for specimens with different CFRP thicknesses T3a (2
plies) and T9a (6 plies).............................................................................................128
Figure 4.36 Effect of CFRP thickness on bond capacity ................................................129
Figure 4.37 Load-slip curves for specimens with different CFRP widths T3a (100 mm),
T12a (50 mm) and T13a (75 mm)............................................................................129
Figure 4.38 Effect of CFRP width on bond capacity......................................................130

xviii
LIST OF FIGURES

Figure 4.39 Load-slip curves for specimens with different support clearances T3a (110
mm), T7a (5 mm) and T8a (70 mm) ........................................................................130
Figure 4.40 Effect of varying support clearance on bond capacity ................................131
Figure 5.1 Preliminary beam dimensions and test set-up ...............................................135
Figure 5.2 Load-displacement curves of two trial beams ...............................................135
Figure 5.3 Beams B1 and B2 after failure ......................................................................136
Figure 5.4 CFRP strain distributions ..............................................................................137
Figure 5.5 Average bond stress distributions in shear span............................................137
Figure 5.6 Photos of beams B1 (a) and B2 (b) near failure ............................................138
Figure 5.7 Applicability of theoretical models to predict IS debond loads for beam B1
.................................................................................................................................139
Figure 5.8 Applicability of theoretical models to predict end debond loads for beam B2
.................................................................................................................................139
Figure 5.9 Typical rectangular beam cross-sectional details (beams E1a and E1b).......141
Figure 5.10 Typical beam longitudinal details (beams E1a and E1b) ............................141
Figure 5.11 A steel clamp ...............................................................................................143
Figure 5.12 Surface texture after water jetting (a) and after priming (b) .......................144
Figure 5.13 Bonding operation (a) Applying a layer of resin; (b) Placing fibres on the
resin; (c) Pressing fibres into the resin with a ripped roller; (d) Finished surface...145
Figure 5.14 Strain gauge locations on beam E1a............................................................146
Figure 5.15 Load and deflection measurements .............................................................146
Figure 5.16 Test rig.........................................................................................................148
Figure 5.17 A support (a) and an anchorage of the reaction frame column (b)..............149
Figure 5.18 Load-deflection behaviour of the control beams.........................................150
Figure 5.19 Photos of beams C1a and C1b after failure .................................................150
Figure 5.20 Load-deflection behaviour of E beams .......................................................151
Figure 5.21 Load-deflection behaviour of beams E1a and E1b .....................................152
Figure 5.22 Photos of the debonded side of E1 beams ...................................................152
Figure 5.23 High-speed video captures for beam E1a at failure (end debond) (rate: 500
frames per second) ...................................................................................................153
Figure 5.24 End cover peeling failure surface in beam E1b...........................................154
Figure 5.25 Influence of stirrup extrusion on debond cracks in beam E1a ....................154
Figure 5.26 Load-deflection behaviour of beams E2a and E2b .....................................155
Figure 5.27 Photos of the debonded side of E2 beams ...................................................155

xix
LIST OF FIGURES

Figure 5.28 Load-deflection behaviour of beams E3a and E3b......................................156


Figure 5.29 Photos of the debonded side of E3 beams ...................................................156
Figure 5.30 Load-deflection behaviour of beams E4a and E4b......................................157
Figure 5.31 Photos of the debonded side of E4 beams ...................................................157
Figure 5.32 Load-deflection behaviour of beams E5a and E5b......................................158
Figure 5.33 Photos of the debonded side of E5 beams ...................................................158
Figure 5.34 Load-deflection behaviour of S beams........................................................159
Figure 5.35 High-speed video captures of beam S1a at failure (combination of end and
intermediate span debond) (rate: 500 frames per second) .......................................160
Figure 5.36 Photos of the debonded side of beams S1a and S1b....................................161
Figure 5.37 Full loading curves of beams S2a and S2b..................................................162
Figure 5.38 High-speed video captures of beam S2b at failure (intermediate span
debond) (rate: 500 frames per second).....................................................................163
Figure 5.39 Photos of the debonded side of beams S2a and S2b....................................164
Figure 5.40 Failure surface in beam S2a.........................................................................164
Figure 5.41 Photos of the debonded side of beams S3a and S3b....................................165
Figure 5.42 Loading curves for all beams tested until concrete crushed........................165
Figure 5.43 Crack patterns in control beams ..................................................................167
Figure 5.44 Crack patterns in S beams ...........................................................................167
Figure 5.45 Crack patterns in E beams ...........................................................................168
Figure 5.46 Tensile steel strain distributions in beam C1a .............................................171
Figure 5.47 Stirrup strain development in beam C1a .....................................................171
Figure 5.48 CFRP strain distributions in beam E1a on unclamped side.........................172
Figure 5.49 Bond stress distributions in beam E1a on unclamped side..........................172
Figure 5.50 CFRP strain development in beam E1a on unclamped side........................172
Figure 5.51 Tensile steel strain distributions in beam E1a on unclamped side ..............173
Figure 5.52 Stirrup strain development in beam E1a on unclamped side.......................173
Figure 5.53 CFRP strain distributions in beam S1a on unclamped side.........................174
Figure 5.54 Bond stress distributions in beam S1a on unclamped side..........................174
Figure 5.55 CFRP strain development in beam S1a on unclamped side ........................174
Figure 5.56 Tensile steel strain distributions in beam S1a on unclamped side ..............175
Figure 5.57 Stirrup strain development in beam S1a on unclamped side.......................175
Figure 5.58 CFRP strain distributions in beam S2a on unclamped side.........................176
Figure 5.59 Bond stress distributions in beam S2a on unclamped side..........................176

xx
LIST OF FIGURES

Figure 5.60 CFRP strain development in beam S2a on unclamped side ........................176
Figure 5.61 Tensile steel distributions in beam S2a on unclamped side ........................177
Figure 5.62 Stirrup strain development in beam S2a on unclamped side.......................177
Figure 5.63 Effect of CFRP and tension steel stiffness ratio..........................................179
Figure 5.64 Effect of CFRP bond length ........................................................................179
Figure 5.65 Crack patterns on the clamped side of E1a and S1a....................................181
Figure 5.66 Non-prestressed U strap system ..................................................................184
Figure 5.67 Prestressed U strap system ..........................................................................184
Figure 5.68 Non-prestressed U-strap system..................................................................186
Figure 5.69 Prestressed U-strap system..........................................................................186
Figure 5.70 Load and deflection measurement for beams tested in three-point bending
.................................................................................................................................187
Figure 5.71 Strain gauge locations on U-straps..............................................................187
Figure 5.72 Load-deflection behaviour of retested beams .............................................188
Figure 5.73 Load-deflection behaviour of beams A1a, A1b, A2a and A2b ...................188
Figure 5.74 Photo of the debonded side of beam E3b2 ..................................................189
Figure 5.75 Photos of the debonded side of beams anchored with a non-prestressed U
strap..........................................................................................................................190
Figure 5.76 Sliding of the longitudinal CFRP and local debonding and rupture of a strap
leg in beam E3a........................................................................................................190
Figure 5.77 Photos of the debonded side of beams anchored with a prestressed U strap
.................................................................................................................................191
Figure 5.78 Debonding of a leg of the prestressed strap in beam A1a ...........................191
Figure 5.79 Photos of beams anchored with three non-prestressed U straps after failure
.................................................................................................................................192
Figure 5.80 Photos of beams anchored with three prestressed U straps after failure .....193
Figure 5.81 Crack patterns in beam E3b2.......................................................................194
Figure 5.82 Crack patterns in beams with one U-strap...................................................194
Figure 5.83 Crack patterns in beams with three U-straps...............................................195
Figure 5.84 Longitudinal CFRP strain distributions in beam A1a .................................197
Figure 5.85 Longitudinal CFRP strain development in beam A1a.................................197
Figure 5.86 Strain development in the U-strap legs in beam A1a ..................................197
Figure 5.87 Longitudinal CFRP strain distributions in beam A1b .................................198
Figure 5.88 Longitudinal CFRP strain development in beam A1b ................................198

xxi
LIST OF FIGURES

Figure 5.89 Strain distributions along U strap height in beam A1b at different load levels
..................................................................................................................................198
Figure 5.90 Strain development in the U-strap legs in beam A1b..................................199
Figure 5.91 Strain development in the tension reinforcement in beam A1b ..................199
Figure 5.92 Longitudinal CFRP strain distributions in beam A2a .................................200
Figure 5.93 Longitudinal CFRP strain development in beam A2a.................................200
Figure 5.94 Strain development in the U-strap legs in beam A2a ..................................200
Figure 5.95 Strain development in the tension reinforcement in beam A2a...................201
Figure 5.96 Longitudinal CFRP strain distributions in beam A2b .................................202
Figure 5.97 Longitudinal CFRP strain development in beam A2b.................................202
Figure 5.98 Strain distributions along U strap height in beam A2b at different load levels
..................................................................................................................................202
Figure 5.99 Strain development in the U-strap legs in beam A2b..................................203
Figure 5.100 Strain development in the tension reinforcement in beam A2b ................203
Figure 6.1 Two-dimensional mesh of a typical specimen (T3) ......................................208
Figure 6.2 CQ12M element (a) and CQ16M element (b)...............................................209
Figure 6.3 Loading-unloading for total strain crack models (de Witte and Kikstra, 2003)
..................................................................................................................................210
Figure 6.4 Concrete stress-strain curve in tension (a) and compression (b) ...................212
Figure 6.5 Modified Newton-Raphson iteration (de Witte and Kikstra, 2003) ..............213
Figure 6.6 Deformed shapes of model T5 at different load levels (magnification factor =
60) ............................................................................................................................215
Figure 6.7 Deformed shapes of model T1 at different load levels (magnification factor =
60) ............................................................................................................................216
Figure 6.8 Crack patterns in models T1 to T6 ................................................................217
Figure 6.9 Crack patterns in models T7 and T8..............................................................217
Figure 6.10 Crack patterns in models T9, T10 and T11 .................................................217
Figure 6.11 Crack patterns in models T12 and T13........................................................217
Figure 6.12 Correlation of peak loads predicted by the FE model and measured in the
experiment................................................................................................................219
Figure 6.13 Effect of varying CFRP bond length (Experiment and FE) ........................220
Figure 6.14 Effect of varying support block height H (Experiment and FE) .................220
Figure 6.15 Effect of varying CFRP thickness (Experiment and FE) ............................221
Figure 6.16 Effect of varying CFRP width (Experiment and FE) ..................................221

xxii
LIST OF FIGURES

Figure 6.17 Comparison of CFRP strain distributions in specimens T1a, T2a and T3a 223
Figure 6.18 Comparison of CFRP strain distributions in specimens T4a, T5a and T6a 224
Figure 6.19 Comparison of CFRP strain distributions in specimens T7a and T8a ........225
Figure 6.20 Comparison of CFRP strain distributions in specimens T9a, T10a and T11a
.................................................................................................................................226
Figure 6.21 Comparison of CFRP strain distributions in specimens T12a and T13a ....227
Figure 6.22 Comparison of load-slip curves for specimens T1a to T6a.........................228
Figure 6.23 Comparison of load-slip curves for specimens T9a, T10a and T11a ..........229
Figure 6.24 Comparison of load-slip curves for specimens T7a, T8a, T12a and T13a..229
Figure 6.25 Comparison of local bond-slip relationships in specimen T5a (bf = 100 mm,
Lf = 180 mm) ...........................................................................................................230
Figure 6.26 Comparison of local bond-slip relationships in specimen T6a (bf = 100 mm,
Lf = 220 mm) ...........................................................................................................231
Figure 6.27 Comparison of local bond-slip relationships in specimen T12a (bf = 50 mm,
Lf = 100 mm) ...........................................................................................................231
Figure 6.28 Comparison of local bond-slip relationships in specimen T13a (bf = 70 mm,
Lf = 180 mm) ...........................................................................................................232
Figure 6.29 Gaussian integration points in the CFRP and adhesive elements ...............233
Figure 6.30 CFRP strain distributions in model T5........................................................234
Figure 6.31 CFRP strain distributions in model T11......................................................235
Figure 6.32 Prediction of CFRP strains near the loaded end at different load levels using
different mesh sizes .................................................................................................237
Figure 6.33 Comparison of crack patterns after peak in specimens T5 (left) and T1 (right)
predicted by the models with different mesh sizes ..................................................238
Figure 6.34 Comparison of load-slip curves predicted by the models with different mesh
sizes..........................................................................................................................238
Figure 6.35 Different meshes used .................................................................................239
Figure 6.36 Comparison of crack patterns after peak in specimens T5 (left) and T1 (right)
predicted by the models with different mesh types .................................................240
Figure 6.37 Comparison of load-slip curves predicted by the models with different mesh
types .........................................................................................................................240
Figure 6.38 Comparison of crack patterns after peak in specimens T5 (left) and T1 (right)
predicted by the models with different Gf values ....................................................242

xxiii
LIST OF FIGURES

Figure 6.39 Comparison of load-slip curves predicted by the models with different Gf
values .......................................................................................................................242
Figure 6.40 Comparison of crack patterns after peak in specimens T5 (left) and T1 (right)
predicted by the models with different fct values .....................................................243
Figure 6.41 Comparison of load-slip curves predicted by the models with different fct
values .......................................................................................................................243
Figure 6.42 Comparison of crack patterns after peak in specimens T5 (left) and T1 (right)
predicted by the models with different Ea values.....................................................245
Figure 6.43 Comparison of load-slip curves predicted by the models with different Ea
values .......................................................................................................................245
Figure 6.44 Two-dimensional mesh of a typical specimen (T3) ....................................246
Figure 6.45 L7BEN element ...........................................................................................247
Figure 6.46 3+3 node structural interface element (CL12I) ...........................................247
Figure 6.47 Idealised bond-slip relationships .................................................................247
Figure 6.48 Correlation of peak loads predicted by the bond-slip based model and
measured in the experiment .....................................................................................248
Figure 6.49 Comparison of the load-slip curves of specimens T5a and T1a..................249
Figure 6.50 Comparison of CFRP strain distributions of specimens T5a and T1a ........249
Figure 7.1 Two-dimensional mesh of beams loaded in 4-point bending........................251
Figure 7.2 Q8MEM element (a) and L7BEN element (b) ..............................................252
Figure 7.3 2+2 node structural interface element (L8IF)................................................252
Figure 7.4 Steel and CFRP reinforcement locations in model E1 ..................................253
Figure 7.5 Reduced concrete element width...................................................................253
Figure 7.6 Modelling of the steel clamp .........................................................................254
Figure 7.7 Concrete shear strength after cracking ..........................................................255
Figure 7.8 Bond-slip relationships adopted for steel bar/concrete interface (a) and
CFRP/concrete interface (b).....................................................................................257
Figure 7.9 Quasi-Newton iteration (de Witte and Kikstra, 2003)...................................258
Figure 7.10 Crack pattern in model C1...........................................................................259
Figure 7.11 Crack patterns in E1 (unclamped side)........................................................260
Figure 7.12 Crack patterns in model S2 (clamped side) .................................................261
Figure 7.13: Crack patterns in model S1 (unclamped side).............................................262
Figure 7.14 Deformed shapes of the unclamped side of model E1 near the peak load
(magnification factor = 20) ......................................................................................262

xxiv
LIST OF FIGURES

Figure 7.15 Deformed shapes of the clamped side of model S2 near the peak load
(magnification factor = 20) ......................................................................................263
Figure 7.16 Deformed shapes of the unclamped side of model E1 near the peak load
(magnification factor = 20) ......................................................................................263
Figure 7.17 Induced tooth-like crack..............................................................................264
Figure 7.18 Flexure-shear crack debond.........................................................................264
Figure 7.19 Crack patterns in S models at the peak loads ..............................................265
Figure 7.20 Crack patterns in E models (unclamped side) at the peak loads .................266
Figure 7.21 Correlation of peak loads as predicted by the FE model and measured in the
experiment ...............................................................................................................266
Figure 7.22 Comparison of load-displacement curves for S beams ...............................267
Figure 7.23 Comparison of load-displacement curves for E beams ...............................268
Figure 7.24 Gauss integration points in a CFRP beam element .....................................269
Figure 7.25 Variation of CFRP strain along the top and bottom halves in model S2N .269
Figure 7.26 Comparison of CFRP strain distributions in beam E1a ..............................270
Figure 7.27 Comparison of CFRP strain distributions in beam S2a...............................270
Figure 7.28 Comparison of CFRP strain distributions in beam S1a...............................271
Figure 7.29 Comparison of tensile steel strain distributions in beam E1a .....................272
Figure 7.30 Comparison of tensile steel strain distributions in beam S2a......................272
Figure 7.31 Comparison of tensile steel strain distributions in beam S1a......................273
Figure 7.32 Comparison of crack patterns at the peak loads in beams E1 (a) and S2C (b)
predicted by models with different Gf values ..........................................................275
Figure 7.33 Comparison of load-deflection curves predicted by models with different Gf
values .......................................................................................................................276
Figure 7.34 Comparison of crack patterns at the peak loads in beams E1 (a) and S2C (b)
predicted by models with different fct values...........................................................276
Figure 7.35 Comparison of load-deflection curves predicted by models with different fct
values .......................................................................................................................277
Figure 7.36 Comparison of crack patterns at the peak loads in beams E1 (a) and S2C (b)
predicted by models with different β values............................................................278
Figure 7.37 Comparison of load-deflection curves predicted by models with different β
values .......................................................................................................................278

xxv
LIST OF FIGURES

Figure 7.38 Behavioural trend of retrofitted beams with different CFRP cross sectional
areas .........................................................................................................................279
Figure 7.39 Behavioural trend of retrofitted beams with different tension reinforcement
cross sectional areas .................................................................................................280
Figure 7.40 Behavioural trend of retrofitted beams with different CFRP bond lengths.280
Figure 7.41 Behavioural trend of retrofitted beams with different concrete covers .......281
Figure 7.42 Mesh of a beam loaded in three-point bending ...........................................282
Figure 7.43 Modelling of reinforcements .......................................................................283
Figure 7.44 Location of interface elements in the FE mesh for model A2a ...................284
Figure 7.45 Crack patterns in retested beams at peak.....................................................285
Figure 7.46 Crack patterns in A beams at peak ..............................................................286
Figure 7.47 Deformed shapes after peak of beams with three straps (magnification factor
= 10) .........................................................................................................................286
Figure 7.48 Correlation of peak loads as predicted by the FE model and measured in the
experiment................................................................................................................287
Figure 7.49 Comparison of load-deflection curves for retested beams ..........................287
Figure 7.50 Comparison of load-deflection curves for A beams....................................288
Figure 7.51 Comparison of longitudinal CFRP strain distributions in beam A1a..........289
Figure 7.52 Comparison of strain development on U-strap sides in beam A1a .............289
Figure 8.1 Hillerborg’s Fictitious Crack Model (Saouma, 2002) ...................................293
Figure 8.2 Interface idealization and notations (Saouma, 2002) ....................................293
Figure 8.3 Failure surface (Saouma, 2002).....................................................................294
Figure 8.4 Bi-linear softening laws (Saouma, 2002) ......................................................294
Figure 8.5 Stiffness degradation in the equivalent uniaxial case (Saouma, 2002) .........295
Figure 8.6 Initial crack locations in specimens T1-T6....................................................297
Figure 8.7 Final transverse crack paths in specimens T1-T6..........................................297
Figure 8.8 Two typical meshes of the shear-lap specimens............................................298
Figure 8.9 Three-node plane stress element (a) and four-node interface element (b) ...298
Figure 8.10 Deformed shapes of model T5 at different load levels (magnification factor =
60) ............................................................................................................................302
Figure 8.11 Deformed shapes of model T1 at different load levels (magnification factor =
60) ............................................................................................................................303
Figure 8.12 Discrete crack opening in models T1 to T6 at the peak loads (magnification
factor = 60)...............................................................................................................304

xxvi
LIST OF FIGURES

Figure 8.13 Comparison of load-slip behaviours between the experiments, FEA using
smeared crack modelling and FEA using discrete/smeared crack modelling .........305
Figure 8.14 Comparison of CFRP strain distributions in specimens T1a, T2a and T3a
(exp. vs discrete/smeared crack model) ...................................................................306
Figure 8.15 Comparison of CFRP strain distributions in specimens T4a, T5a and T6a
(exp. vs discrete/smeared crack model) ...................................................................307
Figure 8.16 Comparison of load-slip curves predicted by the models with different GIf
values .......................................................................................................................308
Figure 8.17 Comparison of load-slip curves predicted by the models with different fct
values .......................................................................................................................308
Figure 8.18 Comparison of load-slip curves predicted by the models with different GIIf
values .......................................................................................................................309
Figure 8.19 Discrete crack locations in model E1 ..........................................................310
Figure 8.20 A typical mesh of retrofitted beams ............................................................311
Figure 8.21 Steel clamp location in model S2C .............................................................312
Figure 8.22 Deformed shape of the unclamped side of beam E1 at the peak load
(magnification factor = 40) ......................................................................................314
Figure 8.23 Deformed shape of the clamped side of beam S2C at the peak load
(magnification factor = 30) ......................................................................................315
Figure 8.24 Deformed shape of the unclamped side of beam S1N at the peak load
(magnification factor = 30) ......................................................................................316
Figure 8.25 Deformed shapes of the unclamped side of other E models at the peak loads
(magnification factor = 30) ......................................................................................317
Figure 8.26 Deformed shapes of other S models at the peak loads (magnification factor =
30) ............................................................................................................................318
Figure 8.27 Comparison of load-displacement curves for E beams ...............................320
Figure 8.28 Comparison of load-displacement curves for S beams ...............................321
Figure 8.29 Comparison of CFRP strain distributions in beam E1a (exp. vs
discrete/smeared crack model).................................................................................322
Figure 8.30 Comparison of CFRP strain distributions in beam S2a (exp. vs
discrete/smeared crack model).................................................................................322
Figure 8.31 Comparison of CFRP strain distributions in beam S1a (exp. vs
discrete/smeared crack model).................................................................................323

xxvii
LIST OF FIGURES

Figure 8.32 Comparison of tensile steel strain distributions in beam E1a (exp. vs
discrete/smeared crack model).................................................................................323
Figure 8.33 Comparison of tensile steel strain distributions in beam S2a (exp. vs
discrete/smeared crack model).................................................................................324
Figure 8.34 Comparison of tensile steel strain distributions in beam S1a (exp. vs
discrete/smeared crack model).................................................................................324
Figure 8.35 Comparison of load-displacement curves predicted by the models with
different GIf values ...................................................................................................325
Figure 8.36 Comparison of load-displacement curves predicted by the models with
different fct values ....................................................................................................326
Figure 8.37 Comparison of load-displacement curves predicted by the models with
different GIIf values ..................................................................................................327
Figure 9.1 Strain distribution before (a) and after installation (b) of FRP......................330
Figure 9.2 Flow chart for implementation of sectional analysis.....................................331
Figure 9.3 Local variations in steel strain and bond stress .............................................332
Figure 9.4 Calculation model for a RC member under pure flexure (Comite Euro-
International du Beton, 1985) ..................................................................................333
Figure 9.5 Sectional analysis results along beam E1a at maximum load level ..............334
Figure 9.6 Sectional analysis results along beam S1a at maximum load level...............335
Figure 9.7 Comparison of FRP strain distributions between experimental results and
predictions by beam theory in beam E1a .................................................................336
Figure 9.8: Comparison of FRP strain distributions between experimental results and
predictions by beam theory in beam S1a .................................................................336
Figure 9.9 Comparison of measured and calculated CFRP strains at the ultimate load
level for retrofitted beams without anchorage .........................................................337
Figure 9.10 Failure mechanisms .....................................................................................337
Figure 9.11 Average shear stress concept.......................................................................339
Figure 9.12 Variation of maximum average bond stress in beams failed by end debond
(a) and maximum tensile force in CFRP in beams failed by intermediate span
debond or mixed debond (b) ....................................................................................340
Figure 9.13 Prediction of end and intermediate span failure loads for the beams tested in
the present study ......................................................................................................342
Figure 9.14 Comparison between the theoretical predictions and experimental results.343
Figure 9.15 Comparison of the predicted trends with different CFRP thicknesses........344

xxviii
LIST OF FIGURES

Figure 9.16 Comparison of the predicted trends with different tensile steel amounts ...344
Figure 9.17 Comparison of the predicted trends with different CFRP bond lengths .....345
Figure 9.18 Comparison of the predicted trends with different concrete covers............345
Figure 9.19 Cross sections of T beams used...................................................................349
Figure 9.20 Side view of T beams and loading positions...............................................350
Figure 9.21 Design point and reliability index according to the first order reliability
method for normally distributed uncorrelated variables (Eurocode 2, 2002)..........356

xxix
LIST OF TABLES

LIST OF TABLES

Table 1.1 Typical properties of fibres (fib Bulletin 14, 2001) ............................................3
Table 2.1 Experimental studies on beams strengthened in flexure...................................22
Table 2.2 Different predictions for concrete tensile strength ...........................................61
Table 2.3 Different predictions for fracture energy..........................................................63
Table 3.1 Summary of the database of shear-lap tests......................................................80
Table 3.2 Variability of test parameters and results in the shear-lap test database ..........80
Table 3.3 Summary of validation results of strength models for shear-lap tests..............85
Table 3.4 Summary of the database of beam tests............................................................88
Table 3.5 Variability of some test parameters in the beam test database .........................89
Table 3.6 Summary of validation results of debond models for beams .........................100
Table 4.1 Variables in the experimental program...........................................................105
Table 4.2 Mechanical properties of MBrace FRP system as given by the manufacturer
.................................................................................................................................106
Table 4.3 Measured elastic modulus of the composite...................................................107
Table 4.4 Shear-lap test results .......................................................................................114
Table 5.1 Mechanical properties of materials used ........................................................134
Table 5.2 Variables in experimental program No. 1.......................................................142
Table 5.3 Mechanical properties of steel reinforcement.................................................144
Table 5.4 Summary of beam test results.........................................................................166
Table 5.5 Maximum CFRP strains and bond stresses ....................................................178
Table 5.6 Variables in experimental program No. 2.......................................................183
Table 5.7 Experimental results .......................................................................................194
Table 5.8 Maximum U-strap leg strains .........................................................................204
Table 6.1 Concrete properties adopted in the finite element model ...............................212
Table 6.2 Comparison between the experimental and the FEA results ..........................218
Table 6.3 Mesh size variations .......................................................................................236
Table 6.4 Mesh type variations.......................................................................................239
Table 6.5 Concrete tensile property variations ...............................................................241
Table 6.6 Adhesive modulus variations..........................................................................244
Table 6.7 The parameters defining Popovics’ bond-slip curve ......................................248
Table 7.1 Concrete properties used in the finite element model ....................................256

xxxi
LIST OF TABLES

Table 7.2 Steel properties used in the finite element model ..........................................256
Table 7.3 Concrete tensile property variations ...............................................................274
Table 7.4 Shear retention factor variations .....................................................................277
Table 8.1 CFRP and adhesive material model parameters ..............................................299
Table 8.2 Concrete smeared crack model parameters for shear-lap blocks....................299
Table 8.3 Interface crack model parameters for shear-lap blocks ..................................300
Table 8.4 Mode-II fracture energy variations .................................................................309
Table 8.5 Concrete smeared crack model parameters for S beams ................................312
Table 8.6 Interface crack model parameters for S beams ...............................................313
Table 8.7 Steel model parameters ...................................................................................313
Table 9.1 Design summary of T beam configurations....................................................350
Table 9.2 Calculation of model errors ............................................................................351
Table 9.3 Statistical parameters for some random variables from CONTECVET (2002)
..................................................................................................................................352
Table 9.4 Statistical parameters for FRP ultimate strength from Atadero et al. .............354
Table 9.5 Summary of variable statistical properties......................................................354
Table 9.6 Simulation results for the shear resistance R ..................................................355
Table 9.7 Capacity reduction factors corresponding to β = 3.25....................................358
Table 9.8 Capacity reduction factors recommended by AS5100....................................358

xxxii
NOTATION AND ABBREVIATION

NOTATION AND ABBREVIATION

Af cross-sectional area of a FRP composite


Al cross-sectional area of a laminate
As cross-sectional area of tension reinforcement
Asv cross sectional area of shear reinforcement
A’s cross-sectional area of compression reinforcement
a shear span of a beam loaded in four-point bending; or
constant in Popovics’ equation
ba width of an adhesive layer
bc, b width of a concrete member
bf width of a FRP composite
bfl width of the top flange of a T-beams
bl width of a laminate
bv effective width of a web for shear
bw width of a web
c cover to reinforcing steel
d effective depth of a cross-section
d0 effective depth of a cross-section (beam shear context)
da maximum aggregate size
db diameter of a reinforcement bar
dc depth of a concrete shear block
df distance from the top fibre of the concrete to the centroid of tension
reinforcement
dn distance from the top fibre of the concrete to the neutral axis of a cross-
section
ds distance from the top fibre of the concrete to the centroid of tension
reinforcement
d’s distance from the top fibre of the concrete to the centroid of compression
reinforcement
Ea modulus of elasticity of adhesive
Ec modulus of elasticity of concrete
Ef modulus of elasticity of a FRP composite

xxxiii
NOTATION AND ABBREVIATION

El modulus of elasticity of a laminate


Es modulus of elasticity of tension steel reinforcement
Esv modulus of elasticity of shear reinforcement
E’s modulus of elasticity of compression steel reinforcement
Ff force in a FRP composite
Ff.max maximum force in a FRP composite
Fs force in tension reinforcement
F’s force in compression reinforcement
Fc force in the concrete compressive area of a cross-section
fc compressive strength of concrete
fcm mean value of the compressive strength of concrete at the relevant age
fcu compressive cube strength of concrete
fct tensile strength of concrete
fcv shear strength of concrete
fcv,d shear strength of the end peeling debond surface
ft tensile strength
f’c characteristic compressive cylinder strength of concrete
f’ct characteristic principle tensile strength of concrete
fint internal force (finite element context)
ffu tensile strength of a FRP composite
fsy yield strength of tension reinforcement steel
f’sy yield strength of compression reinforcement steel
fsy,f yield strength of shear reinforcement steel
Ga shear modulus of an adhesive
Gf shear stiffness of FRP reinforcement; or
Gf, GIf specific mode-I fracture energy of concrete
GIIf specific mode-II fracture energy of concrete
Gs shear stiffness of steel reinforcement
H support block clearance of a shear-lap specimen
h total depth of a cross-section
Ic second moment of area of a concrete beam cross-section about the
centroidal axis
If second moment of area of a FRP composite cross-section about the
centroidal axis

xxxiv
NOTATION AND ABBREVIATION

Itr,f cracked second moment of area of a beam cross-section transformed to


FRP
Itr,c uncracked second moment of area of a beam cross-section transformed to
concrete
ku neutral axis parameter, being the ratio, at ultimate strength, of the depth to
the neutral axis from the extreme compression fibre to ds
L total beam span
Le effective bond length
Lf FRP bond length on concrete block or FRP bond length on the shear span
of a beam
L0 distance from a support to the nearest end of the bonded FRP composite
lb anchorage length of a FRP composite bonded on a beam
M bending moment
M0 bending moment at the cut-off point of a bonded laminate
Mcal calculated maximum bending moment that a beam can support
Mcr flexural cracking bending moment of a cross-section
Mexp actual maximum bending moment that a beam can support
Mu,0 ultimate bending moment capacity of a beam cross-section without FRP
Mu,frp ultimate bending moment capacity of a beam cross-section retrofitted with
FRP
nf number of FRP layer(s)
ns number of tension reinforcement bar(s)
P, Pexp actual maximum force that a single shear-lap can take
Pcal calculated maximum force that a single shear-lap can take
q external distributed load applied on a concrete beam
R resistance (reliability context)
s centre-to-centre spacing of shear reinforcement
s1 slip at which the bond stress reaches its maximum value (bond-slip curve
context)
s2 maximum slip of a linear bond-slip curve
srm spacing of cracks in a beam
ta thickness of an adhesive layer
tf total thickness of a FRP composite
tf,0 thickness of a layer of FRP

xxxv
NOTATION AND ABBREVIATION

tl thickness of a laminate
uf FRP-to-concrete average bond strength
us steel-to-concrete average bond strength
*
V design shear force
V0 shear force at the cut-off point of a bonded laminate
Vcal calculated maximum shear force that a beam can support
Vexp actual maximum shear force that a beam can support
Vdebond actual debonding shear strength of a beam
Vmax, Vu,no debond calculated ultimate shear strength of a retrofitted beam failing by
concrete crushing or rupture of the laminate
Vu,end debond ultimate shear strength of a retrofitted beam failing by end debond
Vu,IS debond ultimate shear strength of a retrofitted beam failing by intermediate span
debond
yf distance from the neutral axis to the centroid of FRP reinforcement
α calibration factor for the maximum FRP tensile force in a retrofitted beam
failing by intermediate span debond; or
ratio of the FRP stiffness to the concrete stiffness (EfAf / EcAc); or
model error (reliability context)
β shear retention factor for concrete (finite element context); or
reliability index (reliability context)
∆L distance between two strain gauge positions
∆u0 displacement increment (finite element context)
δui ith iterative displacement increment (finite element context)
ε, εc concrete strain
ε0 concrete strain at which the concrete stress reaches its maximum value
(concrete stress-strain curve context)
εcu concrete crushing strain
εf FRP reinforcement strain
εfu FRP ultimate strain
εs tension steel reinforcement strain
εsy yield strain of tension steel reinforcement
ε’s compression steel reinforcement strain
ζ distribution coefficient

xxxvi
NOTATION AND ABBREVIATION

µ mean value
µR mean of the resistance (reliability context)
νc Poisson’s ratio of concrete
σ stress; or
standard deviation (reliability context)
σ0 normal stress at a laminate cut-off point
σR standard deviation of the resistance (reliability context)
ΣObar total perimeter of tension reinforcing bars
ρf FRP reinforcement ratio
ρs steel reinforcement ratio
τ average bond stress between two gauge positions
τ0 shear stress at a laminate cut-off point
τave average shear stress on the end peeling debond surface
τmax maximum average bond stress between two positions
τult interface shear strength
φ strength reduction factor
AFRP aramid fibre reinforced polymer composites
CFRP carbon fibre reinforced polymer composites
FCD flexural crack debond
FE finite element
FEA finite element analysis
FRP fibre reinforced polymer composites
FSCD flexure-shear crack debond
GFRP glass fibre reinforced polymer composites
IS debond intermediate span debond
RC reinforced concrete
SCD shear crack debond

xxxvii
INTRODUCTION

CHAPTER 1 - INTRODUCTION

1.1 General background


Nowadays, upgrading civil engineering structures is a very important issue,
particularly for concrete bridges and buildings. For concrete bridges, increase in traffic
load and deterioration caused by environmental factors are the two main reasons for a
decrease in safety. For example, in the United States, almost 40 % of highway bridges are
classified as deficient and in need of rehabilitation or replacement (Chen and Duan,
2000). In Australia, more than 50 % of bridges were built prior to 1948 and most of these
bridges are deemed to be inadequate, particularly those already damaged (Al-Mahaidi,
2002). For concrete buildings, there is also a significant need to repair or strengthen
structural members. This strengthening is to address problems with concrete aging and
changes of function or to improve seismic resistance (Fukuyama, 1999; Hollaway and
Leeming, 1999).

There are a number of strengthening solutions: adding members, adding supports,


increasing member cross sections, post-tensioning or bonding external laminates. Among
these methods, laminate bonding has received much attention. The first type of laminate
bonding is steel plate bonding, which has been used since the 1960s. The principle of this
method is to externally bond steel plates on reinforced concrete members as these plates
take tension in the same way as internal steel reinforcement. This method has proved to
be effective in increasing members’ stiffness and ultimate load capacity. However, it has
some drawbacks, such as the possible corrosion of the plates and the difficulties that arise
from the high self-weight of the plates.

Recently, a second type of laminate bonding has been developed, which is fibre
reinforced polymer composite bonding or FRP bonding. This system makes use of a new
excellent material for construction: fibre reinforced polymer composite (FRP). The new
system possesses many advantages over the steel plate system such as high strength of
laminates, easy installation, excellent durability and minimum maintenance requirement.
FRP bonding can be used to strengthen not only reinforced concrete structures but also
masonry or timber structures (Bakis et al., 2002). Although it is a relatively new

1
CHAPTER 1

technique, FRP bonding has been used to strengthen numerous bridges and buildings
worldwide as well as in Australia in the last two decades. The Westgate Bridge
strengthening project in Melbourne Australia is one excellent example of this.

1.2 Background to FRP strengthening


FRP strengthening can be defined as a system in which FRP composites are fixed
externally to a structural element to enhance its load bearing capacity. There are a number
of systems available. They differ mainly in the composite material types and the methods
of bonding. A brief description of the materials involved and possible bonding methods is
presented below.

A fibre reinforced polymer composite consists of fibres embedded in a matrix. The


main function of the fibres is to carry load, while the matrix holds the fibres together and
protects them from mechanical and environmental factors. The properties of the
composite depend on the properties of the fibres, the matrix and their compositions. The
composite properties can be selected for particular applications. In general, FRP
composites exhibit high strength, light weight and good corrosion resistance (Rizkalla,
2002).

Fibres are polymeric materials. They are very small in diameter (for example
carbon fibres are about 8 µm in diameter). The most common types used in construction
are carbon, glass and aramid fibres. Carbon fibres, which can have very high modulus,
can be produced from two main sources: polyacrylonitrile fibres (PAN) and mesophase
pitch. Carbon fibres from the first source are used in construction because they have
higher strength and modulus. Glass fibres, which have lower modulus, are generally
cheaper. Three most popular types of glass fibres are E-glass, S-glass and C-glass.
Aramid fibres are synthesised from aromatic polyamides. Their modulus values lie
between those of carbon and glass fibres (Hull and Clyne, 1996).

The main mechanical properties of different fibres are summarised in Table 1.1.
Carbon fibres have high strength and elastic modulus, excellent fatigue properties, and
excellent moisture and chemical resistance. Glass fibres have lower strength and are
susceptible to alkaline attack, while aramid fibres can be prone to moisture, temperature,

2
INTRODUCTION

UV and alkaline attack (Rizkalla, 2002). Thus, for strengthening purposes, carbon fibres
are possibly the most suitable material.

Table 1.1 Typical properties of fibres (fib Bulletin 14, 2001)


Elastic modulus Tensile strength
Fibre Type
(GPa) (MPa)
High strength 215-235 3500-4800
Ultra high strength 215-235 3500-6000
Carbon High modulus 350-500 2500-3100
Ultra high modulus 500-700 2100-2400
E 70 1900-3000
Glass
S 85-90 3500-4800
Low modulus 70-80 3500-4100
Aramid
High modulus 115-130 3500-4000

The fibres are bonded to a structure using an epoxy resin by an adhesion process.
The epoxy resin normally contains a resin, a hardener and additives. The adhesion
process to a concrete surface generally involves several steps. At first, the epoxy flows
over and into the irregularities and wets the surface. It is then attached to the surface by
molecular forces and solidifies to form the joint. Due to the similar nature of fibres and
epoxy resins, good adhesion is normally assured. A good bond between concrete and
resins is more difficult. The bond mechanism between these two materials is more
complex involving not only mutual atomic/molecular attraction but also mechanical
interlocking (Hollaway and Leeming, 1999).

The adhesion can be implemented using two main methods: adhesive bonding or
wet lay-up. In the first method, the composite is prefabricated and then bonded onto the
concrete substrate using an adhesive. In the wet lay-up method, the composite and bond
are formed at the same time by impregnating layers of fabrics with the resin on the
concrete surface. It is a slow application and the composites can be subjected to non-
uniform wetting and waviness. However, it has great flexibility for field applications, i.e.
the composites can have different thicknesses and sizes, and can be bonded on curved

3
CHAPTER 1

surfaces. Wet lay-up is also generally the cheapest option (Karbhari and Seible, 1999;
Karbhari, 2001).

1.3 Research problem and aims


The effectiveness of FRP strengthening method relies primarily on the bonding
between the composite and concrete element. The quality of currently used adhesives has
been proved to be adequate for transferring the bond stress. However, concrete cannot
take high shear stresses and therefore cracks in high shear regions. This often results in
separation or debonding of the composite from the concrete element.

There has been a significant amount of research on this phenomenon. A number


of experiments have been carried out and a variety of the theoretical models have been
developed. However, the failure mechanisms are still not understood fully and the
validity of the existing theoretical models has not been assessed adequately. As a result, a
reliable design method for debonding is still not available.

The general aim of this research is, therefore, to study the short-term behaviour of
concrete members bonded with CFRP composites using wet lay-up method. The specific
aims are:
• To summarise and assess various theoretical models for predicting the capacity of
a retrofitted concrete member
• To gain further understanding of the local bond characteristics between CFRP and
concrete
• To study different failure mechanisms and the effect of a number of important
parameters to a the capacity of a RC beam
• To develop and verify numerical and theoretical models for prediction of critical
failure loads
• To investigate mechanical anchorage methods for debond prevention.

The outcome from this research will be to contribute to the knowledge that
currently exists in this field so that greater confidence in the prediction of the load-
carrying capacity of retrofitted concrete members can be achieved.

4
INTRODUCTION

1.4 Thesis outline


This thesis is divided into three main parts. Part I presents the initial investigation
into the behaviour of retrofitted concrete members. It consists of two chapters. Chapter 2
presents the literature review of previous experimental studies on retrofitted concrete
members and existing theories and numerical models used to analyse the members’
behaviour. Chapter 3 reports the assessment study of the existing theoretical models in
predicting the capacity of retrofitted members. Part II presents the main experimental
work undertaken in this study. It comprises two chapters. Chapter 4 reports an
investigation into local bond characteristics between the CFRP composite and concrete
using tensile testing to failure of 26 shear-lap specimens. Chapter 5 describes 26 static
bending tests to failure of retrofitted RC beams to investigate different debonding failure
mechanisms and the influence of several critical parameters. The development of
numerical and theoretical models to describe the behaviour of the shear-lap specimens
and the RC beams under testing is presented in Part III. It consists of four chapters.
Chapter 6 and Chapter 7 present the finite element models of the shear-lap specimens and
retrofitted beams based on smeared cracking. Chapter 8 reports the finite element models
of the shear-lap specimens and retrofitted beams based on a combination of discrete and
smeared cracking. The development and verification of the new theoretical models for
predicting the capacity of a retrofitted beam and the reliability study are reported in
Chapter 9. Finally, in Chapter 10, major conclusions from this research are presented with
recommendations for future studies.

5
LITERATURE REVIEW

CHAPTER 2 - LITERATURE REVIEW

2.1 Introduction
Numerous experimental and analytical studies on concrete members bonded with
laminates have been conducted covering a number of important aspects of structural
performance and durability. In the present study, only short-term structural behaviour
under static loading of two member types, concrete shear-lap specimens and retrofitted
RC beams, is considered. The literature review presented below is devoted to
experimental, theoretical and numerical work on FRP strengthening of the two concrete
member types.

This chapter is organised into three main sections. The first presents a review of
existing experimental and theoretical studies on the characteristics of the bond between
concrete and FRP using shear-lap tests. The second is a review of experimental and
theoretical studies on retrofitted beams, while the third is a review of nonlinear finite
element modelling and its usage to simulate the behaviour of shear-lap specimens and
retrofitted beams under static loading.

2.2 Shear-lap specimens under tensile testing


2.2.1 Experimental work
There have been a number of experimental investigations on the bond between
concrete and FRP (Chajes et al., 1996; Brosens and Van Gemert, 1997; Maeda et al.,
1997; Täljsten, 1997; Izumo et al., 1998; Bizindavyi and Neale, 1999; De Lorenzis et al.,
2001; Nakaba et al., 2001; Xu et al., 2001; Dai and Ueda, 2003; Lee, 2003; Yao et al.,
2005). In these studies, there is a great variation in test configurations, laminate types,
adhesive types and specimen dimensions. There is also a variety of failure modes
reported. In this section, a summary of test set-ups and failure types is presented. It is
followed by a description of the observed bond transfer mechanism and the effect of
several parameters on the bond behaviour. Lastly, studies on the bond-slip relations are
reviewed.

7
CHAPTER 2

2.2.1.1 Test set-ups and failure modes


Figure 2.1 shows the three most common set-ups used to study the bond between
FRP and concrete: the double pull-pull test, single pull-push test and bending test. In a
double pull-pull test, two blocks of concrete placed end to end are held together by FRP
composites bonded on the two sides of the specimen. The concrete blocks are separated
by a small gap and the tensile force is applied by pulling on the reinforcement bar cast
within the concrete block. The tension can also be applied through two steel plates
bonded on the sides of the concrete blocks or by a hydraulic actuator placed between the
concrete blocks. The main advantages of this set-up are that the composite ends do not
need to be anchored and no loading rig is normally required. In a single pull-push test, the
composite is bonded to only one side of the concrete block. The load is applied to the
FRP at the free end. The concrete block is held in place with a support block bearing
against the concrete surface near the loaded end and the rotation of the block is prevented
by a steel bracket. The main advantage of this test configuration is that the FRP and
concrete material usage can be reduced. However, a more complex loading rig is
required. In the third type of testing, the FRP is bonded across a saw cut in the middle of
a beam and the tension is applied on the FRP by bending the beam in three points.

FRP
(a) steel bar concrete

FRP
(b) support concrete
block

(c) saw cut


concrete

FRP

Figure 2.1 Different set-ups for shear-lap tests: a) Double pull-pull test; b) Single
pull-push test; c) Bending test

The observed failure modes reported in the literature are shown in Figure 2.2 with
the colour lines representing the failure surfaces (Neubauer and Rostasy, 1997; Chen and
Teng; 2001 and Yao et al., 2002). Mode 1 is called interfacial debond and it is the most

8
LITERATURE REVIEW

commonly observed. The failure surface is in the concrete a few millimetres beneath the
concrete-adhesive interface. A concrete prism may also be pulled out near the loaded end.
Mode 2 is called shear-tension failure, in which the main crack propagates into the
concrete block from the FRP end. This mode tends to occur only when the laminate is
relatively thick. For modes 3, 4 and 5, the failure occurs in the laminate and adhesive
materials. Mode 3, FRP tensile rupture, can be observed if the FRP cross-sectional area is
very small. Mode 4, cohesion failure through the adhesive, has also been reported. The
composite can also be delaminated as in mode 5, FRP delamination, where the
delamination path can bridge across the adhesive layer and penetrate into the concrete.
These last three modes are not common especially for normal strength concrete, of which
the shear strength is much lower compared to that of the adhesive and FRP. The
following discussion applies for mode 1 only, which is the most widely observed failure
mode.

FRP Mode 5

Mode 3 Mode 4

adhesive

Mode 1

concrete Mode 2

Figure 2.2 Failure modes observed in shear-lap tests

2.2.1.2 Bond transfer mechanism


Despite the wide range of materials and configurations used, a general trend exists
in terms of the force transfer behaviour and failure mechanisms of a composite-concrete
joint (Chajes et al., 1996; Maeda et al., 1997; Lee, 2003). The explanation of the force
transfer mechanism is illustrated in Figure 2.3 for a fixed block bonded with a FRP
composite over a length Lf. At the early stage of loading, the strain in the FRP follows a
shape as curve ‘1’. The strain gradually decreases along the length from the highest value
near the loaded edge. As the load is increased, the strain value increases and the strain
distribution shifts upwards as indicated by curve ‘2’. The strain at the loaded end
eventually exceeds the shear strain capacity of the concrete. At this point, concrete

9
CHAPTER 2

cracking initiates under the adhesive layer near the loaded end leading to a flattening
portion of the strain curve as shown in curve ‘3’. Further increase in the load extends the
crack and gradually shifts the area of active bonding towards the unloaded end until the
whole FRP is debonded from the concrete block. This is characterised by the shift in the
shear stress distribution curve as shown in Figure 2.3c.

concrete
FRP Lf

F
(a)
ε 4
2 3
1 adhesive
F
F
(b)
τ τ
crack
F F
F F
Prior to cracking After cracking
(c)

Figure 2.3 FRP-concrete bond specimen (a) top view (b) strain distribution along
FRP and (c) shear stress distribution along FRP (Lee, 2003)

As a result of the mechanism described above, the tension is transferred to the


concrete via only a certain length. This length has a maximum limit called the effective
bond length, Le. The anchorage strength does not increase once the bond length is greater
than the effective bond length (Chajes et al., 1996; Maeda et al., 1997)

2.2.1.3 Parameters influencing bond strength


Previous experiments indicate that the effective bond length and therefore the
ultimate bond capacity are influenced by a number of factors including FRP stiffness,
concrete strength, concrete surface preparation and adhesive type.

Maeda et al. (1997) reported that the effective bond length decreased as the
stiffness of FRP increased. The stiffness was expressed as the product of the elastic

10
LITERATURE REVIEW

modulus and thickness of the FRP material. Täljsten (1997) reported that there was a
critical strain level, εcr, at which concrete began to fracture, and therefore a governing
factor for failure. It followed that for a given strain level, the higher the FRP stiffness,
the higher the load the bond could stand.

The effect of concrete strength was investigated by Chajes et al. (1996) and Xu et
al. (2001). Chajes et al. carried out bond tests with three different concrete strengths
ranging from 24 to 45 MPa and concluded that the ultimate bond strength was
proportional to f c . Xu et al. varied the concrete strength from 24 to 70 MPa and

reported that the ultimate bond strength was not always proportional to the concrete
tensile strength because different concrete strengths caused different effective bond
lengths.

The effects of concrete surface preparation and adhesive types on the bond
strength of the composite-concrete joint were investigated by Chajes et al. (1996),
Toutanji and Ortiz (2001) and Xu et al. (2001). Chajes et al. studied the bond on three
different surfaces: no surface preparation, a surface ground with a stone to give smooth
finish and a surface abraded mechanically with a wire wheel to leave the aggregate
slightly exposed. They reported that the concrete surface should be mechanically abraded
or sandblasted to achieve the best possible bond. Toutanji and Ortiz conducted a similar
study by comparing two surface preparations: water jet blasting and sanding. The
investigators reported that the surface treatment by a water jet produced higher bonding
strength than the surface treatment by a sander. Xu et al. showed from their experiments
that the bond formed by different adhesive resins had different ultimate strengths. They
also proved that the primer could improve the bond strength and the surface ground
heavily with a disk sander provides a better bond. The conclusions drawn from these
studies were similar. Surface treatments that produced a rough concrete surface produced
a better bond compared to surface treatments that resulted in a smooth concrete finish.

2.2.1.4 Bond-slip relations


As will be presented later in section 2.2.2, most recent theoretical studies are
based on an assumption that there is a relationship between the bond stress and the slip,

11
CHAPTER 2

called the bond-slip relation. This relation has been established experimentally by Nakaba
et al. (2001) and Savoia et al. (2003).

Nakaba et al. (2001) carried out a series of double pull-pull tests with the primary
test variables being the types of composites and concrete strength. The bond-slip curves
were plotted using the readings from the strain gauges placed at 15 mm spacing along the
bond length. The plot is given in Figure 2.4. The researchers concluded that composite
stiffness did not influence the shape of the bond-slip relation. The maximum local bond
stress increased as the concrete compressive strength increased. The relations were clearly
nonlinear and could be represented using Popovics’ equation (Popovics, 1973) given by
τ s a
= (2.1)
τ max s1 (a − 1) + (s / s1 ) a
Nakaba et al. recommended that the constant, a, could be taken to be 3 and the slip at
maximum bond stress, s1, could be taken as 0.065 mm.

Figure 2.4 Measured bond-slip curves for a test series by Nakaba et al. (2001)

Savoia et al. (2003) used the test results from Chajes et al. (1996) to plot the bond-
slip relationships (Figure 2.5). They also recommended that Popovics’s equation could be
used to describe the interface law. The maximum bond stress was found to be 3.5 fc0.19 at
a slip of 0.051 mm. The constant, a, was taken to be 2.86.

12
LITERATURE REVIEW

Figure 2.5 Bond-slip curves plotted by Savoia et al. (2003)

In summary, these past studies have highlighted the complexity of the failure
mechanisms of a concrete-composite joint and the factors affecting the bond’s behaviour
and strength. It is clear that some of the important issues are not yet conclusive. For
example, the effect of FRP width and thickness on the bond-slip curve and therefore the
bond performance needs to be investigated further. Also, the bond behaviour when
transverse shear-tension cracks formed within the bond line is not yet clear. Further
experimental study is, therefore, still required.

2.2.2 Prediction formulae for bond strength


Various formulations have been proposed to predict the bond ultimate capacity P
of a FRP-concrete bond. They are summarised below. In the formulae presented, the
force unit is Newton (N) and the distance is in millimetre (mm), unless noted otherwise.

2.2.2.1 Empirical formulae


An early formulation for bond capacity prediction was reported by Van Gemert
(1980). This formulation is given as:
P = 0.5b f L f f ct (2.2)
where bf is the FRP width, Lf is the bond length and fct is the pull-off tensile strength of
concrete surface. This is a simple method, in which the bond shear stress is assumed to
vary linearly from the highest value at the loaded edge over the whole bond length.
However, experimental evidence has proved that this assumption is not correct for the
composite-concrete bond surface and the ultimate load does not always increase
proportionally to the bond length.

13
CHAPTER 2

Another relatively simple equation was proposed by Chajes et al. (1996), as


follows
P = R × min(L f , L e ) (2.3)
In this equation, the ultimate capacity of the joint depends on two parameters: the average
bond resistance, R, and the bond development length, Le. The drawback of the equation is
that the bond resistance and effective bond length have to be determined directly from
experiments. From their experiments, Chajes et al. found that R = 4.94bf N/mm and Le =
95 mm.

A number of other empirical expressions were developed by researchers in Japan.


Maeda et al. (1997) developed an equation relating the average bond strength to the
stiffness of the composite and strain gradient, as follows
P = b f E f t f × (L e × strain gradient) (2.4)
The gradient was assumed to be constant. From their double pull-pull tests using carbon
fibre sheet, Maeda et al found that the strain gradient was 110.2 × 10-6. The effective
bond length, Le, was found to be dependent on the composite stiffness and its value was
calibrated with the experiments to yield the following formula:
L e = e 6.134−0.580 ln (E f t f ) (2.5)

Izumo et al. (1998) conducted a similar study and developed relationships to


determine the bond strength for a given concrete strength and bond length. The ultimate
2/3
load, P, was assumed to be proportinal to a f c + b , where a and b are material constants

obtained by calibrating the FRP strain at failure and the concrete strength. The term f c2 / 3
is related to the tensile strength of concrete as reported in Japanese Standards.
Calibrations were carried out for specimens with the bond length, Lf, larger or equal to
100 mm and the following expressions were recommended:
P = (3.8fc2/3 + 15.2) Ef Lf bf tf × 10-6 for CFRP (2.6a)
P = (3.8fc2/3 + 69.0) Ef Lf bf tf × 10-6 for AFRP (2.6b)

2.2.2.2 Semi-theoretical formulae


A number of other prediction models were developed based on fracture
mechanics. An early expression was proposed by Holzenkampfer (1994) to calculate the

14
LITERATURE REVIEW

bond strength between steel plates and concrete. It was modified by Niedermeier (1996)
for the bond between CFRP plates and concrete and given as:
⎧ 0.78k f b f E f t f f ct if L f ≥ L e

P=⎨ Lf ⎛ Lf ⎞ (2.7a)
⎪0.78k f b f E f t f f ct L ⎜⎜ 2 − L ⎟⎟ if L f < L e
⎩ e ⎝ e ⎠
where the effective bond length is given by:
Ef t f
Le = (2.7b)
4f ct

Another modification was made by Neubauer and Rostasy (1999). In their paper, a
slightly different formulation was presented for P and Le, as follows
⎧0.64k f b f E f t f f ct if L f ≥ L e

P=⎨ Lf ⎛ Lf ⎞ (2.8a)
⎪0.64k f b f E f t f f ct L ⎜⎜ 2 − L ⎟⎟ if L f < L e
⎩ e ⎝ e ⎠
where
Ef t f
Le = (2.8b)
2f ct

In general, these expressions were derived based on the assumption that the
fracture energy can be calculated as a function of the concrete tensile strength, fct, as
follows
G f = k f2 c f f ct (2.9a)
where kf is a geometrical factor related to the width of the plate, bf, and the width of the
concrete member, bc, and can be calculated using the following expression:
2 − bf / bc
k f = 1.125 (2.9b)
1 + b f / 400

where cf is an empirical constant containing all secondary effects. This constant was
determined in a linear regression analysis using the results of double shear or similar
tests. Neubauer and Rostasy reported that cf equals 0.202 mm.

A more adequately reported derivation was by Yuan et al. (2001). They analysed
the interfacial stress state of externally bonded laminates for single pull-push situation.
The stress state is illustrated in the figure below.

15
CHAPTER 2

FRP
σ1 σ1 + dσ/dx dx
τ
τ

concrete
σ2 σ2 + dσ/dx dx

dx

Figure 2.6 Deformation and stresses of a bonded joint (Yuan et al., 2001)

By assuming that the shear stress is a function of the relative displacement or slip, they
derived a differential equation relating the deformation, δ, with the interfacial fracture
energy, Gf as shown below
d 2 δ 2G f 2
− λ f ( δ) = 0 (2.10)
dx 2 τ 2max
where
τ 2max ⎛ 1 bf ⎞
λ2 = ⎜⎜ + ⎟⎟ (2.11)
2G f ⎝ Ef t f bcEc t c ⎠
Two parameters are introduced in the above expressions: the local bond strength τmax and
the interfacial fracture energy Gf. The solutions were obtained by solving the differential
equation using different assumed relations between bond shear stress and slip (bond-slip
or shear-slip relations). Three bond-slip relations considered by Yuan et al. are illustrated
in Figure 2.7.

Bond stress Bond stress Bond stress

τmax τmax τmax

Gf Gf
Gf slip slip slip
s1 s1 s2 s2

(a) (b) (c)


Figure 2.7 Bond-slip relations: (a) Linear ascending; (b) bilinear relation; and (c)
linear descending

The solution for Case (a), linear ascending with a sudden stress drop, was found to give:

16
LITERATURE REVIEW

τ max b f
P= tanh(λL f ) (2.12)
λ
The solution for Case (c), linear descending, gave:
⎧ τ max b f
⎪ if L f ≥ L e
P=⎨ λ (2.13)
τ b
⎪ max f sin (λL f ) if L f < L e
⎩ λ
π
where L e = and Gf = 0.5τmaxs2.

The solution for Case (b), bilinear shear-slip, yielded a relatively complicated implicit
expression.

A more complete solution was reported later in Wu et al. (2002). In this paper, the
solutions for both pull-pull and pull-push situations were reported. They found that when
the bilinear relationship was used and the bond length Lf is sufficiently large, the ultimate
load only depends on the interfacial fracture energy Gf or the area under the bond-slip
curve. In such cases, the following equations were derived for pull-push (single shear)
and pull-pull (double shear):
2E f t f G f
P = bf for pull-push (2.14a)
1+ α

P = b f 2E f t f G f (1 + α) for pull-pull (2.14b)

where α = Ef Af / Ec Ac.

The equations reported in Yuan et al. and Wu et al. were, however, not verified
with experimental data. A more practical formula was proposed by Chen and Teng
(2001). The researchers noticed that s1 is small compared to s2 and used the solution for
Case (c) in Figure 2.7 (Equation 2.13). They also used the findings from other
experimental studies to derive the formula for the bond strength, as follows

P = 0.427β f β L f c b f L e (2.15a)

where
Ef tf 2 − bf / bc ⎧ 1 if L f ≥ L e
Le = ;βf = ; βL = ⎨ (2.15b)
fc 1 + bf / bc ⎩sin( πL / 2L e ) if L f < L e

17
CHAPTER 2

The constant 0.427 in Equation 2.15a was obtained by calibrating with a database of
shear-lap tests. Chen and Teng showed that the expression agreed well with the test data.
The ratio of the experimental to predicted ultimate bond strength was found to have an
average of 1.00 with the corresponding coefficient of variation of 0.159.

Dai and Ueda (2003) proposed different expressions for the bond-slip curve. The
expressions used are:
λ
⎛s⎞
τ = τ max ⎜⎜ ⎟⎟ s ≤ s1 (2.16a)
⎝ s1 ⎠
τ = τ max e −β ( s −s1 ) s > s1 (2.16b)

By calibrating with in their experiments, they derived expressions to estimate λ, β, τmax


and s1, as follows
λ = 0.575 (2.16c)
β = 0.0035 K (Ef tf)0.34 (2.16d)

− 1.575αK + 2.481α 2 K 2 + 6.3αβ 2 KG f


τ max = (2.16e)

s1 = τmax / αK (2.16f)
where α = 0.028 (Ef tf)0.254 and K= Ga/ta. They also provided the following expression to
estimate the interfacial fracture energy, Gf, based on the adhesive stiffness, K, and the
concrete strength:
G f = 7.554K −0.449 f c0.343 (2.16g)
The researchers showed that the strain distributions and the load-slip curves for their test
specimens were predicted accurately using the proposed formulae.

A different bond-slip relation was used by Kanakubo et al. (2003). The relation
was based on Popovics’ models (Equation 2.1). The shape of the bond-slip plot is shown
in Figure 2.8. By calibrating with their experiments, Kanakubo et al. recommended that
a = 3, τmax = 3.5 fc0.19 and s1 = 0.065 mm. To solve for the ultimate bond strength, they
introduced a concept of ‘equivalent bond stress block’ and derived the following
formulae:
P = k e τ max b f L e (Lf ≥ Le) (2.17a)

P = kτ max b f L e (Lf < Le) (2.17b)

18
LITERATURE REVIEW

2λ f s e
Le = (2.17c)
ke

1− ke ⎛L ⎞ 1+ ke
k= cos⎜⎜ f π ⎟⎟ + (2.17d)
2 ⎝ Le ⎠ 2

where the bond length index λf = tfEf / τmax, se = 0.354 mm and ke = 0.428. It is also worth
noting that the researchers had assumed that the concrete stiffness was significantly
higher than that of the FRP. Therefore, concrete dimensions were not present in the
formulae. Kanakubo et al. also verified their model against a number of tests done in
Japan. They showed that the ratios of the experimental to predicted ultimate strength had
an average of 0.99 and a coefficient of variation of 17 %.

Bond stress

τmax

Gf
slip

s1

Figure 2.8 A bond-slip curve following Popovics’ equation

Ulaga and Vogel (2003) attempted to derive the shape of the bond-slip
relationship using micromechanical considerations. They stated that at a low load, shear
deformation is concentrated in the adhesive layer; whereas at a high load, the crack faces
are in the concrete layer and contain aggregate particles which partly protrude from the
cement matrix. Using ‘aggregate interlock theory’, they derived a bond-slip relationship
with nonlinear softening. To simplify the rather complicated relationship, they adopted a
linear softening curve with τmax = 0.4(fc)2/3. The formulae derived to predict the ultimate
capacity are:
2E f t f G f
P = bf when Lf > Le (2.18a)
(1 + α )
2
(1 + α)τ 2max L f 2E f t f G f
P = sin bf when Lf ≤ Le (2.18b)
2E f t f G f (1 + α )
where

19
CHAPTER 2

π 2E f t f G f
Le = (2.18c)
2 (1 + α)τ 2max

G f ≈ 0.045 f c2 / 3 (2.18d)
Equation 2.18a is the same as Equation 2.14a. However, Ulaga and Vogel (2003) also
derived formula for the case of a short bond length (Equation 2.18b).

In summary, there have been numerous analytical studies on the bond behaviour
between FRP and concrete. There are two main methodologies: empirical and semi-
empirical. The empirical methods are based mainly on the result of a certain set of
experiments and tend to yield simpler equations. The semi-empirical methods are based
on fracture mechanics using an idealized local bond-slip relationship. The shape of the
bond-slip curves and therefore the prediction formulae were generally calibrated with a
limited set of tests. Since, both methods relied on a certain set of experiments which
usually had similar configurations, the accuracy of these models still needs to be further
investigated.

2.3 Retrofitted beams under bending


There have been numerous experimental and theoretical studies on retrofitted
beams under bending. In this section, a review of existing experimental studies is
presented first, in which three main failure modes are identified. It is followed by reviews
of existing theoretical models of the three modes.

2.3.1 Experimental investigations


2.3.1.1 Strengthening schemes and failure modes
Table 2.1 lists the experimental studies of beams strengthened in flexure found in
the literature. The materials used for strengthening included steel, CFRP, GFRP and
AFRP. The FRP composites were either thin flexible fabrics/sheets or thick plates. The
failure modes and the variables under investigation are also included in the table.

In the listed experimental studies, most of the beams are rectangular and simply
supported. The laminates are bonded in various ways as illustrated in Figure 2.9 and
Figure 2.10. Different anchorage methods are developed using bolts, clamps, angles or
wraps.

20
LITERATURE REVIEW

A variety of failure modes have been observed (Figure 2.11). They can be
grouped as follows.

Flexural failure: This failure includes yielding of longitudinal steel bars followed by
secondary compression failure of the concrete or tensile rupture of the laminate.

Shear failure: It can occur in beams with insufficient shear capacity. The typical critical
shear crack extends from the end of the laminate to the point of loading. The crack
initiated close to the support has also been reported.

Debonding failure: Debonding is the most frequently observed mode. About two-thirds
of the specimens tested for FRP flexural strengthening failed due to laminate debonding.
More than one debonding mode exist. They can be classified into two main categories
based on the initial starting point. End debond is the failure that originates near the plate
end and propagates in the concrete either along the tension steel reinforcement (end cover
peeling) or near the bond line (end interfacial debond). Intermediate span debond is the
failure that originates either from a wide flexural crack (flexure crack debond), a wide
flexure-shear crack (flexure-shear crack debond) or a wide diagonal shear crack (shear
crack debond). The failure then propagates from the crack tip to the laminate end parallel
to the adhesive/concrete interface. Flexure-shear crack debond is also often accompanied
by cracking along the tension reinforcement layer.

In addition to these modes, failures in the adhesive or FRP layer have also been
reported. Failure can also result from the unevenness of the concrete surface. However,
these are not common and can be prevented by using proper adhesion and suitable FRP
materials and by carrying out correct surface preparation.

21
CHAPTER 2

Table 2.1 Experimental studies on beams strengthened in flexure


Failure mode3
Laminate Laminate
( \ : not clearly
material1 form2
described)
Reference Variable4
1
S C G A f p & 3 4 5 6 7 8
2
Swamy et al. (1987) x x x x x x ta, tl
Jones et al. (1988) x x x x x anc
Swamy et al. (1989) x x x ta, pre
Ritchie et al. (1991) x x x x x x mat, anc, ta
Saadatmanesh and Ehsani (1991) x x x x x As, pre, she
Oehlers (1992) x x x x x a, L0, she
Triantafillou et al. (1992) x x \ \ pre
Sharif et al. (1994) x x x x x tl, anc
Hussain et al. (1995) x x x x x tl, anc, pre
Vichit-Vadakan (1995) x x x x x x she, L, sur
Quantrill et al. (1996a) x x x x x x x fc, ta, Al, anc
Quantrill et al. (1996b) x x x x x x Al, bl/tl, mat, anc
Garden et al. (1997) x x x x x bl/tl, a/D, anc
Arduini et al. (1997) x x x x x x \ \ El, tl
Arduini and Nanni (1997) x x x x x tl, pre
Norris et al. (1997) x x \ \ fibre orientation
Takahashi et al. (1997) x x x -
He et al. (1997) x x x x El, anc
Saafi and Buyle-Bodin (1997) x x x tl
Garden et al. (1998) x x x x x a/d
Spadea et al. (1998) x x \ anc
GangaRao and Vijay (1998) x x x x \ pre, side wrap
Tumialan et al. (1999) x x x x L, she, anc, bl, tl
Ahmed and van Gemert (1999) x x x \ x tl
Swamy and Mukhopadhyaya (1999) x x x \ \ As, fc, she, anc
Ross et al. (1999) x \ x As
Grace et al. (1999) x x x \ shc
Beber (1999/2001) x x x \ \ tl
David et al. (1999) x x x \ tl, pre
Hearing (2000) x x \ x L0, she, pre
Bonacci and Maalej (2000) x x \ \ pre, anc
Nguyen et al. (2001) x x x x x fc, c, As, L0
Fanning and Kelly (2001) x x x x L0
Rahimi and Hutchinson (2001) x x x x x x x x x tl, pre, mat
Zarnic and Bosiljkov (2001) x x x x x x mat, bl/tl
Shahawy et al. (2001) x x x \ pre
Spadea et al. (2001) x x x x anc
As, tl, sch, pre, c, sur,
Naaman (2001) x x x x x
anc
Kishi et al. (2001) x x x \ x As, a/D, tl
Shehata et al. (2001) x x \ x Al
Xiong et al. (2001) x x x x As, Al, anc
Continued in next page

22
LITERATURE REVIEW

Continued from previous page


Failure mode3
Laminate Laminate
( \ : not clearly
material1 form2
described)
Reference Variable4
1
S C G A f p & 3 4 5 6 7 8
2
Gao et al. (2001) x x x x tl, adh
Wight et al. (2001) x x x pre, continuous span
Stepanek and Podolka (2001) x x L
Shokrieh and Malevajerdy (2001) x x x pre
Maalej and Bian (2001) x x tl
Zhao et al. (2001) x x x tl, L0
Smith and Teng (2001) x x x x L0, a, anc
Leong and Maalej (2003) x x x beam size
Pornpongsaroj and Pimanmas 03) x x x x x anc
Sawada et al (2003) x x x anc
Nurchi et al.(2003) x x x x x anc
Joh et al. (2003) x x x x mat
Takahashi and Sato (2003) x x x x tl, anc, adh
Valcuende and Benlloch (2003) x x x x Asc, anc
Zhang et al (2003) x x x x D, mat
Niu and Wu (2003) x x tl
Khomwan et al. (2004) x x x fc
Brena and Macri (2004) x x x x tl, bl, Al

Notes: 1) S: steel; C: CFRP; G: GFRP; A: AFRP


2) f: fabric/sheet; p: plate
3) refer to Figure 2.11. In case the model is not clearly described, the closest mode as judged by
the author is listed
4) Al: laminate cross-sectional area; adh: adhesive property; anc: anchorage; bl : laminate width;
mat: laminate material; pre: preloading; pres: prestressing; sch: strengthening scheme; she: shear
reinforcement; sur: surface preparation; tl : laminate thickness.

23
CHAPTER 2

grooves nails
plate

bolt tapered end

angle several curtailment lengths

clamp

I jacket

side plate

U jacket
confining plate

FRP wrap

Figure 2.9 Flexural strengthening using FRP or steel plates

anchorage U anchor
fabric

FRP rod mortar

anchorage wrap

Figure 2.10 Flexural strengthening using FRP fabrics/sheets

24
LITERATURE REVIEW

1. Concrete crushing
3. End shear

5. End interfacial 4. End cover 6. Shear crack 2. FRP rupture


debond peeling debond
8. Flexure-shear 7. Flexural crack
crack debond debond
Figure 2.11 Observed failure modes

It is difficult to determine which debonding mode is more critical. While end


debond is likely in beams with short shear span loading, a laminate curtailment far from
support and a stiff laminate, intermediate span debond is often associated with high shear
span loading, a laminate curtailment near support and a thin laminate. On the one hand, it
could be argued that end debond is more complex and structurally detrimental compared
with intermediate span debond, which is often local. The literature search has also shown
that end debond is the most common debonding mode. On the other hand, close
examination of these experiments reveals that most of them were carried out on small to
medium model beams under four-point loading and boned with relatively thick plates. In
practice, where long shear span loading is frequently the case and the laminates used are
fairly thin, intermediate span debond would be more critical.

Shear strengthening
In practice, flexural strengthening is often combined with shear strengthening by
bonding laminates on both the soffit and the sides of the beam. The laminates on the sides
act like steel stirrups to cross diagonal cracks. They can be narrow strips placed at certain
spacing or wide plates covering the entire shear span. The main direction of the fibres can
be vertical or inclined to intersect shear cracks. The shear laminates can also be anchored
by being wrapped around the beam or by mechanical anchorage (Figure 2.12).

25
CHAPTER 2

FRP bolt

anchorage hole

U-anchor
U-anchor
with FRP
without
rod
FRP rod

lap

Figure 2.12 Shear strengthening using steel plates, FRP plates or FRP sheets

There have been a number of experimental and theoretical studies on RC beams


strengthened in shear, which have been reviewed by Triantafillou and Antonopoulos
(2000). Several shear related failure mechanisms of FRP shear laminates have been
identified including tensile rupture, debonding through the concrete near the bond
interface or localised failures. The ultimate shear strength can be calculated as the sum of
the portions contributed by concrete, steel stirrups and FRP shear laminates. Expressions
used to calculate the portion of the shear force contributed by FRP laminates were mainly
based on the simple truss analogy (Al-Sulaimani et al., 1994; Chaallal et al., 1998a) or
more rigorous compression field theory (Malek et al., 1998; Gendron et al., 1999). To
predict the maximum force that FRP shear laminates can undertake, a reduction in the
FRP capacity due to stress concentration or debonding needs to be considered. Several
simplified empirical formulae for this purpose have been proposed by Al-Sulaimani et al.
(1994), Izumo et al. (1997), Triantafillou (1997) and Triantafillou (1998).

26
LITERATURE REVIEW

2.3.1.2 Parameters influencing beam strength


2.3.1.2.1 Adhesive properties
Several studies have been performed to select suitable types of adhesives for
retrofitting RC beams. Saadatmanesh and Ehsani (1990) tested retrofitted beams using
four different two-component epoxies. They concluded that to secure the success of the
strengthening technique, the epoxies should have sufficient strength to transfer the
bonding shear stress as well as toughness to prevent brittle bond failure as a result of the
cracking of concrete. They recommended the use of rubber toughened epoxies. Further
investigation was carried out by Gao et al. (2001). They varied the rubber content in the
resin from 0 to 20 % and found that the beam capacity and ductility increased slightly.

The effect of different adhesive thicknesses was investigated by Swamy et al.


(1987) and Quantrill et al. (1996). Swamy et al. tested steel plated beams with the plate
thickness varying from 1.5 to 6 mm. Quantrill et al. tested beams bonded with GFRP with
two different adhesive thicknesses: 1 mm and 2 mm. Both research groups found that the
adhesive thickness had only a modest influence on the behaviour of the beams.

2.3.1.2.2 Concrete cover


Little experimental work has focused on the influence of concrete cover depth.
Nguyen et al. (2001) tested four beams 1330 mm long and 120 mm deep. The beams
were retrofitted with CFRP. The beam concrete covers varied from 5 to 25 mm while the
beam total depth stayed unchanged. They found that the beams with the smaller concrete
cover had higher ultimate loads. However this finding might not be valid since the beams
with the smaller cover also had a greater tension reinforcement depth. Naaman et al.
(2001) tested two T-beams retrofitted with CFRP with the concrete covers of 25 mm and
50 mm. They found that the beam with the smaller concrete cover had a similar ultimate
load and larger ultimate deflection.

2.3.1.2.3 Concrete strength


The effect of concrete strength has also been studied by a number of research
groups. Quantrill et al. (1996) tested two series of retrofitted RC beams with concrete
cube strengths of 42 MPa and 70 MPa. They found that the beams with the higher
concrete strength experienced greater percentage increases in the load bearing strength.
From their analytical study, they claimed that higher-strength concrete was able to sustain

27
CHAPTER 2

higher level of both shear and normal stress. A similar observation was also made by Cha
et al. (1999) for strengthened prestressed concrete beams.

2.3.1.2.4 Shear span and laminate bond length


The influence of the ratio of the bending moment to the shear force at the steel
plate end on debond was investigated by Oehlers (1992). He varied the shear span and
plate length in his experiments and found two distinct failure modes: ‘shear peeling’
when the moment to shear ratio was low and ‘flexural peeling’ when the ratio was high.
The first mode of failure was due to peeling of the concrete cover near the plate end and
occurred very rapidly. The second mode of failure was due to propagation of horizontal
peeling cracks induced by flexural cracks, which occurred more gradually. A combination
of these two mechanisms was observed for the specimens with an intermediate value of
the bending moment to shear force ratio.

Instead of varying the shear span and the laminate length, Garden et al. (1997) and
Garden et al. (1998) changed the shear span and the beam depth. In their studies, the
shear span to depth ratio was varied from 3.00 to 7.72. A cantilever set-up was utilised to
obtain high ratios. It was found that as the ratio increased, the failure modes shifted from
end cover peeling to flexure-shear crack debond and then to shear crack debond. There
was inconsistency with the trend and it was claimed to be the result of the small amount
of shear reinforcement. Garden et al. (1998) also showed that the ratio of the shear span
to the beam depth did not significantly affect the ultimate shear capacity but affected the
ultimate moment capacity. Furthermore, they suggested that beyond a shear span to depth
ratio of 5.9, there was no further increase in the ultimate moment capacity.

Several researchers varied the laminate bonded length solely to study its effect on
the behaviour of retrofitted beams. Hearing (2000) fabricated nine rectangular beams
retrofitted with CFRP and tested them in four-point bending with a shear span of 600 and
a total span of 1800 mm. The laminated lengths varied from 600 to 1600 mm. Initial
delamination cracks were introduced near the laminate ends. This method forced all the
beams to fail by end interfacial debond. He reported that the beams with a longer
laminated length had a higher delamination initiation load and a higher ultimate failure
load, and also demonstrated a more brittle failure. Fanning and Kelly (2001) tested six
beams spanning 2800 mm and loaded in four-point bending with a shear span of 1100

28
LITERATURE REVIEW

mm. The CFRP bonded lengths were 715, 638 and 550 mm. All of these beams failed due
to end cover peeling with average ultimate loads of 102, 81 and 72 kN, respectively.
Fanning and Kelly also found that for the beams of shorter bonded length, there was
significant inconsistency in the results. Nguyen et al. (2001) also investigated the effect of
the laminate bond length in their experiments. They noticed that the ultimate loads only
slightly increased with longer bond lengths. The CFRP lengths were 950, 1000 and 1150
mm and the ultimate loads measured were 56, 57 and 59 kN, respectively.

2.3.1.2.5 Beam and laminate aspect ratios


Beam and laminate shapes also prove to be important parameters. Flat flexible
sheets and shallow beams are generally less susceptible to debond since they generally
have greater bond areas. However, studies focusing on these aspect ratios are limited.
Garden et al. (1997) varied the plate aspect ratio and kept the plate area the same in their
experiments. Three plate cross-sections were used: 0.5 x 90 mm, 0.7 x 65 mm and 1.0 x
45 mm. The plates were bonded on RC beams, which were 1000 mm long with a cross
section of 100 x 100 mm. The failure mode observed was end cover peeling. They found
that the beams with the plates of high width-to-thickness ratios had significantly higher
ultimate load capacities. Quantrill et al. (1996) also studied the effect of the plate aspect
ratio in two beams of similar dimensions. However, they found that the two beams had
similar load capacities.

2.3.1.2.6 Steel and FRP reinforcement


Most of the beams strengthened in flexure were reinforced with steel stirrups so
that they had a sufficient shear capacity. However, the shear reinforcement can determine
the spacing and opening of diagonal cracks, which might influence debonding failure.
This aspect was investigated by Oehlers (1992), Tumialan et al. (1999) and Hearing
(2000) for end debond failure mode. All these studies showed that stirrup spacing had
insignificant effect on the end debond failure load.

Laminate stiffness relative to tension reinforcement stiffness is an important


parameter deciding the behaviour of a retrofitted beam. Bonacci and Maalej (2001)
summarized the work done by other researchers and identified the relative stiffness of the
FRP to the tension rebars, AfEf / AsEs, as an important factor determining retrofitted
beams’ behaviour. Figure 2.13 shows a general trend that FRP efficiency improves when

29
CHAPTER 2

the ratio reduces. There is only one experimental study by Kishi et al. (2001), in which
the effect of the relative stiffness was investigated directly. They carried out testing on
RC beams strengthened with AFRP sheets and varied both the tension reinforcement and
AFRP area. They found that the beams with the area ratio Af / As ranging from 1.3 to 1.7
failed by debonding, whereas the beams with this ratio ranging from 0.6 to 0.8 failed in
flexure.

1.0 FRP Rupture


Special Detail/No Rupture
Others
0.8
FRP efficency ratio

0.6

0.4

0.2

0.0
0.00 0.25 0.50 0.75 1.00 1.25
AfEf / AsEs

Figure 2.13 FRP efficiency ratio (max. reported strain / strain at rupture) versus
relative stiffness (Bonacci and Maalej, 2001)

The most common method to investigate the effect of laminate stiffness is to


increase the laminate thickness. Most studies reported an increase in the beam stiffness
and less ductility. However, it was found that shifting of the failure mode occurred very
often and the improvement in ultimate load carrying capacity depended a great deal on
the failure mechanism (Sharif et al., 1994; Hussain et al., 1995; Arduini et al., 1997; Saafi
and Buyle-Bodin, 1997; Ahmed and Van Gemert, 1999; Tumialan et al., 1999; Kishi et
al., 2001; Naaman et al., 2001). The failure mode often changed from flexural to
debonding. When the failure mode was flexural failure or flexural crack debond, the
thicker laminate resulted in the higher ultimate load. On the contrary, when end cover
peeling was the failure mode, the ultimate load varied slightly or even reduced when the
laminate thickness was increased.

30
LITERATURE REVIEW

Some researchers varied tension reinforcement ratios (Saadatmanesh and Ehsani,


1991; Ross et al., 1999; Swamy and Mukhopadhyaya, 1999; Naaman et al., 2001; Xiong
et al., 2001). The general conclusion made was that lightly or moderately reinforced RC
beams had higher strengthening efficiency in terms of the percentage increase in the
ultimate load capacity. Little attention was paid to the change of failure mode in those
studies. The most detailed study on the amount of tension reinforcement was by Ross et
al. (1999), in which the tension reinforcement in their beams was the only variable. They
noticed a shift from delamination to flexure as the reinforcement ratio increased.
However, their beams failed in an uncommon mode: delamination between the FRP
laminate and the adhesive, which could be the result of poor bonding.

2.3.1.2.7 Anchorage
Various anchorage methods have been used (Figures 2.9 and 2.10). Basically,
there are two main anchorage types corresponding to two main debonding failures: end
anchorage and intermediate span anchorage. Some forms of end anchorage are bolts, side
angles, I jackets or U anchors. It has been shown that end anchorage improves the beam
ductility since premature debond is delayed (Hollaway and Leeming, 1999). The beam
strength can also be enhanced by end anchorage (Hussain et al., 1995; Hollaway and
Leeming, 1995). Intermediate span anchorage is usually provided by U-shaped transverse
straps (Swamy and Mukhopadhyaya, 1999). These U-straps have been found to be able to
reduce the opening of diagonal shear cracks, prevent spalling of concrete cover and limit
the propagation of debonding cracks to the portion between the straps. They can therefore
increase deflection and load carrying capacities for beams susceptible to intermediate
span debond.

2.3.2 Existing theoretical studies for flexural failure


Some researchers have shown that the behaviour of a retrofitted section can be
described using the classic beam theory (An et al., 1997; Arduini et al., 1997). This
theory is also known as Bernoulli compatibility truss model or bending theory. It is based
on strain compatibility and equilibrium equations. A detailed derivation of the theory for
a retrofitted section is presented in Appendix A.1.

31
CHAPTER 2

2.3.3 Existing theoretical studies for intermediate span debond


There are three types of intermediate span debond: flexural crack debond, shear
crack debond and flexure-shear crack debond. The mechanisms of these sub-modes are
slightly different. The theoretical models developed were not always clearly specified if
they should be applied to a particular sub-mode. Rather, they were developed to cover all
debonding modes occurring in the intermediate span. It was sometimes referred to as
‘intermediate crack-induced debonding’ or just ‘debonding’. However, for clarity,
theoretical models developed for intermediate span debond are grouped into three
different categories below. It is also worth noting that out of these three modes, flexure-
shear crack debond is the most common in beams with sufficient shear strength and a
normal amount of tension reinforcement.

2.3.3.1 Flexural crack debond


Flexural crack debond occurs when a large flexural crack is present, which is only
likely when the beam is very lightly reinforced or when there is a sudden change in the
beam cross section, such as by notching or corrosion of the tension reinforcement. When
the tensile force in the laminate at the tip of a flexural crack causes a shear stress higher
than the concrete shear strength, the laminate debonds from the concrete. The shear stress
can also lead to rupture of a concrete prism at the tip of the crack. The tensile force then
transfers to the adjacent bonding area and debonding propagates away from the initial
location (Figure 2.14).

Figure 2.14 Flexural crack debond (Sebastian, 2001)

32
LITERATURE REVIEW

The simplest method to design against this failure is to limit the FRP strain in the
vicinity of flexural cracks. Arya and Farmer (2001) suggested two strain limits: 0.008 if
the applied load was uniformly distributed and 0.006 if both high shear forces and
bending moments were present. They also proposed a limit on the steel reinforcement
strain of five times the yield strain. The failure mode referred to was called ‘debonding’.
Similar recommendation was made by Shehata et al. (2001), in which the limiting strain
was 0.005. This value was found from testing of two RC beams. They also proposed a
limit for ‘tearing off the concrete cover’, which will be discussed later in Section 2.3.4.

Some researchers believed that flexural crack debond mechanism is similar to that
in a shear-lap test (Figure 2.15). Consequently, the maximum force Ff in the FRP can be
calculated using the results from a shear-lap test. Maruyama and Ueda (2001) reported
the design recommendation by the Japan Society of Civil Engineers, which states that to
neglect peeling in beams strengthened by FRP sheet, the tensile stress at the location of
the flexural cracking should be limited to
2G f E f
ff ≤ (2.19)
n f t f ,0

where nf is the number of attached FRP layers and tf,0 is the thickness of a layer. The
interfacial fracture energy, Gf, is introduced in this formulation, which is to be obtained
from a double shear-lap test or taken as 0.5 N/mm.

Instead of limiting the FRP stress, Teng et al. (2001) proposed to limit the tensile
force in the FRP. To calculate the limiting force, they used Chen and Teng’s formulation
for a shear-lap test (Equation 2.15). A modification factor, α, was introduced to account
for the differences to the beam bending situation. The limiting force is calculated as
follows

P = αβ f β L f c b f L e (2.20)

By calibrating with several experimental results, α was found to be 0.43.

33
CHAPTER 2

FRP

Ff
Ff
Shear-lap test
Peeling crack

Figure 2.15 Similarity between the shear-lap test and the situation near a flexural
crack

For most RC beams, there exist more than just one wide flexural crack. The
cracks are generally spaced at a certain distance over which the shear stress is transferred.
The crack spacing, therefore, might affect debonding failure. To account for that, analyses
were carried out by Malek et al. (1998) and Niu and Wu (2001b).

Malek et al. derived a formula to calculate the interfacial shear stress distribution
near a flexural crack using a closed-form solution. They indicated that the shear stress
near the tip of a flexural crack should be compared with the interfacial shear strength of
the adhesive.

Niu and Wu analysed the forces acting on an infinitesimal element (Figure 2.16)
and used a bilinear shear bond-slip model to describe the bond between FRP and
concrete. They derived expressions for the shear stress distribution near a flexural crack
or between two flexural cracks. The expressions were verified with a finite element
model and an experiment. A stress-slip model with τmax = 8 MPa, s1 = 0.05 mm and s2 =
0.3 mm was used. Close results were found. Niu and Wu also explained the debonding
mechanism in details (Figure 2.17) for plain concrete and RC beams. They found that for
the latter case, crack spacing played an important role in the stress distribution and
therefore debond propagation.

34
LITERATURE REVIEW

Load

M1 M1 + dM1

N1 RC concrete N1 + dN1

V1 V1 + dV1

τ τ

N2 FRP N2 + dN2
dx

Figure 2.16 Forces acting on an element of a composite beam (Niu and Wu, 2001)

Debonding propagation

Flexural crack

Figure 2.17 Debonding mechanism due to flexural cracks (Niu and Wu, 2001)

Both the expressions by Malek et al. and Niu and Wu are, however, cumbersome.
The proposed formulae were also based on a simplification that all materials had linear
elastic properties.

2.3.3.2 Flexure-shear crack debond


The most common type of intermediate span debond is flexure-shear crack
debond. This mode occurs in the shear span and its failure mechanism is similar to that of
flexural crack debond with an exception that there exists some vertical movement at the
flexure-shear crack tip in addition to the horizontal movement. For most cases, the FRP
laminate is relatively thin, the dowel or bending effect that results from vertical
movement is generally not significant. Therefore, the models developed for flexural crack
debond presented above might be used to predict flexure-shear crack debond failure
loads. There are also some theoretical models developed specifically for debonding
between cracks in the shear span. These models are presented below.

Niedermeier (2000) suggested a failure mechanism, in which the teeth between


adjacent cracks subjected to a differential tensile force (Figure 2.18). Peeling-off at the

35
CHAPTER 2

FRP end and at flexural cracks was treated with the same model. To calculate the
debonding load, two main steps were involved.

In the first step, the most unfavourable spacing of flexural cracks was determined
assuming that the mean bond strengths along the steel-concrete and FRP-concrete
interfaces are constant. The spacing was given as
M cr 1
s rm = 2
z m (τ fm b f + ∑ τ sm d b π )
(2.21a)

The bond strength of steel, τsm, was taken as 1.85 fct. The bond strength of FRP, τfm, was
taken as 0.44 fct. The cracking moment was calculated as
kf ct bh 2
M cr = (2.21b)
6
In above equation, the factor k was taken as 2 to take into account the difference between
flexural and direct tensile strength. The mean lever arm, zm, was calculated using the
following formula:
(hE f A f + d s E s A s )
z m = 0.85 (2.21c)
(E f A f + E s A s )

The second step involved checking the tensile force transfer between two
subsequent cracks. The increase in the FRP tensile stress, ∆σf, should be less than the
maximum allowable increase, max∆σfd, between two cracked sections. The researcher
stated that max∆σfd reduced as σf increased. The envelope for max∆σfd is illustrated in
Figure 2.19. Point A corresponds to the end of FRP, where the maximum tensile stress
was calculated as the following:

c1 E f f c f ct
σ fad ,max = (2.21d)
γc tf

The effective anchorage length was given as


Ef tf
l b ,max = c 2 (2.21e)
f c f ct

If the bond length lb was smaller than lb,max, the maximum tensile stress was calculated as

36
LITERATURE REVIEW

lb ⎛ l ⎞
σ fad = σ fad ,max × ⎜2 − b ⎟ (2.21f)
l b ,max ⎜⎝ l b ,max ⎟

To find point B, where the linear decrease ends, the following equations were provided:
c3E f s
σ fB = − c 4 f c f ct rm (2.21g)
s rm 4t f

1 ⎡⎢ c1 E f f c f ct ⎤
2

max ∆σ = B
fd + (σ f ) − σ f ⎥
B 2 B
(2.21h)
γc ⎢ tf ⎥
⎣ ⎦
Point C is defined by the stress limit reaching the tensile strength of FRP.

The disadvantage of this method is its complexity, which makes it difficult to


apply as a practical engineering model. The constants, c1, c2, c3 and c4, also need to be
calibrated directly with shear-lap tests.

L0

lb srm srm srm srm srm

Ff,1 Ff,i Ff,i+1

Ff,1

Ff,i+1 - Ff,i
Ff envelope

Figure 2.18 Niedermeier’s methodology.

37
CHAPTER 2

Figure 2.19 Maximum possible increase in tensile stress between two subsequent
cracks (fib Bulletin 14, 2001).

A simple design method was proposed by Matthys (2000). The shear stress at the
FRP-concrete surface was limited to the bond strength of concrete, fcv, which was
calculated as:
f cv = 1.8f ct (2.22a)
Matthys adopted a number of simplifications and proposed the following formulae to
calculate the final design shear force:
⎛ A E ⎞
Vd = f cv 0.95db f ⎜⎜1 + s s ⎟⎟ if tensile steel does not yield (2.22b)
⎝ Af Ef ⎠
Vd = f cv 0.95db f if tensile steel yields (2.22c)

2.3.3.3 Shear crack debond


Shear crack debond is not a common failure mode. It is caused by a combination
of both horizontal and vertical displacements (Figure 2.20). Shear crack debond is only
likely in beams where there exists a shear crack with its tip opening wide enough. For
most cases, the presence of stirrups and tension reinforcement prevents this from
happening.

38
LITERATURE REVIEW

steel

concrete

FRP

Figure 2.20 Displacements near the tip of a shear crack

There are only a few studies that identified and analysed this mechanism. The first
attempt was by Triantafillou and Plevris (1992). For simplification, bending action was
ignored. The ultimate shear force was believed to be proportional to the total shear
stiffness of the steel reinforcement and the FRP, as follows
⎛v⎞
V ~ ⎜ ⎟ (G s A s + G f A f ) (2.23)
⎝ w ⎠ cr
where v and w are the vertical and horizontal movements at a crack tip. The researchers
indicated that peeling occurred when the ratio of vertical to horizontal displacement
reached a critical value. However, no detailed analysis and verification were reported.

Neubauer and Rostasy (1999) presented a detailed analysis of these shear crack
mouth opening displacements using a truss model with shear crack friction (Figure 2.21).
The mouth displacements were given as
w = 2s cr (ε x − ε cw cos 2 θ) (2.24a)

v = 2s cr tan β cr (ε x − ε cw cos 2 θ) (2.24b)

where scr is the crack spacing, βcr is the angle of inclined shear cracks and θ is the angle
of the compression strut inclination. εx, εz and εcw are the strains in the longitudinal,
perpendicular and compression strut directions given by
ηB − 1 ⎛M τb v d ⎞
εx = ⎜⎜ + ⎟⎟ (2.24c)
2η B E f A f ⎝ Z l 2 tan θ ⎠
τ tan θ
εz = (2.24d)
E s µ sw

39
CHAPTER 2

τ(1 + tan 2 θ)
ε cw =− (2.24e)
E c tan θ
V
where τ = and ηB is the degree of strengthening in bending. However, many
bvd

approximations were needed and a clear design implication for shear crack debond was
not available.

Figure 2.21 Differential crack mouth opening displacements causing peeling


(Neubauer and Rostasy, 1999)

The vertical and horizontal movement of the concrete tip cause peeling of the FRP
from the concrete surface possibly in a similar manner as in the peel tests described by
Karbhari and Engineer (1996) and Karbhari et al. (1997) and illustrated in Figure 2.22. In
the first paper, different specimens with a number of peeling angles, α, were tested to
determine the interfacial fracture energy for mode I (tension) and mode II (shear).
Karbhari and Engineer found that as the angle decreased, GfI increased and GfII decreased.
The total interfacial fracture energy for the bonding between CFRP and concrete was
found to be around 0.56 N/mm. In the second paper, the specimens were exposed to
various conditions (ambient temperature, immersion in fresh water, immersion in sea
water, low temperature and freeze-thaw cycles) to study the environmental effects on the
bond. The authors also listed four generic mechanisms of adhesion with the emphasis on
the dominance of mechanical interlocking.

40
LITERATURE REVIEW

δy
FRP
α δx

roller

concrete

Figure 2.22 A peel test (Karbhari et al., 1997)

To design for shear crack debond, a simple design method was described in
Blaschko (1997), in which the acting shear force was limited to the modified concrete
shear capacity without shear reinforcement. The proposed formulae are
VR = k × b v d (1.2 + 40ρ l ) τ Rk (2.25a)

Ef
ρ l = (A s + A f ) /( b v d ) (2.25b)
Es
The factor k was to take size effect into account and given as k = 1.6 − d ≥ 1 , where d is
in meters. The shear strength, τRk, was found by calibrating with a number of test results
and given by:
τ Rk = 0.18(f c )1 / 3 (2.25c)

Another design model was developed by Mohamed Ali (2000). This study was
based on Zhang’s model (Zhang, 1997), which used the theory of plasticity to calculate
shear strength of reinforced concrete beams. For plain concrete beams, Zhang’s model
postulates that shear cracks generally extend from the load point at the top of the beam to
the position near the support. The closer the crack to the support, the larger the cracking
load, Vcr, (due to a larger crack area) and the smaller the shear strength, Vu, are (due to a
larger sliding angle, α). The critical diagonal crack is at a position where Vcr equals Vu.
For beams bonded with laminates, Mohamed Ali made modifications for Vcr (to account
for the additional resistance due to the tensile force in the laminates) and for Vu (by

41
CHAPTER 2

adding an additional term for the longitudinal reinforcement) and proposed the following
formulae:
Lb

Pb

V Vu

Vuc Vcr

(Ld)crit Ld
(a) Unplated (b) Plated

Figure 2.23 Crack sliding model (Mohamed Ali, 2000)

For beams with laminate bonded on the tension face, the cracking load was given as
⎛ x2 + h2 ⎞⎛ f tef b c mf ct b f t f (h + 0.5t f ) ⎞
Vcr = ⎜⎜ ⎟⎜
⎟ 2 + ⎟ (2.26a)
⎝ a ⎠⎝ h2 ⎠

where x is the distance from the crack tip to the load point, h is the total depth of the
beam, fct is the direct tensile strength of concrete and m is the ratio of FRP to concrete
stiffness Ef / Ec. The tensile strength, ftef, was calculated as the following:
−0.3
⎛ h ⎞
f tef = 0.156f c
2/3
⎜ ⎟ (2.26b)
⎝ 100 ⎠
The shear strength was given as

1 ⎛ ⎛x⎞
2
x⎞
Vu = γ s γ 0, tfp f c 1 + ⎜ ⎟ − ⎟b c h
⎜ (2.26c)
2 ⎜ ⎝h⎠ h⎟
⎝ ⎠
where
γs = 0.5; γ0,tfp = λ.f1(fc).f2(h).f3(ρf); λ = 1.6 (2.26d)
3.5
f 1 (f c ) = 5 MPa < fc < 60 MPa (2.26e)
fc

⎛ 1 ⎞
f 2 (h ) = 0.27⎜⎜1 + ⎟⎟ 0.08 m < h < 0.7 m (2.26f)
⎝ h⎠

42
LITERATURE REVIEW

f 3 (ρ f ) = 0.15(ρ s + ρ f ) + 0.58 (2.26g)

As A tfp ,eff
ρs = 100 ; ρ f = 100 (2.26h)
bch bch
The effective area, Atfp,eff, was to account to the fact that the FRP might delaminate from
the concrete. Mohamed Ali suggested the following equations, which were developed for
a shear test between a steel plate and concrete:
Atfp, eff = Pb / ffu (2.26i)
Pb = 1.09 Lb fct bf ≤ 71ft tf bf (2.26j)
where Lb = (a – x – L0).

2.3.4 Existing theoretical studies for end debond


End debond comprises end interfacial debond and end cover peeling. With proper
bonding, the crack tends to propagate along the weakest layer, which is at the tension
reinforcement level. Consequently, end cover peeling is by far the most common failure
mode, while end interfacial debond is rarely observed. There are numerous theoretical
models developed for end cover peeling. However, since both end cover peeling and end
interfacial debond result from the high interfacial shear stress near FRP ends, there are a
number of models proposed to address both types of failure.

In this review, prediction models for end debond are classified into four main
approaches. They are interfacial localised failure based models, shear capacity based
models, concrete tooth scenario based models and shear-lap scenario based models. Some
models are the combination of several approaches. It is important to emphasize that
although some models were developed for steel plated beams, they are expected to be
applicable for FRP plated beams with or without modification. In the formulae presented
here, the units used are Newton (N) and millimetre (mm), unless noted otherwise. For
comparison, some equations are put in a slightly different form from the originals. Since a
number of models are dependent on the studies on the interfacial stress concentration near
the FRP end, a summary of those studies is presented first in the next section.

2.3.4.1 Interfacial stresses near FRP ends


In various attempts to explain end debond phenomenon, many researchers
concluded that high local interfacial shear and normal stress concentrations exist at the

43
CHAPTER 2

end of the laminate. Several approximate closed-form solutions were derived to predict
the peak stresses near the laminate ends.

Shear

Stress

Normal

Distance from FRP end

Figure 2.24 Predicted shear and normal stress distributions near the plate end
(Roberts, 1989)

The earliest attempt to derive a simple approximate formulation was by Roberts


(1989). Roberts’ prediction of shear and normal stress distributions are shown in Figure
2.24 with the maximum shear and normal stresses at the end of FRP, denoted as τ0 and
σ0, respectively, given by the following formulae:
τ 0 = C R1 V0 (2.27a)

σ0 = C R 2 τ0 (2.27b)
where
⎡ ⎛ K ⎞
1/ 2
M0 ⎤ bf t f
C R1 = ⎢1 + ⎜⎜ s
⎟⎟ ⎥ (d f − d n ) (2.27c)
⎢⎣ ⎝ f f t f
E b ⎠ V0 ⎥ I tr ,f b a

1/ 4
⎛ Kn ⎞
CR2 = t f ⎜⎜ ⎟⎟ (2.27d)
⎝ 4E f I f ⎠
Ga ba Ea ba
Ks = Kn = (2.27e)
ta ta
M0 and V0 are the bending moment and shear force at the FRP end, respectively. By
comparison with more rigorous solutions and test data, Roberts recommended that M0
should be replaced by M*, which is the bending moment evaluated at a distance (h + tf) / 2
from the end of the laminate.

44
LITERATURE REVIEW

Several modifications of Roberts’ solution have also been proposed. Ziraba et al.
(1994) calibrated Roberts’ equations with their finite element analyses. They proposed
the following formulae to calculate the maximum shear and normal stresses:
5/ 4
⎛C V ⎞
τ 0 = 35f ct ⎜⎜ R1 0 ⎟
⎟ (2.28a)
⎝ fc ⎠
σ 0 = 1.1C R 2 τ0 (2.28b)
El-Mihilmy and Tedesco (2001) carried out a calibration study with nine experimental
programs and proposed additional empirical coefficients for Roberts’ equations. Their
formulae to calculate the peak shear and normal stress near the cut-off point are:
Ea tf
τ 0 = τ + c1 σx (2.29a)
Ef
1/ 4
⎛ 3E t ⎞
σ 0 = ⎜⎜ a f ⎟⎟ τ0 (2.29b)
⎝ Ef ta ⎠
where τ and σx are the elastic shear and longitudinal laminate stress, respectively. They
are calculated as the following:
V0 t f y f
τ= (2.30a)
I tr .f

M0 yf
σx = (2.30b)
I tr ,f

Using a linear regression, the researchers found that the constant c1 could be taken to be
0.28.

Independently, Malek et al. (1998) derived equations to predict the stress


distributions near the FRP end for a beam under an applied bending moment M(x)
expressed by
M (x) = a1x2 + a2x + a3 (2.31a)
They developed a methodology based on the linear elastic behaviour of the materials and
the compatibility of deformation (Figure 2.25) and presented a closed-form solution. The
maximum shear and normal stresses were found to be:
τ0 = t f (b3 A + b 2 ) (2.31b)

K n ⎛ Vf V0* + β*M 0 ⎞ qE f I f

σ 0 = *3 ⎜ − ⎟⎟ + (2.31c)
2β ⎝ E f I f E c Ic ⎠ bf E c Ic

45
CHAPTER 2

where
Ga
A= (2.31d)
t a t f Ef

y f a 1E f
b1 = (2.31e)
I tr ,c E c

yf Ef
b2 = (2a 1 L 0 + a 2 ) (2.31f)
I tr ,c E c

⎡ yf t t ⎤
b3 = Ef ⎢ (a 1 L20 + a 2 L 0 + a 3 ) + 2b1 a f ⎥ (2.31g)
⎣ I tr , cE c Ga ⎦

V0* = V0 − 0.5hb f t f (b 3 A + b 2 ) (2.31h)

Vf = −0.5b f t f2 (b 3 A + b 2 ) (2.31i)
1/ 4
⎛K b ⎞
β = ⎜⎜ n f
*
⎟⎟ (2.31j)
⎝ 4E f I f ⎠
q is the external distributed load applied on the beam, yf is the distance from the neutral
axis to the centroid of FRP plate, Itr,c is the moment of inertia of the transformed section
and Kn = Ea / ta.
fn

σ FRP σ + dσ/dx dx

dx

Figure 2.25 Stresses acting on a FRP laminate element near the cut-off point (Malek
et al, 1998)

In a recent study by Smith and Teng (2001), a more rigorous analysis was carried
out to derive the formulae for shear and normal stresses. The solution was later verified
with a linear elastic finite element model using a very dense mesh. However, their
formulation is rather complicated, which makes it difficult to be used in design.

Experimental information on these peak stresses is very limited. The only


verification was done by Roberts (1989). He compared the bond stress distributions along

46
LITERATURE REVIEW

the retrofitted steel plates in the beams tested by Jones et al. (1988) with his theory. No
verification of the end peak stresses has been made for beams bonded with FRP
laminates. It might be due to the difficulty in measuring the stress state close to the FRP
ends since the peak stress region there is much less clear for these beams (Aprile et al.,
2001; Rahimi and Hutchinson, 2001)

2.3.4.2 Models based on interfacial localised failure


In the interfacial localised failure based models, the stress state near the bonding
surface is considered. There are two main approaches. In the first, the interfacial stresses
along the bond length in the shear span are considered as the determining factor. The
studies using this approach are summarised below.

Jones et al. (1988) used the interfacial shear capacity as the failure criteria. Based
on their experimental work on steel plated beams, the researchers stated that anchorage
failure occurred when the peak shear stress exceeded 2 times the concrete tensile
strength. The peak shear stress was estimated to be about twice the elastic shear stress.
The elastic shear stress was given by
V0 A f ( E f / E c ) y f
τ= (2.32)
I tr ,c b f

This is a relatively rough estimation based on very limited experimental data.

Sharif et al. (1994), Ziraba et al. (1994) and Chaallal et al. (1998) suggested that
the interfacial shear stress in the adhesive should be limited to avoid debonding failure.
The peak shear stress was calculated using Roberts’ analytical solution. Sharif et al.
(1994) checked for end debond by comparing the interfacial shear stress with the shear
capacity of the adhesive. They recommended a shear strength of 3.5 MPa. Ziraba et al.
(1994) and Chaallal et al. (1998) differentiated between end interfacial debond and end
cover peeling. The latter mode was checked using another approach and will be described
later. The former mode was checked using Mohr-Coulomb criterion for the stress state in
the adhesive. Failure was deemed to occur when τ0 + σ0tanφ exceeded the allowable shear
cohesion, c, for the epoxy. φ is the internal friction angle. Ziraba et al. (1994) calculated
the peak shear stress at the laminate end using their modified expressions of Roberts’
equations. The cohesion, c, and internal friction angle, φ, were found to be 5.36 MPa and

47
CHAPTER 2

28 degrees, respectively. Chaallal et al. (1998) calculated the shear stress using Roberts’
solution without modification. The cohesion and internal friction angle recommended
were 5.4 MPa and 33 degrees, respectively.

Shehata et al. (2001) suggested that the average interfacial shear stress along the
shear span should be limited to
τlim = 0.3fct (2.33a)
to avoid end debond. The corresponding limit on the plate stress was found using an
equilibrium condition for the plate and given as
σlim tf = τlim a (2.33b)
This recommendation was based on the observation of tearing-off of the CFRP strips used
for shear reinforcement in their experiments. No verification for end debond of
composites bonded on a beam soffit was carried out.

Another form of the first approach is based on laminate strain distributions.


Nguyen et al. (2001) identified three distinct zones in the strain distributions (Figure 2.26)
and believed that failure occurred when the strain at the transition point reached a limiting
value. The location of this transition point was given by
d f 4.61
l dev = c + + (2.34a)
2 λ
1 GaGc
λ2 = (2.34b)
Ef tf Gcta + Gac

where Gc is the concrete cracked shear modulus taken as 0.4Ec / 2(1 + νc). They
postulated that the limiting strain depended mainly on the interfacial shear strength,
which could be found from a shear test. In their tests, the limiting strain varied from
0.0017 to 0.0024.

48
LITERATURE REVIEW

P/2

ldev
lb
CFRP strain
(1) (2) (3)

transition point

Figure 2.26 Definition of bond development length for retrofitted beams by Nguyen
et al. (2001)

In another study, Fanning and Kelly (2001) suggested that the limiting strain
gradient controlled the ultimate loads. This was based on the results of six specimens
strengthened with CFRP composites, in which the CFRP strain gradient varying from 4.2
to 5.5 (strain/mm × 10-6). They reported that the strain dropped from the maximum value
under the load point to zero at the plate end.

In the second approach, the stress state at the cut-off point is used to check for end
debond failure. End cover peeling is deemed to happen when the maximum principle
stress in the concrete just above the cut-off point exceeds the concrete strength. The
concrete at this location undergoes a biaxial stress state with three stress components
present: interfacial shear stress, interfacial normal stress and flexural longitudinal stress
(Figure 2.27). There are differences among the researchers in the calculation of these
stresses and the selection of a concrete strength model.

Figure 2.27 Stresses acting at the cut-off point of FRP (Saadatmanesh and Malek,
1998)

49
CHAPTER 2

Saadatmanesh and Malek (1998) proposed that the interfacial shear and normal
stresses were calculated using their formulation presented in Section 2.3.4.1. To find the
flexural longitudinal stress, the researchers suggested that the applied bending moment at
the plate curtailment location must be increased to an amount given by the moment Mm,
as follows
M m = L 0 t f b f y c (b 3 A + b 2 ) (2.35)

where y c is the distance between the centroidal axis and the bottom of concrete beam.
The strength model for concrete under bi-axial stresses presented by Kupfer and Grestle
(1973) was used. According to this model, the strength of concrete was approximated by:
σ2 σ
= 1 + 0 .8 1 biaxial compression-tension (2.36a)
f ct fc

σ 2 = f ct = 0.295(f c ) 2 / 3 biaxial tension-tension (2.36b)

where σ1 and σ2 are the principle stresses.

Tumialan et al. (1999) developed a similar model to predict end debond failure
loads. However, the interfacial shear and normal stress were calculated using Roberts’
formulae. The longitudinal stress was also determined from a bending analysis but the
additional moment was not included. End cover peeling was deemed to occur when the
maximum principle stress reached the concrete tensile strength, which was recommended
as

f ct = 0.689 f c (2.37)

El-Mihilmy and Tedesco (2001) used their formulae (Section 2.3.4.1) to find the
interfacial shear and normal stress. To find the longitudinal stress, they proposed to
replace the bending moment M0 at the plate curtailment with M* given by
⎛ L ⎞ L
M * = ⎜1.35 − 12.5 0 ⎟M 0 ; 0 ≤ 0.1 (2.38)
⎝ L ⎠ L
where L is the total span length. They also suggested that the longitudinal stress may be
ignored since it is quite small near the support. The strength model used for concrete
under bi-axial state was based on Tasuji and Slate’s criterion (Tasuji and Slate, 1978).

50
LITERATURE REVIEW

This criterion stated that the failure occurred if the principle tensile stress was greater
than the tensile strength, ftu, given by:
⎛ σ ⎞
f tu = f ct ⎜⎜1 + 2 ⎟⎟ biaxial compression-tension (2.39a)
⎝ fc ⎠

f tu = 0.59 f c biaxial tension-tension (2.39b)

Different values for the moment of inertia, Itr, were used in the above three
models. Saadatmanesh and Malek, and Tumialan et al. used the moment of inertia of a
uncracked section; whereas El-Mihilmy and Tedesco used the moment of inertia of a
cracked section.

2.3.4.3 Models based on shear capacity


In the shear capacity based models, end debond is related to the beam shear
capacity. It is reasonable to relate to the shear strength since end debond is the result of
the shear stress near the bond surface, which depends on the shear force applied.

Oehlers (1992) separated two cases of peeling in beams bonded with steel plates.
When the moment to shear ratio near the laminate end, M0/V0, is high, flexural peeling
occurs. When this ratio is low, shear peeling occurs. For the first case, peeling was
assumed to occur when the sum of the normal stresses due to the beam curvature and due
to the axial strain near the plate end exceeded the concrete tensile strength. The
corresponding moment at which flexural peeling occurs, Mdb,f, was then calculated and
calibrated with the test results reported in Oehlers and Moran (1990). For the second case,
Oehlers observed that the amount of stirrups did not influence the shear peeling strength
and proposed a limitation on the shear capacity of the concrete RC beam without stirrups,
Vbd,s. For a general case, peeling was deemed to occur when the sum of V0/Vbd,s and
moment M0/Mdb,f at the cut-off point was greater than 1.17. The proposed formulae are:
M0 V
+ 0 ≤ 1.17 (2.40a)
M db ,f Vbd ,s
1/ 3
⎛A f ⎞
Vbd ,s = β1β 2 β 3 b v d 0 ⎜⎜ st c ⎟
⎟ (2.40b)
⎝ bvd0 ⎠
E c I tr ,c f ct
M db ,f = (2.40c)
0.474E f t f

51
CHAPTER 2

Smith and Teng (2003) modified these equations slightly based on their tests.
Their formulae are:
M0 V
0.4 + 0 ≤ 1.0 if Vdb,end ≥ 0.6Vdb,s (2.41a)
M db ,f Vbd ,s

M 0 ≤ M db ,f if Vdb,end < 0.6Vdb,s (2.41b)

In their paper, the flexural debonding moment was taken as


E c I tr ,c f ct
M db ,f = (2.41c)
0.901E f t f

Jansze (1997) used the ‘fictitious shear span’ concept and modified the model
described in Kim and White (1991) to compute the beam shear resistance for RC beams
bonded with steel plate. In the study by Kim and White for RC beams, the shear stress
distributions in a cracked beam and the reduction of internal moment arm length due to
arch action were considered to compute the shear force Vcr,s required for the initiation of
an inclined-shear crack. Another expression for the shear force Vcr,f at which a crack was
initiated for crack was derived based on cracking moment of the section. Where these two
curves intersected was the location of the critical inclined shear crack. Noticing that for a
retrofitted beam, the critical location was at the end of the steel plate, Jansze, by analogy,
derived an expression for the fictitious shear span, aL, and calculated the shear force
capacity, VRd, as the following:
VRd = τ Rd b c d s (2.42a)

ds ⎛ ⎞
τ Rd = 0.183 3 ⎜1 + 200 ⎟3 100ρ s f c (2.42b)
aL ⎜ d s ⎟⎠

aL = 4
(1 − ρs )2

dL3 < a (2.42c)


ρs

a > L + ds (2.42d)
Jansze stated that this design method may be used for both plate-end shear and end cover
peeling.

Ahmed and Van Gemert (1999) modified the study by Jansze (1997) to account
for the difference between the shear stress generated due to shear force when replacing a

52
LITERATURE REVIEW

steel plate with an equivalent FRP laminate. The expressions to calculate the maximum
shear capacity, VRd, was reported as
VRd = τ FES b c d s (2.43a)

⎛ S S ⎞ ⎛ τ − 4.121 ⎞
τ FES = τ PES + τ PES b c d s ⎜⎜ s − f ⎟⎟ + 6188.5⎜⎜ ⎟⎟ (2.43b)
⎝ Is bf If ba ⎠ ⎝ b d
c s ⎠

⎛ 17.1366ρ s d s ⎞ A sv f sy
τ = ⎜ 0.15776 f c' + ⎟ + 0.9 (2.43c)
⎝ a ⎠ sb c

ds ⎛ 200 ⎞3
τ PES = 0.183 3 ⎜1 + ⎟ 100ρ s f cm (2.43d)
aL ⎜ d ⎟
⎝ ⎠

(1 − ρ s ) 2
aL = 4 dL3 (2.43e)
ρs

Ss and Is are the first and second moment of area of the cracked section transformed to
concrete with an equivalent steel plate with the thickness of tf ×(ffu /fsy). It is not clear
how the reported formulae were derived.

Ziraba et al. (1994) checked for end cover peeling based on the modified shear
capacity Vuc + kVus where Vuc and Vus are the shear contribution of plain concrete and
stirrups, respectively, and k is the empirical modification coefficient used to account for
the stress concentration near the cut-off point. From a regression analysis of a set of
experimental data containing several plated beams, they obtained an expression for k, as
follows
6
k = 2.4e −0.08×10 C R 1C R 2
(2.44a)
where CR1 and CR2 are given in Equations 2.27a and 2.27b, respectively. The shear
capacities, Vuc and Vus, were given in Ziraba et al.’s paper as
1
Vuc = (f c + 100ρ s )b c d s (2.44b)
6
A sv f sy ,f d s
Vus = (2.44c)
s

2.3.4.4 Models based on concrete tooth scenario


These models make use of the concept of a concrete tooth between two adjacent
cracks behaving like a cantilever under the action of the horizontal interfacial shear stress
(Figure 2.28). Debond is deemed to occur when the tensile stress at the root of a tooth

53
CHAPTER 2

exceeds the concrete tensile strength. To calculate the crack width and the interfacial
shear stress, different methods were used among the researchers.
Tension bar

lmin

Figure 2.28 Teeth scenario (Zhang and Raoof, 1995)

A model based on concrete tooth scenario was developed by M. Raoof and his
associates (Zhang and Raoof, 1995; Raoof and Zhang, 1997; Raoof and Hassanen, 2000).
They assumed that the crack patterns for plated and unplated beams were similar. The
crack spacing was calculated from the classical theory of cracking by Watstein and
Parsons (1943) as the following:
A e f ct
l min = (2.45a)
u s ∑ O bar + u f b f

l max = 2l min (2.45b)

where Ae is the area of concrete in tension, ∑O bar is the total perimeter of the tension

reinforcing bars, us and uf are the steel-to-concrete and FRP-to-concrete average bond
strengths, respectively. Raoof and Hassanen (2000) used the following formulae to
calculate these bond strengths:
u s = 0.28 f cu (2.45c)

u f = 0.8 MPa (2.45d)


where fcu is the concrete cube strength. The mean shear stress, τ, acting on the bottom of
the teeth was balanced by the axial stress in the laminate over an effective length, Lp. By
calibrating with a number of tests, the following expressions were recommended:
L p = l min (11.6 − 0.17l pmin ) if l min ≤ 56.5 mm (2.45e)

L p = 2l min if l min > 56.5 mm (2.45f)

With a known Lp, the limit axial stress σ in the plate was calculated using the equilibrium
condition for the plate (σ = τLp / tf). The lower bound of FRP tensile stress was found to
be

54
LITERATURE REVIEW

A e f ct2 L p bc
σ f (min) = (2.45g)
6ct f (u s ∑ O bar + u f bf ) bf

The upper bound was


σ f (max) = 2σ f (min) (2.45h)

To find the tensile strength of concrete, the researchers used the following formula:
f ct = 0.36 f cu (2.45i)

Another concrete tooth scenario based model was proposed by Chaallal et al.
(1998). The crack spacing was simply taken as the stirrups spacing. Only the last tooth
near the FRP end was considered and the shear stress acting on the tooth was calculated
using Robert’s expressions. The ultimate shear peeling stress was calculated as
f ct s b c
τ peeling
ult = (2.46a)
6c b f
The researchers recommended that the tensile strength of concrete was taken as
f ct = 0.53 f c (2.46b)

2.3.4.5 Models based on shear-lap scenario


In this method, end peeling is checked for using the maximum tensile force in the
laminate. This is done based on the studies on shear-lap tests (Figure 2.29). This model
was explained in Neubauer and Rostasy (1997) and Arya and Farmer (2001). A truss
model of the beam was suggested with the tension stringer comprised of the internal and
external reinforcement. The anchorage strength was given by Neubauer and Rostasy’s
formulation for shear-lap tests (Equation 2.8a). The shear-lap scenario based model is not
consistent with the general observation that the crack starts from the laminate end and
extends to the tension rebar. Moreover, the bond strength depends on the distance from
the FRP end to the outermost crack, lb, of which the location is not always obvious. Arya
and Farmer (2001) suggested an additional condition: the longitudinal shear stress
between the FRP and the substrate should be smaller than 0.8 MPa. The stress was
calculated using a simple linear elastic theory without considering the stress concentration
(Equation 2.32).

55
CHAPTER 2

FRP

lb

Figure 2.29 Shear-lap scenario for end debond

2.3.5 Design guidelines


Recently, a number of design guidelines have become available. Different
guidelines adopt different approaches to check for debonding. Three guidelines are
summarised below.

2.3.5.1 fib Bulletin 14: Externally bonded FRP reinforcement for RC structures
This report has the most complete treatment for debonding or loss of composite
action. The failure modes are divided into: ‘peeling-off caused at shear cracks’, ‘peeling-
off at the end anchorage and at flexural cracks’, ‘end shear failure’ and ‘peeling-off
caused by the unevenness of the concrete surface’. For the first mode, the design method
recommended is by Blaschko (1997). To check for the second modes, three approaches
can be used. Approach 1 is based on Ayer and Farmer (2001) and Neubauer and Rostasy
(1999) models. Approach 2 is based on Neidermeier (2000). Approach 3 is based on
Matthys (2000). For end shear failure, the model developed by Jansze (1997) is
recommended.

The beam theory is also recommended to analyse retrofitted sections. It is


suggested that for the case of steel yielding followed by FRP fracture, the concrete
rectangular stress block is modified using two factors β and γ given by

⎧ε c (425 − 70833ε c ) for ε c ≤ 0.002


⎪ 1
β=⎨ (2.47a)
1− for 0.002 < ε c ≤ 0.003
⎪⎩ 1765ε c

56
LITERATURE REVIEW

⎧ 4 − 500ε c
⎪⎪ 6 − 1000ε for ε c ≤ 0.002
γ=⎨ c
(2.47b)
1000ε c (3000ε c − 4) + 2
⎪ for 0.002 < ε c ≤ 0.003
⎪⎩ 1000ε c (3000ε c − 2)

2.3.5.2 Reports by ACI Committee 440


Two recent publications by ACI Committee 440 are ‘Guidelines for the selection,
design and installation of fibre reinforced polymer (FRP) systems for externally
strengthening of concrete structures’ published in 1999 and ‘Guide for the design and
construction of externally bonded FRP systems for strengthening concrete structures’
published in 2001. To check for debonding, the first guideline recommends the method
by Blaschko (1997) or Malek et al. (1998). In the second publication, two failure modes
are distinguished as ‘cover delamination’ and ‘FRP debonding’. It is recommended that in
order to prevent debonding of the FRP laminate, the strain in the FRP should be limited to
a maximum strain, which can be found by multiplying a factor κm to the FRP ultimate
strain. The factor κm is given as
⎧ 1 ⎛ n f E f t f ,0 ⎞
⎪ ⎜⎜1 − ⎟ ≤ 0.9 n f E f t f ,0 ≤ 180000
⎪ 60ε fu ⎝ 360000 ⎟⎠
κm = ⎨ (2.48)
⎪ 1 ⎛⎜ 90000 ⎞⎟ ≤ 0.9 n f E f t f ,0 > 180000
⎪ 60ε fu ⎜⎝ n f E f t f , 0 ⎟⎠

where nf is the number of plies of FRP and tf,0 is the nominal thickness of one ply of FRP.
This recommendation is based on a general recognised trend and on the experience of
engineers practicing the design of bonded FRP system. The second report also specifies
detailing requirements to prevent cover delamination. Two requirements for simply
supported beams are:
1) The FRP anchorage length shall pass the cracking point a distance d.
2) The shear force at the FRP end shall be less than 2/3 of the concrete shear
strength.
The beam theory is also recommended to analyse a retrofitted composite section. In
addition, the document recommends the use of an additional strength reduction factor of
0.85 to be applied to the flexural contribution of the FRP reinforcement.

57
CHAPTER 2

2.3.5.3 ISIS Canada: Strengthening reinforced concrete structures with externally-


bonded fibre reinforced polymers – Design Manual No. 4
The manual provides limited guidance to design against debonding. It states that
the debonding ‘can be avoided by using sufficient anchorages or by allowing sufficient
bonding length to develop the full capacity of the FRP reinforcement’. The manual
recommends that the development length of the FRP can be calculated as
lfrpd = kd Ef tf (2.49a)
where kd is a factor relating the development length to the stiffness of the external FRP
reinforcement and given as
b f ε fu
kd = (2.49b)
bc k fc

k is the proportionality factor relating the mean bond strength to the concrete strength. k
depends on the type of FRP reinforcement, the modulus of rupture of the concrete and the
type of application. k can be taken to be 0.184 as found from the study by Bizindavyi and
Neale (1999).

2.4 Finite element modelling


Finite element modelling (FEM) has become an indispensable tool used in
structural engineering. In the past, the majority of analyses were restricted or carried out
only in the linear elastic range. Concrete, however, tends to behave in a non-linear
manner primarily due to cracking, non-linear material behaviour and time dependent
effects such as creep, shrinkage, temperature and loading history. Therefore, linear
elastic analysis is simply inadequate in describing the complete behaviour of concrete and
non-linear analysis is required. For the last few decades, there has been much
development in the nonlinear finite element analysis (NLFEA) of reinforced concrete.
This method has proved to be a powerful tool to study the behaviour of concrete
structures.

In a nonlinear finite element analysis, there is a certain level of approximation,


which leads to a certain margin of error in the model. The error can be inherent in the
method itself as for any numerical simulation. It can also come from approximation in the
mathematical models to represent concrete, steel, CFRP and bonding between them.

58
LITERATURE REVIEW

Hence, it is usually necessary to have some test results against which the output from the
finite element analysis may be compared.

2.4.1 Review of concrete properties and modelling methods


2.4.1.1 Concrete mechanical properties
On the meso-level, normal strength hardened concrete consists of a complex
arrangement of inclusions in a more or less homogeneous matrix. The inclusions can be
pores and course aggregates. The homogeneous matrix consists of smaller aggregates
embedded in a hardened cement paste. Concrete mechanical properties, therefore, can be
influenced by a large number of parameters such as the ratio between the matrix and
inclusion stiffness, the matrix composition, the type of the aggregates, the texture and
roughness of aggregates, the maximum size of the aggregates as well as their size
distribution (RILEM Technical Committee 90-FMA, 1989). In a very early age of a
concrete structure, i.e. during hardening of the concrete, microcracks may appear due to
several sources, such as voids caused by shrinkage or thermal movements and incomplete
compaction. When a load is applied, additional microcracks form due to the incompatible
deformation of the aggregate and cement matrix. The region around the tip of the cracks
is subjected to a high stress concentration and thus a high strain energy concentration.
Crack propagation occurs when this energy exceeds the energy capacity of the material.
This helps to release any excessive strain energy due to external loading. At or just before
the peak load, the existing microcracks start to grow and/or new microcracks originate.
With increasing deformation, the microcracks coalesce to form a single macrocrack
(Kotsovos and Pavlovic,1995).

On the macro-level, concrete is considered as a homogeneous isotropic material,


even though concrete material properties, especially concrete cracking, are linked with
the lower structural level, i.e. the meso-level. According to RILEM Technical Committee
90-FMA (1989), there are three basic types of failure: tension (mode-I), shear (mode-II)
and tear (mode-III). Since concrete fails most easily in mode-I, cracks tend to branch in
the direction of maximum principle stress. This is why mode-I failure has been the focal
of concrete cracking research and there is little information available about other two
modes. A mode-I crack propagates when the principle tensile stress exceeds the concrete
tensile strength, fct. Near the tip of a crack, there exists a zone, called ‘fracture process
zone’, over which the stress decays with further displacement. The descending branch is

59
CHAPTER 2

characterised by the specific fracture energy, Gf, dissipated during complete crack
formation. It is the area under the complete stress-displacement curve (discrete crack) or
stress-strain curve (crack band) in pure tension. Consequently, concrete cracking is
influenced primarily by two properties: the tensile strength, fct, and the specific fracture
energy, Gf.

The tensile strength of concrete can be determined by a direct tension test, a


cylinder split test or a flexure test. It is more variable than its compressive strength. There
are different expressions recommended by different codes to calculate the concrete tensile
strength. In AS3600 (Standards Australia, 2001), an approximate expression for the
characteristic principle tensile strength at 28 days is given as:

f ' ct = 0.4 f ' c (2.50)

where both strengths are in MPa. An estimate of the characteristic flexural tensile
strength can be made using this expression:

f ' cf = 0.6 f ' c (2.51)

According to Comite Euro-International du Beton (1991), the concrete characteristic


tensile strength lies within a range of which the lower and upper bound values may be
estimated from the characteristic compressive strength using the following equations:
2/3
⎛ f' ⎞
f ' ct ,min = f ' ct 0,min ⎜⎜ c ⎟
⎟ (2.52a)
⎝ f 'c0 ⎠
2/3
⎛ f' ⎞
f ' ct ,max = f ' ct 0,max ⎜⎜ c ⎟
⎟ (2.52b)
⎝ f 'c0 ⎠
where f’c0 = 10 MPa; f’ct0,min = 0.95 MPa; f’ct0,max = 1.85 MPa. The relationship between
the characteristic compressive strength, f’c, and the mean compressive strength, fcm, may
be estimated by
fcm = f’c + 8 (2.52c)
where all strengths are in MPa. The mean value of tensile strength may also be estimated
from this equation:
2/3
⎛ f' ⎞
f ctm = f ' ct 0,m ⎜⎜ c ⎟⎟ (2.52d)
⎝ f 'c0 ⎠
where f’ct0,m = 1.40 MPa.

60
LITERATURE REVIEW

The table below summarises the predictions of the tensile strength of the concrete
with an average compressive strength of 53.7 MPa, which is mainly used in the present
research.

Table 2.2 Different predictions for concrete tensile strength


Tensile strength
Source
(MPa)
AS 3600 principle tensile strength 2.93
AS 3600 flexure tensile strength 4.40
CEB-FIB 90 lower bound 2.60
CEB-FIB 90 upper bound 5.10
CEB-FIB 90 mean 3.86

The fracture energy of concrete, Gf, can also be variable. It is relatively difficult to
be predicted and therefore there are very limited formulations recommended in the
current codes. The fracture energy can be found using testing. There have been a large
number of tests from different laboratories, based on different fracture specimens and
different testing methods. A wide range of values for Gf has been reported (Bazant and
Becq-Giraudon, 2002; Wittmann, 2002). Gf has been found to be dependant on several
parameters, including but not limited to the aggregate sizes and properties, the water-
cement ratio, the test method and the specimen size (RILEM Technical Committee 90-
FMA, 1989; Wittmann et al., 1990; Bazant, 2002; Wittmann, 2002). To estimate the
fracture energy from the basic characteristics of concrete, a number of empirical formulae
have been proposed. In the following expressions, the concrete strength and modulus are
in MPa, the maximum aggregate size, da, is in mm and the fracture energy, Gf, is in
N/mm.

A simple formula for the mean fracture energy was developed by Bazant and Oh
(1983) on the basis of the data of notched specimens. The energy is given by

( )
G f = 2.5 2.72 + 0.0214f ' ct f ' ct
2 da
Ec
(2.53)

where f’ct is the concrete tensile strength and Ec is the elastic modulus of the concrete.

61
CHAPTER 2

Another mean estimate of Gf is recommended in CEB-FIB Model Code 1990


based on the test data reported in the literature up to the late 1980s, as follows
0.7
⎛f' ⎞
G f = (0.0469d − 0.5d a + 26)⎜⎜ c ⎟⎟
2
a (2.54)
⎝ 10 ⎠
The code allows a 30 % variation for the value calculated using the formula.

A simpler formula relating Gf to concrete compressive strength is reported in van


Mier (1997):
Gf = 0.97 fc + 41.8 (2.55)

Trunk and Wittmann (1998) studied the relationship between the fracture energy
and the maximum aggregate size. They proposed a power function to calculate Gf, as
follows
Gf = a . dan (2.56)
From their data, they found that a = 80.6 and n = 0.32.

In a recent study by Bazant and Becq-Giraudon (2002), the fracture energy was
stated to be influenced by the maximum aggregate size, the water-cement ratio, w/c, the
concrete elastic modulus, the unit weight of the concrete and the concrete strength. From
a relatively large database, the researchers derived a statistical formula to compute the
concrete fracture energy as the following:
0.46 0.22 −0.30
⎛ f' ⎞ ⎛ d ⎞ ⎛w⎞
G f = 2.5α 0 ⎜⎜ c ⎟⎟ ⎜1 + a ⎟ ⎜ ⎟ (2.57)
⎝ 0.051 ⎠ ⎝ 11.27 ⎠ ⎝c⎠

where the coefficient α0 equals 1.12 for crushed aggregates. The coefficient of variation
was found to be 29.9%.

These formulae produce different values for Gf. For example, the table below
summarises the predictions for the concrete with the compressive strength of 53.7MPa
with the maximum aggregate size of 14 mm and the water/cement ratio of 0.59. These are
the properties of the concrete used mainly in the present research.

62
LITERATURE REVIEW

Table 2.3 Different predictions for fracture energy


Source Gf (N/mm) Allowed variation
Bazant and Oh (1983) 0.039 N/A
CEB-FIB 90 0.091 30%
van Mier (1996) 0.094 N/A
Trunk and Wittmann (1998) 0.188 N/A
Bazant and Becq-Giraudon (2002) 0.124 29.9%

2.4.1.2 Concrete modelling


In modelling of concrete, compressive and tensile behaviour are often considered
separately. For concrete under uniaxial compression, the relation between stress and
strain is described by a non-linear curve (Hognested, 1951; Desayi and Krishnan, 1964;
Thorenfeldt et al., 1987). Under biaxial or triaxial compression, concrete shows a
pressure-dependent behaviour, i.e. the strength and ductility increase with increasing
isotropic stress. The compressive stress-strain relationship for the uniaxial case can be
modified to incorporate that behaviour. To model concrete in tension, there are basically
two main approaches: smeared crack approach and discrete crack approach. A detailed
description of these approaches can be found in RILEM Technical Committee 90-FMA
(1989). Briefly, in the first approach, cracks are assumed to be distributed across an
element. It is achieved by reducing the stiffness of the element in the direction of the
principle strain when the principle tensile strain in an element reaches a limiting value.
This approach has been used widely because it enjoys the advantage of permitting
automatic crack propagation with a relatively small effort. However, since this method
maintains displacement continuity across a crack, a realistic prediction of the distribution
of strains in regions adjacent to cracks can be difficult. In the second approach, discrete
cracking, displacement discontinuities across a crack are accounted for by disconnecting
elements at the nodal points along their boundaries. The main problem in this approach,
however, is the difficulty arising from the introduction of additional nodal points required
by the altered topology of the analytical model. The cracks usually have to be inserted
manually into the mesh.

63
CHAPTER 2

2.4.2 Finite element modelling of shear-lap tests


Maeda et al. (1997) carried out a nonlinear finite element analysis to simulate
their double shear-lap tests. A FE code called WCOMR was used, which adopted a
smeared crack model. A relatively course mesh was used. Bond elements were inserted
between the CFRP and concrete elements. The analysis was to simulate the strain
distributions only and the ultimate loads were not predicted. The comparison showed
good agreement at low load levels but large discrepancy at higher loads. The authors
stated that further study was necessary.

(a)

8000 3kN (exp.)


5kN (exp.)
9kN (exp.)
6000
Microstrain

3kN (FEM)
5kN (FEM)
9kN (FEM)
4000

2000

0 50 100 150
Location (mm)

(b)
Figure 2.30 Finite element mesh adopted (a) and comparison of strain distributions
reported (b) by Maeda et al. (1997)

Chen et al. (2001) carried out a study on the stress distributions in several
common test arrangements using a linear elastic finite element analysis. A very fine mesh
was used near the pulling end and the plate end (Figure 2.31). The stress concentrations in
the plate-adhesive interface were found showing the peaks within a very short distance
(0.5 mm) near the plate ends. The stress variation across and along the composite plate

64
LITERATURE REVIEW

were also investigated. It was found that the stress difference on the upper and lower
surfaces gradually decreased and became insignificant at a certain distance from the
pulling end. The effect of the support block clearance in the pushing tests was also
studied. The maximum shear stress near the pulling end was not significantly affected
once the clearance was greater than 37.5 mm. Chen et al. also found that the bond
strength could be dependant on the test methods. The bending tests gave higher bond
strengths than those obtained using shear tests. Little difference was found between the
corresponding double and single tests. However, these findings were only for the linear
elastic range and cracking in the concrete was not accounted for in the analyses.

Pulling end Plate end

Plate

Adhesive

Concrete

Figure 2.31 A typical FE mesh adopted by Chen et al. (2001)

A recent investigation was by Lee (2003), where concrete cracking was taken into
account by smeared cracks modelling. Lee showed similar normal cracks in the concrete
between his FE model and experiment and was able to predict the ultimate failure loads
for his series of specimens. However, interfacial cracks along the interface were not
observed even though it was the dominant failure surface observed in his experiments.

65
CHAPTER 2

CFRP nodes free to move in


Y-direction, fixed in X-direction
CL Adhesive element
CFRP element
Concrete element

X CL
- Nodes free to move in X-direction, fixed in Y-direction
- Uniform displacement applied along the nodes

Figure 2.32 The FE mesh (above) and the crack patterns for two example specimens
with bond lengths of 50 mm (below left) and 130 mm (below right) as reported in
Lee (2003)

Camata et al. modelled Lee’s tests using a different technique. The researchers
used a combination of smeared and discrete crack approaches. The analyses were carried
out using the program MERLIN (MERLIN User's Manual, 2002). An auto-remeshing
technique was used to find the crack paths in the concrete blocks. The researchers were
able to simulate shear-tension failure and reported good agreement between the numerical
analysis and the experimental results for a number of specimens.

Figure 2.33 Boundary conditions and mesh discretization by Camata et al. (2003)

2.4.3 Finite element modelling of beam tests


A number of numerical studies of retrofitted beam testing have been reported in
the literature. Both two- and three-dimensional models were built and both linear and

66
LITERATURE REVIEW

nonlinear analyses were carried out. To simulate debond failures, both discrete and
smeared crack modelling were used.

Malek et al. (1998) and Teng et al. (2002) built linear elastic models to investigate
the stress distributions in RC beams strengthened with a bonded soffit plate. Malek et al.
used the FE analysis to verify their theoretical predictions of the local interfacial stress
concentrations near the FRP cut-off point and near the tip of a flexural crack. Relatively
fine meshes were used at those locations. Different meshes with different element types
and different mesh sizes were chosen. The interfacial shear and normal stress, and the
tensile stress of the plate predicted by the FEA were in good agreement with the
theoretical solution. Teng et al. adopted a much finer mesh (Figure 2.34a). They were
able to show that the stresses vary strongly in the adhesive layer. In particular, near the
end of the plate, the interfacial normal stress is tensile along the adhesive-concrete
interface but compressive along the FRP-adhesive interface (Figure 2.34b). Teng et al.
also found that the effect of a spew fillet at the end of the plate formed from excess
adhesive can reduce the interfacial stresses. They also pointed out that a closed-form
solution based on the assumption of uniform stresses in the thickness direction is
incapable of predicting accurately such stress concentrations but can provide reasonably
close predictions of stresses along the middle-thickness section of the adhesive layer.

67
CHAPTER 2

(a)

Normal stress Shear stress


(b)
Figure 2.34 A FE mesh adopted (a) and the stress contours near the plate end
reported (b) by Teng et al. (2002)

Niu and Wu (2001a) and Niu and Wu (2001b) built FE models based on discrete
cracks to study interfacial peeling of FRP from flexural and flexure-shear cracks. The
main cracks were inserted manually. In the first paper, the commercial FE package
ABAQUS was used and the main focus was on the interfacial stress distributions in
beams with flexural cracks. Concrete was assumed to be linear elastic. The cracks were
simulated using a tension-softening model. The researchers compared the FE results with
their theoretical solutions and found good agreement. No comparison with experimental
results was however made. In the second paper, the FE package DIANA version 7.2 was
used to simulate the behaviour of a retrofitted beam under testing. Both flexural and
flexure-shear cracks were inserted in the model as shown in Figure 2.35a. Plane stress
and truss elements were used to model concrete and FRP, respectively. Concrete material
was modelled as nonlinear in compression. The interfacial fracture behaviour was
simulated using bilinear relationships between the bond stress and the slip and between
the normal stress and the normal displacement. The researchers found that the

68
LITERATURE REVIEW

propagation of peeling depends mainly on the interfacial facture energy consumed for
mode II. Comparison of the load-displacement curves was done for two beams as shown
in Figure 2.35b. The predicted curves were found to have a larger stiffness than those of
the experiments. That was attributed to the fact that the model did not take into account
distributed cracks in concrete.

(a)

(b) (c)
Figure 2.35 FE discretization model (a), loading curve (b) and FRP stress
distributions (c) as reported by Niu and Wu (2001)

A number of nonlinear analyses using smeared cracks have also been carried out.
Arduini et al. (1997) modelled RC beams strengthened with CFRP plates and sheets using
ABAQUS. A two-dimensional mesh was used for the beams strengthened with CFRP
plates, while the beams strengthened with CFRP sheets were modelled in three
dimensions. The concrete elements used were eight-node and twenty-node, respectively.
The meshes were relatively coarse (50 and 100 mm mesh size). A perfect bond was
assumed between the FRP composite and concrete. The researchers reported that the
numerical models showed good accordance to the experimental results in terms of the
load-deflection response, the load-FRP strain response and the evolution of cracking.
However, the assumption of a perfect bond and the course mesh used led to an
overestimation of the beam stiffness. Debond cracks were not observed. In addition,
convergence difficulty was encountered and therefore the models fell short in the
identification of the maximum load carrying capacity of the beams.

69
CHAPTER 2

David et al. (1999) used a FE code developed in France to model their tests on
beams retrofitted with CFRP. These beams failed by end cover peeling. Three-node
triangular elements and two-node linear elements were used to model the concrete and the
CFRP, respectively. An elasto-plastic model with a softening post-peak phase was used
for the concrete in compression. In tension, the concrete was modelled using a smeared
fixed crack approach. The CFRP-concrete interface was assumed to have a perfect elastic
behaviour. The author reported that the model was able to predict correctly the flexural
behaviour of the strengthened beams. However, comparison of only one loading curve
was showed. The crack patterns and predicted failure modes were not presented.

Ross et al. (1999) modelled their beams using another commercial package
(ADINA), which is capable of nonlinear modelling. Eight-node plane stress elements and
three-node truss elements were used to represent the concrete and the FRP composite,
respectively. Good agreement with the experiments was observed in terms of the beam
stiffness but the failure mode and therefore the ultimate load were not predicted by the
model. The crack pattern at failure of a beam is illustrated in Figure 2.36, showing no
delamination crack even though it was the dominant failure mode of the beams.

Figure 2.36 A crack pattern at failure reported by Ross et al. (1999)

Hearing (2000) used the same FE package (ADINA) and built different two-
dimensional models with different mesh types and mesh sizes to simulate debonding
failure. The nonlinear FE models were able to produce the global behaviour, i.e. the load-
deflection stiffness, of the specimens tested by the researcher. However, the analyses
were not able to capture local debond mechanisms.

Another FE study using a smeared crack approach was done by Rahimi and
Hutchinson (2001). The cracking model used was incorporated with an isotropic damage
model to simulate the nonlinear behaviour of the concrete. Four- or eight-node

70
LITERATURE REVIEW

quadrilateral isoparametric elements were used for concrete. Three- or six-node elements
were also needed in the transition zones. The internal rebars were modelled with two- or
three-node bar elements, while the adhesive layer and external reinforcement were
modelled with a single row of four- or eight-node elements (Figure 2.37). The adhesive
was assumed to be elastic. The FE analysis was used mainly to supplement the
experimental measurement. The FE predictions showed two peak interfacial shear
stresses near the FRP end and under the load point. However, some discrepancy with the
experiments was observed. The FE analyses were also found to be sensitive to the
concrete tensile strength. The beam strengths predicted were found to be within 20 % of
the test results. No crack pattern was reported.

Figure 2.37 A typical two-dimensional model of retrofitted beams adopted by


Rahimi and Hutchinson (2001)

Hormann et al. (2002) used two- and three-dimensional models to simulate


flexural crack debonding in their concrete slabs strengthened with CFRP plates. A perfect
bond was assumed between the plate and concrete. The researchers found that the three-
dimensional model predicted the slab stiffness better than the two-dimensional one.
Debonding was not captured and in order to predict the ultimate load, the researchers
manually reduced the CFRP tensile strain capacity by 65 %.

Fanning and Kelly (2000) used the smeared crack model incorporated in the three-
dimensional eight-node solid isoparametric element, Solid65, in the commercial package
ANSYS to model both the concrete and the adhesive layer. Four-node isoparemetric

71
CHAPTER 2

linear elastic shell elements were utilised to represent the CFRP plates. The researchers
reported that their model was able to simulate the strain distributions in the CFRP
composites and the load-deflection curves for a number of beams in their test series. A
crack pattern was compared for a beam failed in shear. However, the cracking patterns for
the beams failed in end cover peeling were not reported and the deflection and plate strain
were not showed for most of the beams failing by end cover peeling.

Another attempt to model retrofitted RC beams using ANSYS was by Shokrieh


and Malevajerdy (2001). They modelled their tested beams, which were retrofitted with
GFRP plates with end anchorage. However, the model was not capable of simulating
propagation of debonding failure observed in their experiments.

Zarnic and Bosiljkov (2001) used a nonlinear three-dimensional FEA program


where tensile cracking was the only nonlinearity considered. They modelled the beams
and slabs strengthened with CFRP and steel plates reported in Zarnic et al. (1999). A
relatively coarse mesh was used. The main beam body and the concrete cover were
represented by two rows and one row of elements, respectively. It was not clear how the
adhesive was modelled. The stiffness predicted was relatively higher than the test results.
Post peak behaviour and therefore the ultimate capacities were not successfully modelled.

Vecchio and Bucci (1999) developed a nonlinear FE model to analyse repaired


structures by taking account the chronology of the loading, damage and repair. The
methodology defined and employed plastic strain offsets in the context of a smeared
rotating crack model. The ability to engage and disengage elements at various stages of
loading was utilised to model several flexural members retrofitted after they were loaded
and subjected to a certain damage. Shear failure was reported to be simulated correctly. A
reasonably good correlation was achieved.

Aprile et al. (2001b) adopted a different approach to model shallow retrofitted


beams. They used a displacement-based fibre model with bond-slip (Figure 2.38). This
model was able to properly describe the interfacial shear stress distributions that arise
during loading of the RC beam up to failure. An elastic-brittle bond law for the plate-
concrete interface was introduced. Nonlinear behaviour of the materials including the
tension stiffening of concrete in tensile areas was also taken into account. The peeling

72
LITERATURE REVIEW

stresses normal to the rebars and to the external plates were however neglected. The
researchers modelled the same beams reported in Zarnic et al. (1999). Two failure modes,
midspan and end debond, were indicated in the simulation by looking at the bond stress
distributions. However, due to the nature of the models, the failure surfaces were assumed
to be the bond surfaces and the capacities were determined by an assumed strength of the
epoxy resin. Close match was found for the loading curves of all specimens.
y

U31 U22
U21 vB (x) U32
uB (x)
x z
U11 U12
u (x)

U41 U42

Figure 2.38 Node and element displacements for RC beam model with slip in plate
by Aprile et al. (2001)

Wong and Vecchio (2003) also modelled Zarnic et al.’s beams. The researchers
used a smeared rotating crack model to analyse the beams and bond elements to simulate
debonding. Good agreement with the experiments was achieved when the maximum bond
stress was taken as 3 MPa (Figure 2.39).

CFRP (FE)
(perfect bond)
CFRP (FE)
(τmax = 3 MPa)
Total load (kN)

CFRP (Exp.)

Control (Exp.)

Midspan deflection (mm)


Figure 2.39 Comparison of load-deflection curves for Zarnic et al.’s beams by Wong
and Vecchio (2003)

A similar approach was adopted by Wong et al. (2001). This research group built
a three dimensional FE model of FRP plated beams tested by Rahimi and Hutchinson

73
CHAPTER 2

(2001) using a general-purpose FE package (MARC). Eight-node solid elements, four-


node membrane elements and eight-node rebar elements were used for concrete, FRP and
steel rebars, respectively (Figure 2.40). Concrete cracking was simulated using a smeared
crack model. Linear bond-slip relationships were assumed in both tangential and normal
directions for the FRP-concrete interface. The maximum bond stress allowed was chosen
to be 3.2 MPa. The maximum normal traction was 5.2 MPa, which was higher than the
concrete tensile stress adopted (4.3 MPa). Reasonably close predictions were found with
the error less than 7.3 %. However, the failure surface was forced to be in the interface.
No crack pattern was reported.

Figure 2.40 The three-dimensional model of FRP plated beams adopted by Wong et
al. (2001)

Yang et al. (2002) used a discrete crack model based FEA. Linear elastic fracture
mechanics (LEFM) was used in conjunction with a remeshing algorithm. A computer
program based on AUTOFRAP was specially developed to incorporate these features. In
the LEFM approach, all the materials involved in the plated RC beams were assumed to
be linear elastic in both tension and compression. Since LEFM is meant to be applicable
to large concrete structures only, the researchers argued that acceptable accuracy may still
be achieved for normal-sized structures if energy based criteria are properly applied.

In Yang et al.’s model, the bond between the concrete and internal reinforcements
was modelled using interface elements. A perfect bond was assumed initially. Once a
crack intersected an internal rebar, a full loss of bond over the length of one

74
LITERATURE REVIEW

reinforcement element was assumed. The bond-slip behaviour between the concrete-
adhesive and adhesive-plate was modelled using four-node contact elements. A maximum
energy release rate criterion was adopted for the crack propagation condition. To allow
multiple crack propagation, Yang et al. used a combination of two loading algorithms, the
G-scaling procedure and the displacement control procedure, to avoid the problems
associated with each individual method. The G-scaling procedure was used at first to
allow only one crack to initiate at each step. In this procedure, the load level and the
corresponding structural responses were scaled from a linear elastic analysis to meet the
crack propagation condition. When multiple cracks were initiated at a load step, the
displacement control procedure was used thereafter to allow multiple cracks to propagate
within a single step. It was achieved by scaling the load and the structural responses so
that the displacement satisfied a pre-set value.

Yang et al. modelled an example RC beam retrofitted with a CFRP plate and
failing by end cover peeling. The concrete cover on the tension face was modelled using
four rows of elements. Both the adhesive and the CFRP plate were modelled using one
layer of elements. The crack pattern was simulated successfully (Figure 2.41). The effect
of the plate length on the failure mode was also investigated. Shifting from flexural to end
cover peeling was observed as the plate length was reduced. However, very limited
comparison with the experimental measurements was made. Only the load deflection was
compared but the peak load was not captured accurately.

Figure 2.41 Initial FE mesh (left) and crack pattern at a high load level (right)
reported by Yang et al. (2003)

In summary, the FE studies were able to predict the global behaviour of retrofitted
beams. The load versus deflection curves predicted by most FE models matched the
corresponding experimental ones reasonably well. However, most models were not able

75
CHAPTER 2

to capture local debonding failures. Cracking patterns were rarely reported. Several
studies showed the general crack bands but the debond cracks were not observed. There
were attempts to simulate debonding using a predefined interface behaviour. However, in
this model delamination surface location was assumed to be the bond line between the
FRP composite and concrete. Last but not least, most of the reported FE models lack
experimental verification. Further investigation using nonlinear FEA is therefore still
needed.

2.5 Summary of findings


This chapter has presented reviews on experimental, theoretical and numerical
results of shear-lap and retrofitted beam tests. In the first section of the review, the
common test set-ups, the observed failure modes and the general behaviour of shear-lap
tests were described, followed by a summary of existing prediction formulae, ranging
from empirical to theoretical models. From the review, the following items were
highlighted as areas that require further investigation:

• Investigation on shear-tension failure and effect of vertical cracks and concrete prism
formation
• Further investigation on the effect of FRP dimensions and the bond-slip relations
• Verification of existing prediction models.

In the second section of the review, a summary of strengthening schemes and


possible failure modes of retrofitted RC beams was presented first. It was followed by a
description on the effects of several parameters, in which the important ones were
highlighted including concrete cover, FRP bond length and FRP stiffness relative to
tension reinforcement stiffness. A wide variety of theoretical models were also presented.
The following items were identified as the issues to be addressed:

• Further investigation on the debond failure mechanisms and the effects of a number of
parameters
• Further experimental data to verify existing theoretical models
• Development or recommendation of a reliable design methods.

76
LITERATURE REVIEW

The final section of the review was about FE modelling of retrofitted concrete
members. An outline of concrete mechanical properties and modelling techniques was
presented. It was followed by a critical review of FE modelling of shear-lap and beam
tests. The following tasks were highlighted:

• Further investigation using FE on the bond between the FRP composite and concrete
• Further FE modelling of retrofitted beams to capture different debonding
mechanisms.

77
ASSESSMENT OF PREDICTION MODELS

CHAPTER 3 - ASSESSMENT OF PREDICTION MODELS

3.1 Introduction
In Chapter 2, existing experimental and theoretical studies on shear-lap specimens
and retrofitted beams are presented. It was found that there are a number of theoretical
models to predict the load capacity of a joint between FRP and concrete. These models
are based on different experimental data and/or different theoretical assumptions.
Therefore, there is a need to investigate the accuracy of the models. Similarly, there are
many experimental and theoretical studies on the retrofitted beams. Different failure
modes have been described and different approaches have been adopted to predict the
peak loads of retrofitted beams. It was found that not only these prediction methods need
to be validated, but the failure mechanisms also require further investigation.

In this chapter, existing prediction models are assessed. The assessment of the
models for shear-lap tests was done based on a relatively large database of shear-lap tests
collected from past experiments. The assessment of the models for retrofitted beams was
based on a large database of beam tests from past experiments. For completeness, the data
obtained from the experimental programs undertaken by the author was also included in
the databases. The details of these experiments will be reported in Chapters 4 and 5.

3.2 Assessment of analytical studies of shear-lap specimens


3.2.1 Database
An extensive literature survey was carried out to collect available data on testing
of single and double shear-lap specimens. 156 tests were found. They are summarised in
Table 3.1. More details of these specimens can be found in Appendix A.2. Tests that were
not sufficiently well documented for analysis have been excluded. The variability of test
specimen parameters and test results are shown in Table 3.2.

79
CHAPTER 3

Table 3.1 Summary of the database of shear-lap tests


No of Laminate Failure
Reference Test type
tests type mode1
Chajes et al. (1996) 16 Single GFRP 1, 4
Maeda et al. (1997) 8 Double CFRP 1, 3
Bizindavyi and Neale (1999) 4 Single GFRP/CFRP 1
Xu et al. (2001) 31 Single CFRP 1
Lee (2003) 5 Double CFRP 1
Yao et al. (2005) 66 Single GFRP/CFRP 1, 2
Present study 26 Single CFRP 1, 2
Note: 1) Refer to Figure 2.2

Table 3.2 Variability of test parameters and results in the shear-lap test database
Parameter Minimum Maximum Average
Concrete strength, fc (MPa) 18.9 70 35.5
Concrete width, bc (mm) 100 228.6 138.0
FRP width, bf (mm) 15 100 51.8
FRP modulus, Ef (MPa) 22500 380000 216432
FRP thickness, tf (mm) 0.11 2 0.39
Bond length, Lf (mm) 50 700 137
Relative stiffness, Ef Af / Ec Ac 0.0097 1.754 0.366
Ultimate capacity, P (kN) 3.81 42.8 13.7

As shown in Table 3.1 and Table 3.2, the most common setup is single shear-lap
test (143 specimens) and the most common bonded composite is CFRP (152 specimens).
The most frequently observed failure is mode 1, interfacial debond (138 specimens).
Sixteen specimens failed by mode 2, shear-tension failure. Only one specimen failed by
FRP tensile rupture and one failed by cohesion failure of the adhesive.

In a number of theoretical models used to predict the debonding failure load, the
elastic modulus and thickness of the adhesive are needed. However, in the experiments,
these parameters were often not measured. In those cases, the adhesive thickness was

80
ASSESSMENT OF PREDICTION MODELS

taken as 1.0 mm as observed in the experiments by the author, and the adhesive elastic
modulus was taken as 8500 MPa, as recommended by Smith and Teng (2002).

3.2.2 Validation results


Model accuracy is measured on the basis of two statistical parameters: the average
and the coefficient of variation (COV) of the ratios of the experimental maximum load to
the calculated ultimate capacity, Pexp / Pcal.

The validation results for the first two simple formulae by Van Gemert (1980) and
Chajes et al. (1996) are shown in Figure 3.1. It is clear that the two models do not provide
accurate predictions. Van Gemert’s model is based on an incorrect distribution of the
bond stress. The average of Pexp / Pcal is 1.98 and the coefficient of variation is 53 %. In
Chajes et al.’s model, the ultimate load is dependent on two empirical factors, which were
found by calibrating with their limited experimental data. The model predicts non-
conservative results with an average Pexp / Pcal of 0.62 and a coefficient of variation of
35 %.

Mode 1 Mode 2 Mode 1 Mode 2


Mode 3 Mode 4 Mode 3 Mode 4
50 50
Predicted max. load (kN)
Predicted max. load (kN)

40 40

30 30

20 20

10 10

0 0
0 10 20 30 40 50 0 10 20 30 40 50
Measured max. load (kN) Measured max. load (kN)

(a) (b)
Figure 3.1 Validation results for Van Gemert’s model (a) and
Chajes et al.’s model (b)

Figure 3.2 shows the validation results for the models by Maeda et al. (1997) and
Izumo et al. (1998). Izumo et al.’s model was validated using the specimens bonded with
CFRP only. Both of these models are purely empirical, based on a number of tests done

81
CHAPTER 3

in Japan. However, they give relatively accurate results for most specimens in the
database described above. The average values of Pexp / Pcal are 0.95 and 0.99 for Maeda et
al.’s and Izumo et al.’s models, respectively. The validation results for the model by
Izumo et al. have a high degree of scatter. Their coefficient of variation is 39 % compared
to 21 % for Maeda et al.’s model. Both models also tend to be non-conservative for the
specimens with high bond strengths.

Mode 1 Mode 2 Mode 1 Mode 2


Mode 3 Mode 4 Mode 3 Mode 4
50 50

Predicted max. load (kN)


Predicted max. load (kN)

40 40

30 30

20 20

10 10

0 0
0 10 20 30 40 50 0 10 20 30 40 50
Measured max. load (kN) Measured max. load (kN)

(a) (b)

Figure 3.2 Validation results for Maeda et al.’s model (a) and Izumo et al.’ model
(b)

Figure 3.3 presents the evaluation results for three models based on fracture
mechanics. All three models provide good correlations between Pexp and Pcal. The
coefficients of variation are less than 20 % for all models. The model by Chen and Teng
gives the best result in terms of the accuracy with the average Pexp / Pcal of 1.07; whereas
Niedermeier’s model is the least conservative with the average Pexp / Pcal of 0.73.

82
ASSESSMENT OF PREDICTION MODELS

Mode 1 Mode 2 Mode 1 Mode 2


Mode 3 Mode 4 Mode 3 Mode 4
50 50

Predicted max. load (kN)


Predicted max. load (kN) 40 40

30 30

20 20

10 10

0 0
0 10 20 30 40 50 0 10 20 30 40 50
Measured max. load (kN) Measured max. load (kN)

(a) (b)
Mode 1 Mode 2
Mode 3 Mode 4
50
Predicted max. load (kN)

40

30

20

10

0
0 10 20 30 40 50
Measured max. load (kN)

(c)
Figure 3.3 Validation results for Niedermeier’s model (a), Neubauer and Rostasy’s
model (b) and Chen and Teng’s model (c)

The validation results for the other two models by Kanakubo et al. (2003) and
Ulaga and Vogel (2003) are presented in Figure 3.4. These models also produce relatively
good predictions. The former tends to have non-conservative results with the average
Pexp / Pcal of 0.83; whereas the latter is conservative with the average Pexp / Pcal of 1.20.

83
CHAPTER 3

Mode 1 Mode 2 Mode 1 Mode 2


Mode 3 Mode 4 Mode 3 Mode 4
50 50

Predicted max. load (kN)


Predicted max. load (kN)

40 40

30 30

20 20

10 10

0 0
0 10 20 30 40 50 0 10 20 30 40 50
Measured max. load (kN) Measured max. load (kN)

(a) (b)
Figure 3.4 Validation results for Kanakubo et al.’s model (a) and Ulaga and Vogel’s
model (b)

A summary of the validation results is shown in Table 3.3. Of three empirical


models, Maeda et al.’s model gives the best result. Most semi-empirical models based on
bond-slip relationships perform relatively well. Of all models, Chen and Teng’s model
performs best. Thus, Chen and Teng’s model is recommended to predict the bond
strength of FRP-concrete joint considering its solid theoretical grounding; whereas Maeda
et al.’s model can be also used for a quick check.

84
ASSESSMENT OF PREDICTION MODELS

Table 3.3 Summary of validation results of strength models for shear-lap tests
Tests failed by mode 1
Verified against All tests
only
Coefficient Coefficient
Average of Average of
Statistical parameters of variation of variation
Pexp / Pcal Pexp / Pcal
(%) (%)
Van Gemert (1980) 1.98 53 2.02 54
Chajes et al. (1996) 0.62 35 0.63 34
Maeda et al. (1997) 0.95 21 0.98 19
Izumo et al. (1997) 0.99 39 1.04 35
Niedermeier (1996) 0.73 19 0.74 17
Neubauer and Rostasy
0.92 18 0.94 17
(1999)
Chen and Teng (2001) 1.07 16 1.08 16
Kanakubo et al. (2003) 0.83 20 0.86 18
Ulaga et al. (2003) 1.20 23 1.25 20

3.3 Assessment of analytical studies of beam specimens


3.3.1 Introduction
As reviewed in Chapter 2, there has been a significant amount of research on the
behaviour of RC beams retrofitted with FRP. A large number of experiments have been
carried out and a variety of theoretical models have been developed. Since most models
have either not been verified or only been verified with a limited number of tests, it is
difficult to justify their credibility.

Therefore, this section presents an assessment of all those models together. The
assessment was based on a large database of the results from testing of simply supported
retrofitted RC beams done by the author and several other research groups. In the
following sections, some description on the database is presented first. It is followed by
the validation and assessment results of the strength models for three main failure
categories: flexure failure, intermediate span debond and end debond.

85
CHAPTER 3

3.3.2 Existing database and new database


There have been several attempts to collect data of testing on RC beams retrofitted
with FRP or steel plates. Bonacci and Maalej (2001) listed 127 beams and summarized
the general behavioural trends. The failure modes were grouped into beam shear failure,
flexural failures (concrete crushing and FRP rupture) and debonding. The beams with
special anchorage systems were not separated from the others. Mukhopadhyaya and
Swamy (2001) collected data from 26 beams with bonded steel plates, 20 beams with
bonded GFRP plates and 23 beams with bonded CFRP plates. They focused mainly on
the variability of the interfacial shear stress between the concrete and externally bonded
plates. El-Mihilmy and Tedesco (2001) gathered the test results of 26 beams strengthened
with FRP composites and failing by end debond. They used the database to assess six
theoretical models. The most detailed study was by Smith and Teng (2002). A relatively
large database was built including 44 beams failing by end cover peeling, 8 beams failing
by end interfacial delamination and 7 beams failing by a combination of those two.
Validation with a number of theoretical models for end debond was carried out and
relatively scattered results were found. Lastly, a small database of beams failing by
flexural crack debond was compiled by Teng et al. (2001). It included 8 beams and 9
slabs. Teng et al. used the database to verify their strength model for flexural crack
debond.

There are a number of reasons to build a new database. Firstly, even though a
number of studies have showed that the beam theory can predict the ultimate loads for
flexure failures, i.e. concrete crushing or FRP rupture, the conclusion made was based on
a very limited number of test specimens and therefore it needs to be consolidated with a
larger number of tests. Secondly, the main focus of the existing databases is either on the
general beam behaviour or on end debond. There have been very limited assessment
studies on intermediate span debond. Lastly, a large number of new tests have become
available recently for end cover peeling (including tests by the author). Further validation
for this mode is therefore possible to confirm the previous findings.

To build the database, an extensive literature survey was carried out. For inclusion
in the database, the beam chosen must satisfy following requirements. It was of a
rectangular section, simply supported and loaded in three or four-point bending. No
preloading was applied. The laminate material was FRP composite only and not

86
ASSESSMENT OF PREDICTION MODELS

prestressed. For the beam failing by debonding, no form of corresponding anchorage was
present.

A list of the experimental studies included in the database is shown in Table 3.4
together with the number of beams tested and the failure modes observed. The
classification of the failure modes was based mainly on the photos and/or descriptions
provided. More details of those beam tests can be found in Appendix A.3. In total, 181
beams were found. 36 beams failed due to flexural failures, in which 19 beams failed by
concrete crushing and 17 failed by FRP rupture. 58 beams failed by intermediate span
debond, in which 43 beams failed due to flexure-shear crack debond, 14 beams due to
flexural crack debond and one beam due to shear crack debond. 82 beams failed by end
debond, in which the only mode observed was end cover peeling. The variability in some
of the test parameters is shown in Table 3.5.

87
CHAPTER 3

Table 3.4 Summary of the database of beam tests

No. of beams failing in certain mode


Total
No. Reference No. of Mix of
Crushing
test(s) IS End end &
of tensile
debond debond IS
rupture
debond
1 Ritchie et al. (1991) 8 3 0 5 0
2 Saadatmanesh and Ehsani (1991) 1 0 1 0 0
3 Triantafillou and Plevris (1992) 2 2 0 0 0
4 Chajes et al. (1994) 8 8 0 0 0
5 Sharif et al. (1994) 1 1 0 0 0
6 Quantrill et al. (1996) 3 0 0 3 0
7 Arduini et al. (1997) 4 1 1 2 0
8 He et al. (1997) 2 0 0 2 0
9 Garden et al. (1997) 6 0 0 6 0
10 Mukhopadhyaya et al. (1998) 1 1 0 0 0
11 Garden et al. (1998) 6 0 3 3 0
12 Ahmed and Van Gemert (1999) 8 1 0 7 0
13 Beber et al. (1999) 6 0 0 6 0
14 David et al. (1999) 4 0 0 4 0
15 Hau (1999) 3 0 0 3 0
16 Tumialan et al. (1999) 7 0 4 3 0
17 Ross et al. (1999) 9 9 0 0 0
18 Bonacci and Maalej (2000) 1 0 1 0 0
19 Gao et al. (2001) 3 0 0 3 0
20 Fanning and Kelly (2001) 4 0 0 4 0
21 Zarnic and Bosiljkov (2001) 1 0 0 1 0
22 Rahimi and Hutchinson (2001) 12 6 2 4 0
23 Nguyen et al. (2001) 5 1 0 4 0
24 Kishi et al. (2001) 8 0 8 0 0
25 Pornpongsaroj and Pimanmas (2003) 3 0 1 2 0
26 Takahashi and Sato (2003) 3 1 2 0 0
27 Valcuende and Benlloch (2003) 4 0 4 0 0
28 Leong and Maalej (2003) 10 0 10 0 0
29 Smith and Teng (2003) 7 0 0 7 0
30 Khomwan et al. (2004) 2 0 2 0 0
31 Brena and Macri (2004) 11 0 11 0 0
32 Present study: Preliminary experiment 2 0 1 1 0
33 Present study: Main experiment 21 0 7 11 3

Note: IS debond = Intermediate span debond

88
ASSESSMENT OF PREDICTION MODELS

Table 3.5 Variability of some test parameters in the beam test database
Parameter Minimum Maximum Average
Beam span, L (mm) 812 6000 1972
Beam aspect ratio, bc/h 0.45 1.67 0.78
Ratio of FRP length to beam depth, Lf/h 0.50 5.40 2.99
Relative stiffness, EfAf / EsAs 0.03 1.06 0.27
Concrete strength, fc (MPa) 25 80 45.48
FRP modulus, Ef (MPa) 10343 400000 150754

In the theoretical models used to predict debonding modes, there were a few
parameters, which were not measured or mentioned in a number of experimental studies.
In those cases, the parameters were assumed as follows: The elastic modulus of steel, Es,
was taken as 200000 MPa. The concrete elastic modulus was taken as 5050 f c

according to AS3600 (Standards Australia, 2001). The average concrete tensile was
calculated according to CEB-FIB Model Code 1990 (Comite Euro-International du
Beton, 1991) The adhesive thickness was taken as 1 mm as observed in the experiments
by the author. The elastic modulus of the adhesive was taken as 8500 MPa as
recommended by Smith and Teng (2002).

Similar to the validation study of the shear-lap tests, model accuracy is determined
based on two statistical parameters: the average and the coefficient of variation of the
ratios of the capacity measured in the experiments to the predicted ultimate capacity
Mexp / Mcal or Vexp / Vcal.

3.3.3 Validation of beam theory


The implementation of the beam theory was carried out using the formulation
described in Appendix A.1. The validation results are shown in Figure 3.5. It is clear that
the beam theory can predict the sectional moment capacity with relatively good accuracy.
The average of the ratio of the experimental to the predicted moment capacity Mexp / Mcal
is 0.97 with the coefficient of variation being 16 %.

89
CHAPTER 3

120

100

80

Mcal (kN.m)
60

40

20

0
0 20 40 60 80 100 120
Mexp (kN.m)

Figure 3.5 Validation of the beam theory

3.3.4 Validation of existing models to predict intermediate span debond


The nature of all intermediate span debond sub-modes (flexural crack debond,
shear crack debond and flexure-shear crack debond) is similar. They are the result of the
high tensile stress level in the composite near the tip of a crack in the intermediate span.
As a result, most theoretical models were proposed to address all sub-modes together. In
a number of reported experimental studies, identifying different sub-modes of
intermediate span debond was also difficult due to insufficient reporting. Therefore, in
this section, validation of the prediction models was done for all sub-modes of
intermediate span debond.

To assess the debonding strength models, sectional analysis was carried out using
the beam theory (described in Appendix A.1). For the cases when the predicted
debonding failure load exceeded the ultimate load based on the sectional moment
capacity, the latter was used. To find the failure load, a backward substitution method was
used. A procedure was set up using VBA in Microsoft Excel 2002 to vary the applied
shear load until a failure ratio was close to 1. The failure ratio was defined as the ratio of
failure criteria to the corresponding capacity. The allowable error was 1 %. Checks were
also carried out for the parameter limits and the results were ignored if a requirement was
not satisfied.

The validation results of the models by Arya and Farmer (2001) and Shehata et al.
(2001) are shown in Figure 3.6. It was found that the FRP strain limit was always critical

90
ASSESSMENT OF PREDICTION MODELS

in Arya and Farmer’s model. The steel strain did not exceed 0.5εsy for any specimen.
Despite the simplification made, the predicted failure loads agree well with the
experimental results. The average of the ratios Vexp/Vcal is 1.08 with the coefficient of
variation being 15 %. However, for some beams, the calculated load is not conservative.
For Shehata et al.’s model, since the allowable strain limit was reduced from 0.006 to
0.005, the average of Vexp/Vcal increases to 1.15 and the predictions are conservative for
most cases.

350 FSCD 350 FSCD


FCD FCD
SCD SCD
300 300

250 250
Vcal (kN)

Vcal (kN)
200 200
150 150
100 100
50 50
0 0
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
Vexp (kN) Vexp (kN)

(a) Arya and Farmer (2001) (b) Shehata et al. (2001)


Note: 1) FSCD: Flexure-shear crack debond; FCD: Flexural crack debond; SCD: Shear crack debond

Figure 3.6 Validation results for Arya and Farmer’s model (a) and Shehata et al.’s
model (b)

The validation results of the models by Maruyama and Ueda (2001) and Teng et
al. (2001) are illustrated in Figure 3.7. Validation of Maruyama and Ueda’s model was
done assuming that the fracture energy was 0.5 N/mm. The model is generally
conservative with the average of the ratios Vexp/Vcal being 1.47 and the coefficient of
variation being 20 %. Very similar results are found for Teng et al.’s model.

91
CHAPTER 3

350 FSCD 350 FSCD


FCD FCD
SCD SCD
300 300
250 250
Vcal (kN)

Vcal (kN)
200 200
150 150
100 100
50 50
0 0
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
Vexp (kN) Vexp (kN)

(a) Maruyama and Ueda (2001) (b) Teng et al. (2001)


Figure 3.7 Validation results for Maruyama and Ueda’s model (a) and Teng et al.’s
model (b)

Validation results for the models by Blaschko (1997) and Mohamed Ali (2000)
are plotted in Figure 3.8. These models produce relatively scattered results with the
coefficient of variation being 23 % for both cases. Mohamed Ali’s model is very
conservative. The average of Vexp/Vcal for this model is 2.70.

350 FSCD 350 FSCD


FCD FCD
SCD SCD
300 300

250 250
Vcal (kN)

Vcal (kN)

200 200
150 150

100 100

50 50

0 0
0 50 100 150 200 250 300 350 0 50 100 150 200 250 300 350
Vexp (kN) Vexp (kN)

(a) Blaschko (1997) (b) Mohamed Ali (2000)


Figure 3.8 Validation results for Blaschko (1997) model (a) and Mohamed Ali
(2000) model (b)

92
ASSESSMENT OF PREDICTION MODELS

The graphical comparison of the predictions by Niedermeier (2000) with the


experimental data is shown in Figure 3.9. The model gives very scattered and generally
conservative results. It was also found that the checks proposed by Matthys (2000) are
usually not critical. The ultimate shear loads predicted by this model were higher than the
ultimate shear capacities based on sectional bending capacities for most cases and,
therefore, no graphical output is presented here.

350 FSCD
FCD
SCD
300
250
Vcal (kN)

200

150

100
50

0
0 50 100 150 200 250 300 350
Vexp (kN)

Figure 3.9 Validation results for Niedermeier’s model

3.3.5 Validation of existing models to predict end debond


The validation results are reported below in the same order as presented in
Chapter 2. A backward substitution method was used to find the ultimate debond load in
a similar manner as the previous validation study for intermediate span debond.

The simple model by Jones et al. for steel plated beams does not work well for
FRP plated beams. As the predicted failure loads are higher than the ultimate shear
capacities assuming no debond for most cases, the validation results are invalid and not
presented here. Shehata et al.’s model produces relatively good results (Figure 3.10). The
average of Vexp/Vcal for this model is 1.15 with 22 % variation.

93
CHAPTER 3

160
140
120
100

Vcal (kN)
80
60
40
20
0
0 20 40 60 80 100 120 140 160
Vexp (kN)

Figure 3.10 Validation results for Shehata et al.’s model

To verify Nguyen et al.’s model, the limiting strain was taken as the lowest value
found by the researchers, i.e. 0.0017. Similarly for Fanning and Kelly’s model, the limit
strain slope was taken as 4.2 strain/mm × 10-6. The validation results for these two models
are shown in Figure 3.11. Both models produce slightly scattered results with the
coefficient of variation being 25 and 26 %, respectively. Nguyen et al.’s model provides
relatively more accurate predictions with the average of Vexp/Vcal being 1.02.

200 200

160 160
Vcal (kN)

Vcal (kN)

120 120

80 80

40 40

0 0
0 40 80 120 160 200 0 40 80 120 160 200
Vexp (kN) Vexp (kN)

(a) Nguyen et al. (2001) (b) Fanning and Kelly (2001)


Figure 3.11 Validation results for Nguyen et al.’s model (a) and Fanning and
Kelly’s model (b)

94
ASSESSMENT OF PREDICTION MODELS

Figure 3.12 shows the comparison of the predictions of end debond strengths
using the models by Saadatmanesh and Malek (1998), Tumialan et al. (1999) and El-
Mihilmy and Tedesco (2001) with the experimental data. It is clear that these models do
not provide accurate predictions. The average of Vexp/Vcal ranges from 1.77 to 4.27 with
the coefficient of variation ranging from 26 to 35 %.

160 160
140 140
120 120
100 100
Vcal (kN)

Vcal (kN)
80 80
60 60
40 40
20 20
0 0
0 20 40 60 80 100 120 140 160 0 20 40 60 80 100 120 140 160
Vexp (kN) Vexp (kN)

(a) Saadatmanesh and Malek (1998) (b) Tumialan et al. (1999)

160
140
120
100
Vcal (kN)

80
60
40
20
0
0 20 40 60 80 100 120 140 160
Vexp (kN)

(c) El-Mihilmy and Tedesco (2001)


Figure 3.12 Validation results for Saadatmanesh and Malek’s model (a), Tumialan
et al.’s model (b) and El-Mihilmy and Tedesco’s model (c)

Validation results for the shear capacity based strength models are shown in
Figure 3.13 to Figure 3.15. These models tend to give less scattered results with the
variability of around 15 %. Oehlers’ and Smith and Teng’s model base on the shear

95
CHAPTER 3

bending moment interaction diagrams and produce similar results. Both models provide
moderately conservative predictions. The averages of Vexp/Vcal are 1.82 and 1.84,
respectively. The models by Jansze (1997) and Ahmed and Van Gemert (1999) are less
conservative with the corresponding averages of Vexp/Vcal being 1.44 and 1.21,
respectively. In Ziraba et al.’s model, the beam shear strength comprises of the shear
contribution of plain concrete and of the modified shear contribution of stirrups. The
latter was found to be insignificant for most cases. Therefore, the results for this model
were found to be similar to those of other shear based models, where the shear
contribution of stirrups was not considered.

160 160
140 140
120 120
100 100
Vcal (kN)

Vcal (kN)

80 80
60 60
40 40
20 20
0 0
0 20 40 60 80 100 120 140 160 0 20 40 60 80 100 120 140 160
Vexp (kN) Vexp (kN)

(a) Oehlers (1992) (b) Smith and Teng (2003)


Figure 3.13 Validation results for Oehlers’ model (a) and Smith and Teng’s model
(b)

96
ASSESSMENT OF PREDICTION MODELS

160 160
140 140
120 120
100 100

Vcal (kN)
Vcal (kN)

80 80
60 60
40 40
20 20
0 0
0 20 40 60 80 100 120 140 160 0 20 40 60 80 100 120 140 160
Vexp (kN) Vexp (kN)

(a) Jansze (1997) (b) Ahmed and Van Gemert (1999)


Figure 3.14 Validation results for Jansze’s model (a) and Ahmed and Van
Gemert’s model (b)

160
140
120
100
Vcal (kN)

80
60
40
20
0
0 20 40 60 80 100 120 140 160
Vexp (kN)

Figure 3.15 Validation results for Ziraba et al.’s model

Validation results of concrete tooth scenario based models are shown in Figure
3.16. Predictions using the lower limit by Raoof and Hassanen are presented only since
the upper limit was not critical for most cases. The average of Vexp/Vcal for this model is
1.26. A scattered distribution is also observed. Chaallal et al.’s model gives very
inaccurate predictions. The validation results for this model are also very scattered.

97
CHAPTER 3

160 160
140 140
120 120
100

Vcal (kN)
100
Vcal (kN)

80 80
60 60
40 40

20 20

0 0
0 20 40 60 80 100 120 140 160 0 20 40 60 80 100 120 140 160
Vexp (kN) Vexp (kN)

(a) Raoof and Hassanen (2000): lower bound (b) Chaallal et al. (1998)
Figure 3.16 Validation results for Raoof and Hassanen’s model – lower limit (a)
and Chaallal et al.’s model (b)

Graphical comparisons of the predictions by shear-lap scenario based models to


the experimental data are plotted in Figure 3.17. Arya and Farmer’s model uses Neubauer
and Rostasy’s method with an additional limit on the longitudinal shear stress between
the FRP and concrete. The validation results of Neubauer and Rostasy’s model are
relatively consistent with only 19 % variation. However, they are not conservative with
the average of Vexp/Vcal being 0.89. The additional limit recommended by Arya and
Farmer (2001) does improve the model’s conservativeness. The average of Vexp/Vcal
increases to 1.17.

98
ASSESSMENT OF PREDICTION MODELS

200 200

160 160

Vcal (kN)
Vcal (kN)
120 120

80 80

40 40

0 0
0 40 80 120 160 200 0 40 80 120 160 200
Vexp (kN) Vexp (kN)

(a) Neubauer and Rostasy (1997) (b) Arya and Farmer (2001)
Figure 3.17 Validation results for Neubauer and Rostasy’s model (a) and Neubauer
and Rostasy’s model with Arya and Farmer’s limit (b)

Table 3.6 summarises the validation results for the models discussed above. For
intermediate span debond, the simple models by limiting FRP strain level (Arya and
Farmer’s and Shehata et al.’s) tend to perform best despite the fact that they assume a
constant FRP tensile strain at peeling-off. Recent test results have demonstrated that the
FRP tensile strain at failure depends on a broad range of parameters, such as the
properties of FRP and concrete (fib Bulletin 14, 2001). These models can, however, be
used for a quick check. For end debond, most models tend to give scattered and/or
inaccurate results. Even though Nguyen et al.’s model has the average of Vexp/Vcal closest
to 1, the assumption of a constant FRP strain at failure for all cases has not been proved
by experiments.

99
CHAPTER 3

Table 3.6 Summary of validation results of debond models for beams


Coefficient
Average
of
Failure type Category Model of
variation
Vexp/Vcal
(%)
Flexural Arya and Farmer (2001) 1.08 15
crack Shehata et al. (2001) 1.15 13
debond Maruyama and Ueda (2001) 1.47 20
Intermediate
Teng et al. (2001) 1.61 23
span
debond Shear crack Blaschko et al. (1997) 1.39 23
debond Mohamed Ali (2000) 2.70 23
Flexure-
shear crack Niedermeier (2000) 1.59 39
debond
Shehata et al. (2001) 1.15 22
Nguyen et al. (2001) 1.02 25
Interfacial
Fanning and Kelly (2001) 1.64 24
localised
Saadatmanesh and Malek (1998) 3.59 26
failure
Tumialan et al. (1999) 1.77 35
El-Mihilmy and Tedesco (2001) 4.27 13
Oehlers (1992) 1.82 13
End
Shear Smith and Teng (2003) 1.84 13
debond
capacity Jansze (1997) 1.44 17
based Ahmed and Van Gemert (1999) 1.21 19
Ziraba et al. (1994) 1.59 16
Concrete Raoof and Hassanen (2000):
1.25 36
tooth Lower
scenario Chaallal et al. (1998) 8.27 13
Shear-lap Neubauer (1997) 0.89 19
scenario Arya and Farmer (2001) 1.17 24

100
ASSESSMENT OF PREDICTION MODELS

3.4 Validation of design guidelines


The validation studies of fib Bulletin 14 have been reported previously for the
models by Blascho, Janzse, Neubauer and Rostasy, Arya and Farmer, Niedermeier, and
Matthys.

The validation results of ACI 440 recommendations (published in 2001) are


shown in Figure 3.18. The applicability of Formulae 9-2 in ACI guideline was checked
against the beams failing by intermediate span debond. It is found that the formulae
provide accurate but relatively un-conservative predictions of the peak load. The average
of Vexp/Vcal is 0.92 with the coefficient of variation being 12 %. The beams failing by end
debond were checked using ACI detailing requirement No. 2, which states that the shear
force at the FRP end shall be less than two third of the concrete shear strength. As shown
in Figure 3.18b, this requirement provides a safe design limit. All predictions are lower
than the actual failure loads.

FSCD 160
350 FCD
SCD 140
300
120
250
Vcal (kN)

100
Vcal (kN)

200
80
150
60
100
40
50
20
0
0
0 50 100 150 200 250 300 350
Vexp (kN) 0 20 40 60 80 100 120 140 160
Vexp (kN)

(a) ACI formulae 9-2 (b) ACI detailing requirement No. 2


Figure 3.18 Validation of the methods recommended in ACI 440.

3.5 Summary of findings


This chapter has reported two main validation studies. In the first study, the
performance of nine debonding strength models was assessed. It was based on a database
consisting of 156 specimens. Based on this study, the following conclusions can be
drawn:

101
CHAPTER 3

• Among three empirical models, Maeda et al.’s model performs best.


• Most bond-slip based models provide good predictions of the bond strength. Among
these, Chen and Teng’s model tends to predict most accurately.

The second study involved an assessment of the strength models for three main
failure modes: flexural failures, intermediate span debond and end debond. 26 models
were verified using a database of 181 beams. The main conclusions drawn from this study
are:

• Beam theory is able to predict the full composite action of beams strengthened with
FRP composites.
• Intermediate span debond capacity is closely related to the CFRP tensile strain. Even
though the models based purely on limiting FRP strain yield good correlation with
current database, the effect of other important parameters such as FRP and concrete
properties need to be taken into account. The empirical formulae recommended by
ACI provide good correlation.
• End debond capacities are not predicted well by existing models. Most models fail to
recognise that the failure plane is in the concrete layer at tension reinforcement level.
The ACI guideline provides a safe design limit for this kind of failure.

102
EXPERIMENTAL INVESTIGATION OF SHEAR-LAP SPECIMENS UNDER TENSILE TESTING

CHAPTER 4 - EXPERIMENTAL INVESTIGATION OF


SHEAR-LAP SPECIMENS UNDER TENSILE TESTING

4.1 Introduction
The mechanical performance and durability of FRP composites in strengthening
applications depend primarily on their bond with the main structure. However, FRP
composites tend to delaminate or debond from concrete due to the local shear stress
concentration and the low shear strength of concrete. To be able to use FRP composites
for strengthening, an understanding of bonding and debonding mechanisms is therefore
essential.

This chapter reports an experimental program to study the characteristics of the


bond between CFRP and concrete formed by a wet lay-up method. The program includes
testing of 26 single shear-lap specimens under static loading. The parameters investigated
are the CFRP length, CFRP thickness, CFRP width and support block height.

4.2 Aim and design methodology


The aim of the experiment was to investigate the short-term behaviour of the bond
between CFRP bonded to concrete using a wet lay-up method under static loading. The
specific tasks were to identify the failure modes; to examine the failure mechanisms; to
determine the bond strength of concrete, the CFRP strains at failure and the effective
bond length; and to study the effect of several parameters on the bond performance. The
data from the experiments was also to be used for the assessment of theoretical and
numerical models of the bond between CFRP and concrete.

In order that the results of these shear-lap tests could be used in the study of the
beam tests reported later in Chapter 5, the concrete block width and CFRP cross sectional
dimensions were chosen to be the same as in those beams, and the same fibre, epoxy and
bonding procedures were selected. The blocks were cast together with the beams so that
the concrete strength would be similar. To ensure that the bond length was sufficient and
that the specimens could be easily fitted in the steel cage used to fix the blocks to the bed
of the loading machine, the concrete block length was chosen to be 300 mm. The gripping

103
CHAPTER 4

steel plate length was also chosen to be 250 mm in total, which was long enough such
that the stress was transferred to the CFRP relatively uniformly.

4.3 Experimental set-up


4.3.1 Specimen details
Thirteen test configurations were designed. For each configuration, two identical
specimens were manufactured, which made up a total of twenty six specimens. The
specimen dimensions and the test set-up are shown in Figure 4.1. The variables are listed
in Table 4.1. They were the bond length, Lf, the number of CFRP plies, the support block
clearance, H, and the CFRP width, bf. The configurations were initially labelled as B1 to
B16. For clarity, they were relabelled as T1 to T13. Two identical specimens for each
configuration were labelled as ‘a’ and ‘b’ (for example, configuration T1 has two
specimens: T1a and T1b).

Steel cage H

CFRP

Lf

300

bf

140

140

All dimensions are in mm.


Figure 4.1 Specimen dimensions and test set-up

104
EXPERIMENTAL INVESTIGATION OF SHEAR-LAP SPECIMENS UNDER TENSILE TESTING

Table 4.1 Variables in the experimental program


No. of
Design Lf (mm) H (mm) bf (mm) Label Initial label
CFRP plies
1 60 2 110 100 T1 B6
2 80 2 110 100 T2 B7
3 100 2 110 100 T3 B1
4 140 2 110 100 T4 B8
5 180 2 110 100 T5 B9
6 220 2 110 100 T6 B10
7 100 2 5 100 T7 B2
8 100 2 70 100 T8 B4
9 100 6 110 100 T9 B11
10 140 6 110 100 T10 B15
11 180 6 110 100 T11 B16
12 100 2 110 50 T12 B13
13 100 2 110 70 T13 B14

4.3.2 Materials
4.3.2.1 Concrete
The concrete material used was supplied separately, pre-mixed from a local
supplier. Compressive strength tests were carried out on concrete cylinders for each batch
in accordance with AS1012.9 (Standards Australia, 1999). The concrete strength after 67
days was 53.7 MPa. Further information on the concrete strength development with time
is presented in Appendix B.1.

4.3.2.2 MBrace Primer, MBrace CF 130 Fibre and MBrace Saturant


MBrace FRP ‘Wet lay up’ system was used. The system was provided by MBT
(Australia) Pty Limited. It consists of three main components: MBrace Primer, MBrace
CF 130 Fibre and MBrace Saturant. Their mechanical properties are listed in Table 4.2.
The use of MBrace Primer was to improve the bonding of the composite to the substrate.
It is a two-part epoxy product with low viscosity and 100 % solids content. It has the
ability to penetrate the substrates and to bond to a dry concrete surface. MBrace CF 130

105
CHAPTER 4

Fibres, which are also known as S&P C-sheet 240, are high tensile strength carbon fibres
supplied in unidirectional tow sheets of 300 mm width. The fibres were held together
with a backing grid and rolled together with a plastic sheet. The nominal thickness, based
on the total thickness of the fibres in a unit width, was reported to be 0.176 mm. The fibre
weight was 300 g/m2. MBrace Saturant is a two-part epoxy with 100 % solids content
used to impregnate the fibres to form a composite and to provide bonding to the primed
surface. Its compressive strength was reported to be greater than 80 MPa. The pot life of
the saturant was 30 minutes at 25°C and it reached the design strength after 7 days.

Table 4.2 Mechanical properties of MBrace FRP system as given by the


manufacturer
Elastic modulus Tensile strength
Material (MPa) (MPa)
direct flexural direct flexural
MBrace CF 130 fibre 240000 N/A 3900 N/A
MBrace Saturant >3000 >3500 > 50 >120
MBrace Primer >700 >580 > 12 >24

Figure 4.2 MBrace FRP system components and main tools used for installation

106
EXPERIMENTAL INVESTIGATION OF SHEAR-LAP SPECIMENS UNDER TENSILE TESTING

As the elastic modulus of the composite material was of primary importance, it


was determined using tensile testing accordance to ASTM D 3039/ D 3039M – 00
(ASTM International, 2000). The test results are listed in Table 4.3. These results were
based on the normal thickness of the fibres (0.176 mm). A more detailed description of
the test is presented in Appendix B.3. The average elastic modulus found was slightly
lower than the reported value. The ultimate strengths are not reported here since some
tensile specimens failed due to the stress concentration near the grip.

Table 4.3 Measured elastic modulus of the composite


Specimen label Tensile elastic modulus (MPa)
CT1 202000
CT2 215000
CT3 212000
CT4 194000
CT5 220000
Average 209000

4.3.3 Specimen preparation and instrumentation


The process of applying CFRP to concrete involved two main steps: surface
preparation and bonding. The concrete surface was prepared using high-pressure water
jetting to remove a thin layer of the paste and expose the coarse aggregates. The water jet
operated at 4000 psi or 28 MPa. This method proved to be efficient at providing a
relatively rough surface to improve mechanical bonding. The surface was left to dry and
all loose particles and dust was removed with an industrial vacuum cleaner. MBrace
Primer was then applied on the surface thoroughly with a brush. The primer, which had
low viscosity, was allowed to fully penetrate and produce a nonporous shiny film.

The bonding operation was carried out about 30 minutes to 1 hour after the
application of the primer. The operation included resin under-coating, carbon fibres
placement and resin over-coating. MBrace Saturant was first applied on the primed
surface using a steel scraper to form a coat approximately 2 mm thick, which proved to be
sufficient to achieve a wet-out of the FRP fabrics. Then, a piece of MBrace CF 130 fibre
sheet, which had been cut beforehand into prescribed sizes using scissors, was placed

107
CHAPTER 4

with the fibre side down onto the coating and generally smoothed down by hand. After
that, the surface of the sheet was rolled over along the longitudinal direction of the fibres
using a ribbed roller to impregnate resin into the fibres and remove any air bubbles.
Rolling was continued until the resin was squeezed out between the fibres. A thin layer of
resin was then placed on the fibres and pressed down with the scraper to smooth out any
remaining imperfections. To bond another ply of fibres, the same steps were followed. To
allow for epoxy impregnation, a period of five minutes was allowed between subsequent
applications. The specimens were left to cure at room temperature for at least seven days
before testing.

The specimens were instrumented to record strain, load and extension readings.
Out of two identical specimens manufactured for each configuration, one block (block
‘a’) had more instrumentation than the other (block ‘b’). Strain gauges were used to
measure the strain levels in the composite along its centre line (Figure 4.3). Loads and
extensions were measured using a load cell and a linear variable displacement transducer
(LVDT) in the loading machine. LVDTs were also mounted on the sides of the block in
an attempt to measure the slip of the CFRP. However, it was found that their readings
were not reliable since the slips were too small to be picked up by these LVTDs.

108
EXPERIMENTAL INVESTIGATION OF SHEAR-LAP SPECIMENS UNDER TENSILE TESTING

15
15
15
2@18

3@18

4@18
Lf = 60 mm Lf = 80 mm Lf = 100 mm
6@18 15

15

15
8@18

10@18
Lf = 140 mm Lf = 180 mm Lf = 220 mm

All dimensions are in mm


Figure 4.3 Locations of strain gauges on CFRP shear-lap specimens

The data acquisition equipment made by dataTaker, ‘DT505’, was used in the
experimental program with a channel expander module ‘CEM-AD’. During testing,
readings were recorded once every second using the computer programs ‘Defriend’ and
‘Detransfer’. Crack propagation was captured by a digital video recorder and a digital
high-speed video recorder. The high-speed video recorded at 500 frames per second and
it was used during two tests only.

4.3.4 Testing set-up and procedure


The specimens were loaded using the Baldwin universal testing machine, which
has a 500 kN load capacity and a piston stroke of 30 mm. The loading set-up and testing
machine are shown in Figure 4.4. The cage was designed to fix the block vertically to the
bottom bed of the machine. For gripping, two steel plates were bonded to the free end of
the CFRP composite using MBrace Saturant. To ensure a good bond between the steel
plates and CFRP, the plate surfaces were roughened using a steel brush, cleaned using
acetone and let dry. The bonding process of steel on CFRP was carried out together with

109
CHAPTER 4

the bonding of CFRP on concrete. The plates were connected to the jaw through a
connection rod. The plate thickness was chosen to be 3 mm to ensure that yielding would
not occur near the connection hole. During installation of the specimen, good alignment
was controlled using a plumb-line. The tension bolts were tightened slightly before the
test.

Steel jaw Top


bed

Connection
Tension rod
bolts CFRP limit
Support
block

Stiffener

Steel Bottom
clamp bed

Load cell and LVDT

(a) (b)
Figure 4.4 Loading and gripping method (a) and a specimen before testing (b)

Loading was applied in displacement control, achieved using a LVDT attached to


the actuator that provided constant feedback to a computer program which controlled the
hydraulic distributor system. Based on the readings of the LVDT, the computer program
constantly adjusted the pressure output to the actuator in order to maintain the required
displacement rate. The specimens were loaded at a rate of 0.20 mm per minute. After
each test, the concrete block was taken out and checked for the crack patterns. Some of
the blocks with minimum damage were reused by bonding CFRP on the other undamaged
side.

110
EXPERIMENTAL INVESTIGATION OF SHEAR-LAP SPECIMENS UNDER TENSILE TESTING

4.4 Experimental results


4.4.1 Failure modes and ultimate capacity
Two failure modes observed were interfacial debond and shear-tension failure
(Figure 4.5). The specimens bonded with two plies of CFRP with the bond length greater
or equal 80 mm failed by interfacial debond. Concrete just below the bond line ruptured
in shear with a thin layer of concrete substrate still bonded to the laminate for most of the
area (Figure 4.6a). The failure was very brittle especially for the specimens with a short
bond length. Close examination of the failure surface revealed the failure mechanism.
During bonding, epoxy infiltrated into the rough surface of concrete. The interface shear
strength was mainly provided by the interlocking between the epoxy and the concrete
surface irregularities. When the shear stress increased to a high value, those irregularities
started to rupture with cracks propagating both through the cement paste and the
aggregate (Figure 4.6b). Since there were many irregularities on the surface, the cracking
propagated gradually starting from the loading end and transferring the stress to the next
bond area. In most specimens failing in this mode, a triangular concrete prism was also
pulled off near the loaded end.

The specimens bonded with two plies of CFRP with a very short bond length (60
mm) or with six plies of CFRP of 100 mm bond length failed by shear-tension failure. A
crack initiated from the unloaded CFRP end, propagated deep into the concrete block and
finally terminated near the support block location. The failure was very brittle. The crack
propagated through both the cement paste and coarse aggregate (Figure 4.7).

Some specimens with two plies of CFRP bonded over a very short length or with
six plies of CFRP bonded at a long bond length failed either due to shear-tension failure
or interfacial debond. One reason for the inconsistency was that, in those cases, the stress
transferred well to the unloaded CFRP end, which resulted in a high stress concentration
at both ends of the composite and therefore both failure modes were likely. Another
possible reason for the inconsistency was the difficulty in keeping the CFRP perfectly
aligned with the concrete block. This led to bending of the composites and therefore a
different stress distribution. This could have been exacerbated under high loads due to the
rotation of the concrete block.

111
CHAPTER 4

The observed failure modes and recorded ultimate capacities are summarised in
Table 4.4. The failure loads varied from 18.8 kN to 42.8 kN with an average of 28 kN.
The averages of the average bond stress and the maximum FRP stress were 2.7 MPa and
704.5 MPa, respectively. The table also shows that the model by Chen and Teng (2001)
predicts the bond strengths of the specimens relatively well. The average of the ratios of
the experimental failure load to the prediction is 0.98 with the coefficient of variation
being 13 %.

The photos of all specimens after failure are shown in Appendix C.1. The high-
speed video captures for a specimen are also illustrated in Appendix C.2.

T3: mode 1 T1: mode 2


Figure 4.5 Failure modes

112
EXPERIMENTAL INVESTIGATION OF SHEAR-LAP SPECIMENS UNDER TENSILE TESTING

Carbon fibre

Coarse
Transverse crack aggregate
initiation location Cement paste

(a) A CFRP bond surface after failure

Surface cracks

Force direction

(b) A close-up on a concrete block surface after failure


Figure 4.6 Typical crack surfaces in interfacial debond failure

Force direction
Coarse
aggregate

Cement paste Crack initiation


location

Figure 4.7 A typical crack surface in shear-tension failure

113
CHAPTER 4

Table 4.4 Shear-lap test results

Max CFRP Prediction


Max. Max average by the model
1
tensile
Specimen Variable load bond stress Failure mode by Chen and
stress
(kN) (MPa) Teng (2001)
(MPa)

T1a 20.0 3.3 568 shear-tension 22.2


T1b 18.8 3.1 534 interfacial debond 22.2
T2a 25.8 3.2 733 interfacial debond 26.0
T2b 25.2 3.2 716 interfacial debond 26.0
T3a 25.8 2.6 733 interfacial debond 27.3
T3b 27.3 2.7 776 interfacial debond 27.3
Lf
T4a 26.7 1.9 759 interfacial debond 27.3
T4b 25.9 1.9 736 interfacial debond 27.3
T5a 27.8 1.5 790 interfacial debond 27.3
T5b 31.7 1.8 901 interfacial debond 27.3
T6a 31.7 1.4 901 interfacial debond 27.3
T6b 28.6 1.3 813 interfacial debond 27.3
T7a 33.0 3.3 938 interfacial debond 27.3
T7b 26.9 2.7 764 interfacial debond 27.3
H
T8a 28.5 2.9 810 interfacial debond 27.3
T8b 29.8 3.0 847 interfacial debond 27.3
T9a 28.4 2.8 269 shear-tension 37.4
T9b 29.8 3.0 282 shear-tension 37.4
T10a 37.4 2.7 354 interfacial debond 45.3
tf
T10b 33.3 2.4 315 shear-tension 45.3
T11a 42.8 2.4 405 interfacial debond 47.0
T11b 39.0 2.2 369 shear-tension 47.0
T12a 18.8 3.8 1068 interfacial debond 17.4
T12b 19.3 3.9 1097 interfacial debond 17.4
bf
T13a 21.2 3.0 860 interfacial debond 22.1
T13b 24.2 3.5 982 interfacial debond 22.1
Note: 1) Max average bond stress = Max load / Total bond area

4.4.2 Strain, stress and slip distributions


A cut through the composite revealed that it had two main layer types (Figure
4.8a). The saturant layers contained mainly MBrace saturant. The fibre/saturant layers
contained carbon fibres and MBrace saturant. These layers were formed as a result of the
effort to impregnate the resin into the fibres. The saturant layers were generally thinner

114
EXPERIMENTAL INVESTIGATION OF SHEAR-LAP SPECIMENS UNDER TENSILE TESTING

than the fibre/saturant layers because of the pressure applied during bonding. The
approximate thicknesses of the layers are illustrated in Figure 4.8b.

Fibre/saturant Top saturant


Saturant finish layer Strain gauge location
Saturant
0.7
Fibre/
saturant 0.7

1.0

Concrete

Concrete
All dimensions are in mm
(a) A cut through a 6-ply composite (dark (b) Layers of a 2-ply CFRP composite bonded on
layers contain mainly carbon fibres) concrete

Figure 4.8 Layers of the composite

In the experiments, the CFRP strain was measured near the composite top surface
(Figure 4.8b). The top strain was expected to be different to the average composite strain
as the strain could vary across the thickness of the composite. This variation was very
difficult to measure experimentally. However, as shown by the finite element analysis
presented in Chapter 6, the difference between the top strain and the average value was
only significant near the ends of the CFRP and near the tips of widely open transverse
cracks. Hence, to describe the global behaviour of the bond, it was reasonable to assume
that the strain measured can represent the average CFRP strain.

The strain distributions measured in the specimens with varying bond lengths are
plotted in Figure 4.9 to Figure 4.14. At a low load level, the distributions show a gradual
decline from the peak near the loaded edge to the other end. As the load increases but still
at a relative low level (less than 10 kN), the strain at the first gauge increases most
rapidly. As a result, the distributions have the steepest slope near the loaded edge
(between the first and second gauges from the loaded edge). As the load increases further
to around 15 kN, the strain at the second gauge starts to increase faster than that at the
first gauge. Consequently, the steepest slope moves to the right of the second gauge.

115
CHAPTER 4

With addition of more loads, the steepest portion shifts further to the right. The
distributions start to flatten out near the loaded edge indicating delamination underneath
the CFRP there. This is observed most clearly in the specimens with long bond lengths
(T4, T5 and T6). The effective bond length can be approximated from the strain
distributions of those specimens. It is the distance over which the maximum strain
decreases to near zero. From the plots, the effective bond length is found to be
approximately 100 mm. The maximum CFRP strain is approximately 3000 microstrain.

2500 Load (kN)


5.0
2000 10.0
15.0
Microstrain

1500 18.0
20.0
1000

500

0
0 50 100 150 200
Distance from loaded edge (mm)

Figure 4.9 Strain distributions along bond length in specimen T1a (Lf = 60 mm)

3000 Load (kN)


5.0
2500 10.0
15.0
2000
Microstrain

20.0
1500 22.0
24.0
1000 25.8

500

0
0 50 100 150 200
Distance from loaded edge (mm)

Figure 4.10 Strain distributions along bond length in specimen T2a (Lf = 80 mm)

116
EXPERIMENTAL INVESTIGATION OF SHEAR-LAP SPECIMENS UNDER TENSILE TESTING

3500 Load (kN)


5.0
3000
10.0
2500 15.0

Microstrain
2000 20.0
22.0
1500 25.7
1000
500
0
0 50 100 150 200
Distance from loaded edge (mm)

Figure 4.11 Strain distributions along bond length in specimen T3a (Lf = 100 mm)

3500 Load (kN)


5.0
3000
10.0
2500 15.0
Microstrain

2000 20.0
22.0
1500
24.0
1000 26.0
500 26.7

0
0 50 100 150 200
Distance from loaded edge (mm)

Figure 4.12 Strain distributions along bond length in specimen T4a (Lf = 140 mm)

5000 Load (kN)


4500 10.0
4000 15.0
3500 20.0
Microstrain

3000 22.0
2500 24.0
2000 26.0
1500 27.0
1000
27.8
500
0
0 50 100 150 200
Distance from loaded edge (mm)

Figure 4.13 Strain distributions along bond length in specimen T5a (Lf = 180 mm)

117
CHAPTER 4

4500 Load (kN)


5.0
4000 10.0
3500 15.0
Microstrain 3000 20.0
22.0
2500 24.0
2000 26.0
1500 27.0
28.0
1000 28.5
500 26.9
0
0 50 100 150 200
Distance from loaded edge (mm)

Figure 4.14 Strain distributions along bond length in specimen T6a (Lf = 220 mm)

The strain distributions measured in the specimens with varying support block
heights are plotted in Figure 4.15 and Figure 4.16. There is little difference in the
distributions.

4500 Load (kN)


4000 5.0
3500 10.0
3000 15.0
Microstrain

2500 20.0
2000 25.0
30.0
1500
32.0
1000
33.0
500
0
0 50 100 150 200
Distance from loaded edge (mm)

Note: Only three gauges were installed on this specimen.

Figure 4.15 Strain distributions along bond length in specimen T7a (H = 5 mm)

118
EXPERIMENTAL INVESTIGATION OF SHEAR-LAP SPECIMENS UNDER TENSILE TESTING

4000 Load (kN)

3500 5.0
10.0
3000 15.0

Microstrain
2500 20.0
2000 25.0
26.0
1500 27.0
1000 28.0
500 28.5

0
0 50 100 150 200
Distance from loaded edge (mm)

Figure 4.16 Strain distributions along bond length in specimen T8a (H = 70 mm)

The strain distributions measured in the specimens with varying CFRP widths are
plotted in Figure 4.17 and Figure 4.18. It can be seen that the maximum composite strain
increases as the CFRP width decreases. The maximum CFRP strains increase to
approximately 3500 and 4500 microstrain for the specimens with the CFRP width of
70 mm and 50 mm, respectively.

5000 Load (kN)


4500 5.0
4000 10.0
3500
14.0
Microstrain

3000
15.0
2500
2000 16.0
1500 17.0
1000 18.0
500
0
0 50 100 150 200
Distance from loaded edge (mm)

Figure 4.17 Strain distributions along bond length in specimen T12a (bf = 50 mm)

119
CHAPTER 4

4000 Load (kN)


5.0
3500
10.0
3000
Microstrain 15.0
2500 18.0
2000 19.0
1500 20.0
1000 20.5
500 21.0
0
0 50 100 150 200
Distance from loaded edge (mm)

Figure 4.18 Strain distributions along bond length in specimen T13a (bf = 70 mm)

For the specimens bonded with six plies of CFRP, i.e. specimens T9, T10 and
T11, the maximum applied loads were higher and the measured strains were much lower
than those of the specimens bonded with two plies. Therefore, the measurements on these
specimens were much more sensitive to the alignment errors. The errors were difficult to
be avoided and consequently led to bending of the composite plate during loading. This
especially affected the readings of the gauges near the loaded edge. As a result, the strain
measurements for these specimens were not reliable near the edge. As illustrated in
Figure 4.19 to Figure 4.21, the gauges closest to the loaded edge measured low and even
negative strains indicating that the composite was bended there.

500

0
0 50 100 150 200
Microstrain

-500
Load (kN)
5.1
-1000 10.0
15.0
-1500 20.0
25.0
29.8
-2000
Distance from loaded edge (mm)

Figure 4.19 Strain distributions along bond length in specimen T9a (Lf = 100 mm)

120
EXPERIMENTAL INVESTIGATION OF SHEAR-LAP SPECIMENS UNDER TENSILE TESTING

600 Load (kN)


5.0
500 10.0
400 15.0

Microstrain
20.0
300 25.0
200 30.0
35.0
100 36.0
0
0 50 100 150 200
Distance from loaded edge (mm)

Figure 4.20 Strain distributions along bond length in specimen T10a (Lf = 140 mm)

1400 Load (kN)


1200 10.0
1000 20.0
800 30.0
Microstrain

600 35.0
400 40.0
200 42.8
0
-200 0 50 100 150 200
-400
Distance from loaded edge (mm)

Note: Only five gauges were installed on this specimen.

Figure 4.21 Strain distributions along bond length in specimen T11a (Lf = 180 mm)

The average bond stress between two adjacent strain gauges can be calculated by
dividing the force difference by the bond area and given by the following equation:
E f (ε f ,i +1 − ε f ,i ) t f
τ= (4.1)
∆L
where Ef and tf are the CFRP elastic modulus and thickness, respectively; εf,i+1 and εf,i are
the CFRP strains; and ∆L is the distance between strain gauges. The bond stress
distributions at discrete data points for three specimens with relatively long bond lengths
are plotted in Figure 4.22 to Figure 4.24. The distributions of other specimens are
included in Appendix C.3.

The area under the bond stress curve is proportional to the amount of tension
transferred by the bond. It can be observed from these plots that the tension is transferred

121
CHAPTER 4

via a certain length only, called the ‘active length’. When the shear stress at the loaded
end exceeded the shear capacity, concrete cracking occurred and the ‘active length’
displaces toward the unloaded end of the specimen.

5 Load (kN)
5.0
4 10.0
Bond stress (MPa)

15.0
3 20.0
22.0
24.0
2
26.0
26.7
1

0
0 50 100 150 200
Distance from loaded edge (mm)

Figure 4.22 Average bond stress distributions along bond length in specimen T4a

8 Load (kN)
5.0
7
10.0
6
Bond stress (MPa)

15.0
5 20.0
4 22.0
24.0
3 26.0
2 27.0
1 27.8
0
0 50 100 150 200
Distance from loaded edge (mm)

Figure 4.23 Average bond stress distributions along bond length in specimen T5a

122
EXPERIMENTAL INVESTIGATION OF SHEAR-LAP SPECIMENS UNDER TENSILE TESTING

8 Load (kN)

7 5.0
10.0
6 15.0

Bond stress (MPa)


20.0
5
22.0
4 24.0
26.0
3 27.0
2 28.0
28.5
1 26.9
0
0 50 100 150 200
Distance from loaded edge (mm)

Figure 4.24 Average bond stress distributions along bond length in specimen T6a

The average slip can be calculated as the incremental sum of the CFRP extension,
as follows
ε f ,i +1 + ε f ,i
si = ∆L + s i −1 (4.2)
2
In Equation 4.2, since the concrete block is much stiffer than the CFRP, concrete
elongation has been ignored.

The plots of the average bond stress versus the average slip at several load levels
for some specimens are presented below. The distributions for the other specimens are
shown in Appendix C.4. In these plots, the average bond stresses and slips are calculated
at several locations along the bond length. For example, the bond stress and slip at the
location 24 mm from the loaded edge are calculated using the readings from the strain
gauges at 15 and 33 mm from the loaded edge.

The plots for the specimens bonded with the CFRP composite of 100 mm width
bonded over a relatively long bond length (specimens T4a, T5a, and T6a) are shown in
Figure 4.25 to Figure 4.27. The curves vary significantly, possibly due to the presence of
transverse cracks along the joint. However, it can be seen that the peak shear stress was
around 4 MPa at a slip of approximately 0.05 mm. Similar observation can also be made
from the bond stress slip distributions at several load levels for the specimens with the
support block height equal to 70 mm (Figure 4.28). However, all of these curves appear
to have a nonlinear ascending and descending trends. It was found that these trends can

123
CHAPTER 4

be approximately described using Popovics’ equation (Equation 2.1) with a = 3 as shown


in Figure 4.29.

The maximum bond stresses of the local bond-slip curves seem to depend
significantly on the composite width. The peak bond stresses are approximately 6 and 5
MPa for the specimens with the composite widths of 50 and 70 mm (specimens T12a and
T13a), respectively.

7 Location (mm) 6 Load (kN)


24 42
6 5.0 10.0
5
60 78 15.0 20.0
Bond stress (MPa)

Bond stress (MPa)


5
96 4 22.0 24.0
4 25.0 26.0
3 26.7
3
2
2
1 1

0 0
0.00 0.05 0.10 0.15 0.20 0.25 0.00 0.05 0.10 0.15 0.20 0.25
Slip (mm) Slip (mm)

(a) (b)
Figure 4.25 Bond stress versus slip development in specimen T4a (bf = 100 mm)

7 Location (mm) 8 Load (kN)


24 42
6 7 5.0 10.0
60 78
Bond stress (MPa)

15.0 20.0
Bond stress (MPa)

5 6
96 22.0 24.0
4 5
26.0 27.0
4 27.8
3
3
2
2
1 1
0 0
0.00 0.05 0.10 0.15 0.20 0.25 0.00 0.05 0.10 0.15 0.20 0.25
Slip (mm) Slip (mm)

(a) (b)
Figure 4.26 Bond stress versus slip development in specimen T5a (bf = 100 mm)

124
EXPERIMENTAL INVESTIGATION OF SHEAR-LAP SPECIMENS UNDER TENSILE TESTING

8 Location (mm) 8 Load (kN)


7 24 42 7 5.0 10.0
15.0 20.0
6 60 78
Bond stress (MPa)
6

Bond stress (MPa)


22.0 24.0
5 96 114 5 26.0 27.0
28.0 28.5
4 4
3 3
2 2
1 1
0 0
0.00 0.05 0.10 0.15 0.20 0.25 0.00 0.05 0.10 0.15 0.20 0.25
Slip (mm) Slip (mm)

(a) (b)
Figure 4.27 Bond stress versus slip development in specimen T6a (bf = 100 mm)

8 Location (mm) 8 Load (kN)


7 24 42 7 5.0 10.0
15.0 20.0
Bond stress (MPa)
Bond stress (MPa)

6 6
60 25.0 26.0
5 5 27.0 28.0
4 4 28.5
3 3
2 2
1 1
0 0
0.00 0.05 0.10 0.15 0.20 0.25 0.00 0.05 0.10 0.15 0.20 0.25
Slip (mm) Slip (mm)

(a) (b)
Figure 4.28 Bond stress versus slip development in specimen T8a (bf = 100 mm)

8
7
6
Bond stress (MPa)

Popovics'
5
equation
4
3
2
1
0
0.00 0.05 0.10 0.15 0.20 0.25
Slip (mm)

Figure 4.29 Fitted bond-slip relationships for specimens with two layers of CFRP of
100 mm width

125
CHAPTER 4

8 Location (mm) 8 Load (kN)


5.0 10.0
7 24 42 7
14.0 15.0

Bond stress (MPa)


6
Bond stress (MPa)

6
60 78 16.0 17.0
5 5
18.0
4 4
3 3
2 2
1 1
0 0
0.00 0.05 0.10 0.15 0.20 0.25 0.00 0.05 0.10 0.15 0.20 0.25
Slip (mm) Slip (mm)

(a) (b)

Figure 4.30 Bond stress versus slip development in specimen T12a (bf = 50 mm)

8 Location (mm) 8 Load (kN)

7 24 42 7 5.0 10.0
15.0 18.0
Bond stress (MPa)

6
Bond stress (MPa)

6
60 78 19.0 20.0
5 5
20.5 21.0
4 4
3 3
2 2
1 1
0 0
0.00 0.05 0.10 0.15 0.20 0.25 0.00 0.05 0.10 0.15 0.20 0.25
Slip (mm) Slip (mm)

(a) (b)
Figure 4.31 Bond stress versus slip development in specimen T13a (bf = 70 mm)

4.4.3 Load-slip curves and parametric study


The load-slip curves for the specimens with varying bond length, Lf, i.e.
specimens T1a to T6a, T9a to T11a, are plotted in Figure 4.32 and Figure 4.33. The slip
was the average slip between the second and third gauges from the loaded edge. This
location was chosen because the slip measurements there were not significantly affected
by the local stress variation near the edge and they were still of reasonable high readings.

126
EXPERIMENTAL INVESTIGATION OF SHEAR-LAP SPECIMENS UNDER TENSILE TESTING

It can be observed that the initial stiffness is very similar for the specimens with
two plies of CFRP. For the specimens with six plies, the stiffness varies due to the
alignment errors but the general trend is similar. As expected, the specimens with longer
bond lengths show more ductile behaviour.

35
30
Total load (kN)

25
T1a - Exp.
20 T2a - Exp.
15 T3a - Exp.
T4a - Exp.
10 T5a - Exp.
5 T6a - Exp.
0
0.00 0.10 0.20 0.30 0.40 0.50
Slip at 42 mm from loaded edge (mm)

Figure 4.32 Load versus slip curves for specimens with two layers of CFRP and
variable bond lengths

45
40
35
Total load (kN)

30
25
20 T9a - Exp.
15 T10a - Exp.
10 T11a - Exp.
5
0
0.00 0.05 0.10 0.15
Slip at 42 mm from loaded edge (mm)

Figure 4.33 Load versus slip curves for specimens with six layers of CFRP and
variable bond lengths (subjected to alignment errors)

The effect of the bond length, Lf, on the ultimate load capacity and average
maximum bond stress is illustrated in Figure 4.34. The average bond stress was calculated
by simply dividing the total load by the total bond area. It is clear that for the specimens
with two plies of carbon fibres, the ultimate load does not increase significantly once the
bond length exceeds 100 mm. The average bond stress therefore decreases greatly as Lf
increases beyond that limit. A slight increase in the ultimate load is still observed for the

127
CHAPTER 4

specimens with a very long bond length possibly due to some friction after debonding.
When the failure is dominantly shear-tension (specimens T9a to T11a), an increase in the
bond length results in a significant increase in the bond capacity.

Experiment 2 plies Experiment 6 plies Experiment 2 plies Experiment 6 plies

Average bond stress (MPa)


50 3.5
3.0
40
2.5
Load (kN)

30 2.0
20 1.5
1.0
10 0.5
0 0.0
0 50 100 150 200 250 0 50 100 150 200 250
Lf (mm) Lf (mm)

(a) (b)
Figure 4.34 Effect of bond length on bond capacity

A thicker composite can allow better stress transfer along the bond area. However,
the improvement usually does not outweigh the waste in CFRP materials. Figure 4.36b
demonstrates this point showing that the maximum CFRP stress is reduced significantly
as the nominal fibre thickness increases from 0.352 to 1.056 mm. The specimens with a
thicker composite also have much stiffer behaviour.

30
25
Total load (kN)

20
15 T3a - Exp.
10 T9a - Exp.

5
0
0.00 0.05 0.10 0.15 0.20
Slip at 42 mm from loaded edge (mm)

Figure 4.35 Load-slip curves for specimens with different CFRP thicknesses T3a (2
plies) and T9a (6 plies)

128
EXPERIMENTAL INVESTIGATION OF SHEAR-LAP SPECIMENS UNDER TENSILE TESTING

Experiment L =100 mm Experiment L =100 mm


Experiment L=140 mm Experiment L=140 mm
Experiment L=180 mm Experiment L=180 mm
45 1020

CFRP stress (MPa)


40 820
Load (kN)

35 620
30 420
25 220
20 20
0 0.5 tf (mm) 1 1.5 0 0.5 1 1.5
tf (mm)

(a) (b)
Figure 4.36 Effect of CFRP thickness on bond capacity

Figure 4.37 shows that the composite width has a significant effect on the ultimate
capacity. The specimen with a narrower CFRP width has a lower bond capacity.
However, if the average bond stress is considered, the narrower strip proves to be more
effective (Figure 4.38b).

30
25
Total load (kN)

20
15
T3a - Exp.
10 T12a - Exp.
T13a - Exp.
5
0
0.00 0.05 0.10 0.15 0.20
Slip at 42 mm from loading edge (mm)

Figure 4.37 Load-slip curves for specimens with different CFRP widths T3a (100
mm), T12a (50 mm) and T13a (75 mm)

129
CHAPTER 4

30 5

Average bond stress (MPa)


25
4
20
Load (kN)

3
15
2
10
1
5
0 0
20 40 60 80 100 120 20 40 60 80 100 120
b f (mm) b f (mm)

(a) (b)
Figure 4.38 Effect of CFRP width on bond capacity

Figure 4.39 and Figure 4.40 indicate that the support block clearance has an
insignificant effect on the bond except when the clearance is very small. Specimen T3a
with only 5 mm clearance failed at a slightly higher load. However, the variation between
two identical specimens T3a and T3b is also significant (20 %). Therefore, it is not
possible to draw conclusions about the effect of a small clearance on the bond capacity.

35
30
Total load (kN)

25
20
15 T3a - Exp.
10 T7a - Exp.
T8a - Exp.
5
0
0.00 0.05 0.10 0.15 0.20
Slip at 42 mm from loading edge (mm)

Figure 4.39 Load-slip curves for specimens with different support clearances T3a
(110 mm), T7a (5 mm) and T8a (70 mm)

130
EXPERIMENTAL INVESTIGATION OF SHEAR-LAP SPECIMENS UNDER TENSILE TESTING

35
30
25

Load (kN)
20
15
10
5
0
0 25 50 75 100 125
H (mm)

Figure 4.40 Effect of varying support clearance on bond capacity

4.5 Summary of findings


This chapter has presented the procedures and results from a testing program
consisting of 26 CFRP-concrete shear-lap specimens. The investigated variables were the
CFRP bond length, the CFRP thickness, the CFRP width and the height of the support
block. The findings from this work are:

• Two types of failure were observed. Interfacial debond was dominant in the
specimens with two plies of CFRP bonded over a sufficient length. Shear-tension
failure was likely for the ones with six plies of CFRP or with two plies of CFRP but
bonded over a very short length. The failure surface was in the concrete through both
the cement matrix and aggregates.
• The failure loads varied from 18.8 kN to 42.8 kN with an average of 28.0 kN. The
means of the average bond stress and the maximum FRP stress were 2.7 MPa and
704.5 MPa, respectively.
• A typical CFRP strain distribution showed an exponential curve at early stages of
loading with the maximum value located near the loaded edge. The curve gradually
flattened as the load approached the ultimate capacity.
• The strain distributions also indicated an effective stress transfer length of around 100
mm for specimens with 2 plies of CFRP of 100 mm width. The local bond-slip curves
were nonlinear and varied significantly possibly due to the presence of inclined cracks
intersecting in the bond. The average maximum bond stress was around 4 MPa at an
average slip of approximately 0.05 mm for the specimens with 2 plies of CFRP of 100

131
CHAPTER 4

mm width. Popovics’ equation could be used to describe the relationship


approximately.
• The bond length and the CFRP width had a significant effect on the bond behaviour.
The specimens with a longer bond length had a more ductile failure showing a higher
slip at ultimate. The maximum bond stresses were higher in the specimens with a
narrower CFRP composite.

132
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

CHAPTER 5 - EXPERIMENTAL INVESTIGATION OF


RETROFITTED BEAMS UNDER BENDING

5.1 Introduction
There have been a number of studies on debonding failures in RC beams
strengthened using FRP composites as reported in Chapter 2. However, due to the
complexity of the failures, the mechanisms are still not fully understood and the
influences of several parameters are still not clear.

In practice, RC beams are often retrofitted with longitudinal FRP composites


together with some form of anchorage. The anchorage is to avoid debonding of the
longitudinal composite and/or to improve the shear capacity of the beams. However,
experimental investigations on the performance of anchorage methods are very limited.
The effectiveness of anchorage methods in preventing different debonding failures has
not been clearly demonstrated.

Therefore, in this study, two experimental programs are carried out to further
investigate the behaviour of RC beams retrofitted with CFRP fabrics using a wet lay-up
method. The aim of the first program is to study the failure mechanisms and the influence
of a number of parameters. In this program, a total of eighteen RC beams are constructed
and tested under four-point bending. Two are control beams and sixteen are retrofitted
with CFRP fabrics. The variables are the CFRP bond length, CFRP thickness, steel
tension reinforcement amount, concrete cover and stirrup spacing. The aim of the second
experimental program is to investigate the effectiveness of anchorage systems using
composite U-straps. In the program, the beams are tested under three-point bending. A
total of eight tests are carried out. One is not anchored and seven are anchored with either
non-prestressed or prestressed CFRP U-straps.

This chapter reports these two experimental investigations together with a


preliminary experimental program. The preliminary experiments are to initially assess
and verify existing prediction models for beams retrofitted with CFRP. The preliminary

133
CHAPTER 5

experiments are also to help identify critical failure modes and establish design
parameters for the main experimental programs.

5.2 Preliminary experimental program


5.2.1 Beam Set-up
Two beams were constructed with dimensions 2500 mm long, 140 mm wide and
258 mm deep. Each beam was reinforced with two tension reinforcement bars 16 mm in
diameter and two compression reinforcement bars 12 mm in diameter. The stirrups were
10 mm in diameter at 125 mm spacing. The beams were retrofitted with CFRP fabrics
2000 mm long by 100 mm wide. One beam was strengthened with two layers of CFRP,
whereas the other beam was strengthened with four layers of CFRP. They were labelled
as B1 and B2, respectively. The material properties are listed in Table 5.1.

Table 5.1 Mechanical properties of materials used


Elastic Compressive Yield Tensile
Material modulus strength strength strength
(MPa) (MPa) (MPa) (MPa)
Concrete 29000 33 - -
Steel ( Y16) 200000 - 400 500
Steel ( Y12) 200000 - 535 625
Steel ( Y10) 200000 - 400 550
CFRP fabrics
230000 - - 3500
(SikaWarp 230C)
Adhesive
3120 - - >concrete
(SikaDur 330)

The beams were loaded in four-point loading with a total span of 2300 mm and a
shear span of 700 mm. A steel clamp was used at one end to force the failure at the other
end. A high-speed video camera was utilised to capture the failure sequences. Strain
gauges were also installed on the CFRP and tension reinforcement.

134
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

All dimensions are in mm


Figure 5.1 Preliminary beam dimensions and test set-up

Beam B1 failed due to intermediate span debond (flexure-shear crack debond) and
beam B2 failed by end debond (end cover peeling) (Figure 5.3). Both beams experienced
a very brittle failure. As shown in Figure 5.2, three regions can be clearly identified
according to the changes in the slope when the concrete cracked and when the tension
reinforcement yielded. As expected, beam B1 had lower stiffness after the yielding of
steel as the bonded composite was thinner. The ultimate loads were however very close
for beams B1 and B2 (149 kN and 155 kN, respectively).

100

80
Shear load (kN)

60
B1
40
B2
20

0
0 10 20 30 40
Midspan deflection (mm)

Figure 5.2 Load-displacement curves of two trial beams

135
CHAPTER 5

(B1)

(B2)

Figure 5.3 Beams B1 and B2 after failure

The CFRP strain distributions are shown in Figure 5.4. The maximum strain level
in beam B1 is much higher than that in beam B2. The average bond stress distributions
were derived from the CFRP strain distributions (using Equation 4.1) and are plotted in
Figure 5.5. Beam B1 has a peak value near its midspan whereas B2 has a peak value near
the composite end. This agrees with the failure modes observed.

Examination of the crack patterns reveals the existence of more flexural cracks
near the midspan and less inclined cracks near the CFRP end in beam B1 than in beam
B2. The clamp proved to be very effective in preventing the formation of inclined cracks
leaving no crack in beam B1 and only two thin cracks in beam B2.

The images from the high-speed video illustrated the failure mechanism of end
cover peeling. At first, when the steel had yielded, visible inclined shear cracks were
observed near FRP end. Figure 5.6a shows a sudden crack propagation originating from
those inclined cracks. The concrete teeth failed simultaneously. Flexure-shear crack
debond had a different mechanism. Inclined cracks appeared closer to the load point.
Those cracks became more inclined as the load increased. Vertical shear deformation was
observed clearly before failure (Figure 5.6b). Cracks propagated towards the CFRP end

136
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

along the concrete layer located approximately 1 to 2 mm above the bond surface. The
concrete cover in the bending region was also sheared off.
Load Load
(kN) 8000 (kN)
10000
10 10
8000 20 6000 20
30 30
Microstrain

Microstrain
6000 40 40
4000
50 50
4000
60 60
65 2000
2000 70
70 75
0 74 0 78
0 500 1000 0 500 1000
Distance from FRP end Distance from FRP end
(mm) (mm)

(a) B1 (b) B2
Figure 5.4 CFRP strain distributions

Load Load
7 (kN) 10 (kN)
10 9 10
6
20 8 20
Bond stress (MPa)
Bond stress (MPa)

5 7
30 30
4 6
40 40
5
3 50 4 50
2 60 3 60
65 2 70
1
70 1 75
0 0
74 78
0 275 550 0 275 550
Distance from FRP end Distance from FRP end
(mm) (mm)

(a) B1 (b) B2
Figure 5.5 Average bond stress distributions in shear span

137
CHAPTER 5

CFRP
end
(a) End cover peeling
(high-speed video
capture)

Load
point
(b) Flexure-shear crack
debond

Figure 5.6 Photos of beams B1 (a) and B2 (b) near failure

5.2.2 Applicability of existing theoretical models


Figure 5.7 and Figure 5.8 show the predicted ultimate loads for intermediate span
debond (beam B1) and end debond (beam B2), respectively. For both failure modes, the
existing models provide very different results. For intermediate span debond, most
models are conservative. Three models which performed best are by Niedermeier (2000),
Arya and Farmer (2001), and Shehata et al. (2001). For end debond, three models which
performed best are Raoof and Hassanen (2000), Shehata et al. (2001) and Nguyen et al.
(2001). These models were used in the design of the beams used in the main experimental
programs.

138
End debond failure shear load (kN)

0
10
20
30
40
50
60
70
80
90
100

68
Shehata et al. (2001)
Intermediate span debond failure shear load (kN)

88
Nguyen et al. (2001) 0
10
20
30
40
50
60
70
80
90
100

59
Fanning and Kelly (2001)

21
Saadatmanesh and Malek (1998) Arya and Farmer (2001)
62 60

40
Tumialan et al. (1999) Shehata et al. (2001)

15
El-Mihilmy and Tedesco (2001) Maruyama and Ueda (2001)
58 56

Oehlers (1992) Teng et al. (2001)

failure load

B2
B1

Experimental
47

38 38
Smith and Teng (2003) Blaschko et al. (1997)
25

46
Jansze (1997) Mohamed Ali (2000)
failure load
Experimental

75

59
Ahmed and Van Gemert (1999) Niedermeier (2000)

39
Ziraba et al. (1994)

77
Raoof and Hassanen (2000): Lower

13
Chaallal et al. (1998)
Neubauer and Rostasy (1997)
90 90

Arya and Farmer (2001)


Figure 5.7 Applicability of theoretical models to predict IS debond loads for beam

Figure 5.8 Applicability of theoretical models to predict end debond loads for beam
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

139
CHAPTER 5

5.3 Design methodology of the main experimental programs


To select the beam dimensions and the reinforcement amounts for the main
experimental programs, the experience from the preliminary experimental program was
taken into account. Different configurations were considered. The key objective was to
ensure that laboratory work was viable while still achieving valid experimental data.
More specifically, the dimensions were chosen to ensure that testing was physically
possible with available equipment, to minimise cost and to enable preparation of the
specimens in a reasonable time frame. The beams were designed so that their shear
capacity was significantly higher than their flexural capacity. The amount of CFRP
reinforcement was also chosen to be significantly high to avoid fibre rupture and thus to
ensure debonding failure. The materials used were provided from local providers.
Concrete was ready-mixed with a strength of around 40 MPa at 28 days. The steel bars
used for flexural reinforcement were deformed bars with a yield stress of approximately
500 MPa. The minimum available deformed bar size was chosen, which was 12 mm. To
ensure that the beams were under-reinforced, the ratio of the neutral depth to the effective
depth, ku, was kept at around 0.19 for non-strengthened sections and from 0.22 to 0.37 for
strengthened sections. These values were below the limit set by AS3600 of 0.40. The
above mentioned requirements led to the selection of a 50 % beam scale. The shear span
to depth ratio was selected to be around 3 to ensure the desire failure modes.

5.4 Testing of beams in four-point bending


5.4.1 Beam details
The selected dimensions of a typical beam are illustrated in Figure 5.9 and Figure
5.10. All of the beams had a total length of 2.7 m. They were 140 mm wide with the
effective depth of 220 mm. The beams were reinforced with longitudinal deformed bars
and vertical closed ligatures made of plain bars. The composites bonded on the soffit of
the beams were 100 mm wide.

A total of eighteen RC beams were constructed. Two were control beams and
sixteen were retrofitted with carbon fibre reinforced fabrics using a wet lay-up method.
The variables for the experiment program are listed in Table 5.2. The retrofitted beams
were divided into two main groups. The E group had ten beams retrofitted with a

140
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

relatively thick layer of CFRP (six or nine plies). The S group had six beams retrofitted
with two plies of CFRP. Two identical beams were manufactured for each configuration
and denoted as ‘a’ and ‘b’. For example, E1 configuration has two beams: E1a and E1b.
The test variables included the CFRP bond length, the area of tension reinforcement, the
concrete cover, the number of CFRP plies and the amount of shear reinforcement.

2N12

46
220
R10*
260*

3N12*

100 CFRP*

140

Notes: * varying parameter


All dimensions are in mm

Figure 5.9 Typical rectangular beam cross-sectional details (beams E1a and E1b)

700 900 700


*
75 8@125 550 8@125* 75
Load point Load point Shear
reinforcement

Longitudinal
Support CFRP Support
reinforcement
150* 150*

200 2300 200

* varying parameter S N
All dimensions are in mm.
Figure 5.10 Typical beam longitudinal details (beams E1a and E1b)

141
CHAPTER 5

Table 5.2 Variables in experimental program No. 1

Label ns x db db.sv - s cover (mm) nf x t.0 L0 (mm)

Control 3 x 12 10 - 125 24 N/A N/A


E1 3 x 12 10 - 125 24 6 x 0.176 150
E2 3 x 12 10 - 125 24 6 x 0.176 350
E3 2 x 12 10 - 125 24 6 x 0.176 150
E4 3 x 12 10 - 125 44 6 x 0.176 150
E5 3 x 12 10 - 125 24 9 x 0.176 150
S1 3 x 12 06 - 125 24 2 x 0.176 150
S2 3 x 12 06 - 90 24 2 x 0.176 150
S3 2 x 12 06 - 125 24 2 x 0.176 150

Notes: ns x db: number of tension steel bars x bar diameter (mm)


db.,sv - s: stirrup diameters (mm) - spacing (mm)
cover: concrete cover to stirrup
nf x tf.0: number of plies x ply thickness (mm)
L0: distance from end of FRP to nearest support

On one side of the retrofitted beams, the longitudinal CFRP was anchored with a
steel clamp as shown in Figure 5.11. The clamp was used to force end failure (if it
occurred) at the other side. This allowed the camera to be focused on one side of the
beam. The clamp comprised of two steel plates 20 mm in thickness connected by two
bars 14 mm in diameter. Four nuts were used to tighten the clamp. To monitor the tension
in the bars, a torque wrench was used together with a strain gauge installed on one bar.
The wrench was set to a torque level so that a strain level of approximately 400
microstrain was introduced to the bolts.

142
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

Concrete beam

Strain gauge
location on bolt

CFRP

Figure 5.11 A steel clamp

5.4.2 Materials
5.4.2.1 Concrete
The concrete material used was supplied separately, pre-mixed from a local
supplier. The concrete was poured in two batches. The first batch was for S beams. The
second batch was for E beams. Compressive tests were carried out in accordance with
AS1012.9-1999 (Standards Australia, 1999). The concrete strengths for S beams and E
beams at the day of the first test in the series were 47.7 MPa (82 days) and 53.7 MPa (67
days), respectively. Further information on the concrete strength development with time
is presented in Appendix B.1.

5.4.2.2 Steel reinforcement


The flexural reinforcement used in the study consisted of grade 500 deformed bars
with a nominal diameter of 12 mm (N12). For the shear reinforcement, rounded bars R10
and R6 were used. Tensile tests in accordance with AS1391-1991 (Standards Australia,
1991) were carried out to determine the properties of the steel bars. The test results are
listed in Table 5.3. The typical stress-strain profiles of the reinforcement are contained in
Appendix B.2. The average yield strains for N12, R10 and R6 were 2688, 1637 and 1777
microstrain, respectively.

143
CHAPTER 5

Table 5.3 Mechanical properties of steel reinforcement


N12 R10 R6
Sample E fsy ft E fsy ft E fsy ft
(MPa) (MPa) (MPa) (MPa) (MPa) (MPa) (MPa) (MPa) (MPa)
1 208000 551 649 205000 334 484 220000 410 578
2 205000 548 650 204000 335 482 235000 420 571
3 202000 553 654 203000 333 484 260000 440 580
Average 205000 551 651 204000 334 483 238000 423 576

5.4.2.3 MBrace Primer, MBrace CF 130 fibre and MBrace Saturant


The MBrace FRP ‘Wet lay up’ system used for the shear-lap specimens described
in Chapter 4 was also used for the retrofitted beams.

5.4.3 Specimen preparation and instrumentation


Surface preparation and bonding operation were carried out similar to those for
the shear-lap specimens. The prepared surfaces after water jetting and after priming are
shown in Figure 5.12. The application sequences are illustrated in Figure 5.13. To check
if the bond was formed properly, additional patches were bonded on the beam ends and
pull-off tests were carried out at these locations. Further information on the tests is
included in Appendix B.4

(a) (b)
Figure 5.12 Surface texture after water jetting (a) and after priming (b)

144
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

(a) (b)

(c) (d)
Figure 5.13 Bonding operation (a) Applying a layer of resin; (b) Placing fibres on
the resin; (c) Pressing fibres into the resin with a ripped roller; (d) Finished surface

The beams were instrumented to record strain, load and deflection readings. Strain
gauges were used to measure strain levels in the steel and CFRP reinforcement. Out of
two identical beams manufactured for each configuration, one beam (beam ‘a’) had more
instrumentation than the other. The strain gauge locations for a typical beam are
illustrated in Figure 5.14. The strain gauges were bonded on the horizontal tension
reinforcement on the bottom side of the middle bar. They are labelled as ‘Hi’ in Figure
5.14, where ‘i’ is the gauge number as counted from right to left. The strain gauges on the
stirrups were located at the middle height of the legs. They are labelled as ‘Via’ or ‘Vib’,
where ‘a’ and ‘b’ correspond to the gauges on the front and back legs, respectively. For
the retrofitted beams, the strain gauges were also placed on the composite surface and
labelled as ‘Ci’. All gauges were protected from the environment by a layer of M-Coat.
For the gauges on the steel reinforcement, wax coating was also applied before pouring
concrete to provide a water resistance layer. The deflections of the beams were measured
at three locations using linear variable displacement transducers (LVDTs) as shown in

145
CHAPTER 5

Figure 5.15. The loads and reactions were measured using load cells placed in the
actuator and the supports.

Clamped side (left) Unclamped side (right)

V4b V4a V3b V3a V2b V2a V1b V1a

H5 H4 H3 H2 H1

C7 C6 C5 C4 C3 C2 C1
Two gauges on two 450 450 125 125 125 125
stirrup legs
S N
All dimensions are in mm.
Figure 5.14 Strain gauge locations on beam E1a

425 725 725 425


Load cell
LVDT
LVDT LVDT

Load cell Load cell

S N
All dimensions are in mm.
Figure 5.15 Load and deflection measurements

A digital video recorder and a high-speed video recorder were used to capture the
crack propagation. The high-speed video recorded at 500 frames per second.

5.4.4 Test set-up and procedure


All of the beams were tested in four-point bending with a total span of 2300 mm.
The shear span was chosen to be 700 mm leading to a shear span to depth ratio of
approximately 3.1. The beams were loaded using the Instron universal testing machine
with a 250 kN load capacity with a piston stroke of 250 mm.

The loading set-up and testing machine details are shown in Figure 5.16. The
beam was placed on two steel support blocks, which seated on two low friction bearing

146
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

strips to allow horizontal movements (Figure 5.17a). Load cells of 100 kN capacity were
placed underneath the bearing steel blocks to measure the reaction at the supports. The
load from the actuator was transferred to the beam through an I-beam and two rollers.
The rollers seated on two steel bearing plates. The I-beam stiffness was chosen to ensure
that its maximum deflection under the maximum applied load was negligible (less than
0.25 mm). The bearing plates were 100 mm wide and 20 mm thick. The actuator was
mounted on a loading frame, which consisted of two steel I crossbeams of 540 mm depth
bolted to two steel I-columns of 253 mm depth and 4500 mm high. Each column was
secured in place using two anchor bolts pre-tensioned to 50 kN against the strong floor as
illustrated in Figure 5.17b.

147
148
CHAPTER 5

Reaction frame

Loading jack

Concrete beam

Supports

Figure 5.16 Test rig


EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

(a) (b)
Figure 5.17 A support (a) and an anchorage of the reaction frame column (b)

Loading was in displacement control. The beams were loaded at a rate of 1 mm


per minute for most of the time.

5.4.5 Experimental results


5.4.5.1 Load-deflection curves, crack patterns and failure modes
5.4.5.1.1 Control beams
The load-deflection curves of the control beams are given in Figure 5.18. The load
was measured at the support on the debonded side and the deflection was measured at the
beam middle.

The two control beams failed by typical steel yielding followed by secondary
compression failure of concrete. The load and crack development were very similar for
both beams C1a and C1b. At approximately 15 kN, fine flexural cracks became visible in
the middle of the beams. At around 35 kN, very fine flexure-shear cracks were observed
in the shear spans. As the load increased, more flexural cracks appeared and propagated
deeper into the beams. When the load reached a value of approximately 54 kN, the tensile
steel yielded and the loading curve levelled off. As loading was continued, the beams
showed wide flexural cracks at their midspan. These cracks extended to the compressive
area and the concrete crushed underneath the load point.

In beam C1a, local crushing and excessive flexural cracking occurred near the
right load point when the deflection was 43 mm. When the deflection reached 63 mm,

149
CHAPTER 5

similar crushing and cracking were also seen under the left load point. As the load
increased further, the most severe cracks were observed under the two load points (Figure
5.19a). Beam C1a showed very ductile behaviour. It could still hold a significant load
when the midspan deflection reached 142 mm. In beam C1b, concrete crushed first near
the left load point when the midspan deflection reached 40 mm. As the beam deflection
increased to 55 mm, more crushing between the left load point and the beam middle was
observed (Figure 5.19b). The beam also experienced ductile behaviour.

60

50
Shear load (kN)

40
Loading
30 terminated C1a

20 C1b

10

0
0 25 50 75 100 125 150
Midspan deflection (mm)

Figure 5.18 Load-deflection behaviour of the control beams

Left load Right load


(a) C1a point point

Left load Right load


(b) C1b point point

Figure 5.19 Photos of beams C1a and C1b after failure

150
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

5.4.5.1.2 E beams
All of the beams in E series failed by end debond on the unclamped side. The
loading curves of these beams are shown together in Figure 5.20. The curves of the
control beams are also shown for reference. The stiffness of the retrofitted beams was
increased significantly compared to that of the control beams. However, the loading
curves of the retrofitted beams also show a more brittle failure. A detailed description of
each beam responses is presented in the following paragraphs.

80
70
60 C1a C1b
Shear load (kN)

50 E1a E1b
40 E2a E2b
E3a E3b
30
E4a E4b
20
E5a E5b
10
0
0 5 10 15 20 25 30
Midspan deflection (mm)

Figure 5.20 Load-deflection behaviour of E beams

The load-deflection responses for E1 beams are plotted in Figure 5.21 and the
crack patterns after failure are shown in Figure 5.22. The crack development of beam E1a
is typical for all beams failing by end debond. Vertical flexural cracks appeared first in
the pure bending region. As the load increased, diagonal flexure-shear cracks became
visible in the shear span. A transverse crack originating from the CFRP end was observed
at 39 kN shear load level. As the load increased further, the cracks in the shear span
widened progressively. At the same time, the transverse crack propagated diagonally into
the concrete cover and then parallel along the tension reinforcement level. It finally
joined the adjacent diagonal crack. At 59 kN, this crack opened further and a portion of
about 90 mm length from the CFRP end was widely open. Since the tensile steel
reinforcement was not yet yielded, further application of load led to gradual transferring
of the tensile force in CFRP to the intact portion and the debond crack propagated into the

151
CHAPTER 5

shear span until the load reached a peak of 70.7 kN. The propagation sequences are
illustrated in the high-speed video captures shown in Figure 5.23. Beam E1b behaved in a
very similar fashion as beam E1a. The ultimate load of beam E1b was 74.5 kN.

80
70
Shear load (kN)
60
50
40 C1a
30 C1b
20 E1a
10 E1b
0
0 5 10 15 20 25 30
Midspan deflection (mm)

Figure 5.21 Load-deflection behaviour of beams E1a and E1b

Right load Diagonal cracks


point Inclined
debond cracks
(E1a)

Right load
point

(E1b)

Figure 5.22 Photos of the debonded side of E1 beams

152
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

Right load
point

(1) (2)

(3) (4)

(5) (6)
Figure 5.23 High-speed video captures for beam E1a at failure (end debond) (rate:
500 frames per second)

The debond crack surface was approximately at the mid-height of the rebars
(Figure 5.24). The debonded cover had several tooth-like transverse cracks, which were
spaced at the stirrup spacing. This was due to the interference of the stirrup extrusions to
the cracking surface (Figure 5.25). As the debond crack was propagating along the weak
plane at the tension reinforcement level, its path was intersected by the stirrup extrusions,
which induced a transverse crack that originated from the tip of the extrusions. This

153
CHAPTER 5

mechanism might have helped to increase the delamination strength and therefore the
load was still able to increase when the debond crack at the composite end had already
opened.

Rebar locations

Figure 5.24 End cover peeling failure surface in beam E1b

Cracking surface
Stirrup extrusions

Induced cracks

Figure 5.25 Influence of stirrup extrusion on debond cracks in beam E1a

The load deflection responses of beams E2a and E2b are plotted in Figure 5.26.
Their crack patterns after failure are shown in Figure 5.27. These beams had a much
shorter bond length. A debond crack near the cut-off point was observed very early. The
crack propagated more gradually since the load level was not high. In both beams, the
debond crack opened wide at the load level of around 49 kN. After that, the load still
increased since the tension steel reinforcement could take more loading. The cracks along
the tension reinforcement level finally caused total separation of the composite and the
concrete cover.

154
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

60
50

Shear load (kN)


40 C1a
30 C1b
20 E2a
E2b
10
0
0 5 10 15 20 25 30
Midspan deflection (mm)

Figure 5.26 Load-deflection behaviour of beams E2a and E2b

Right load
point

(E2a)

Right load
point

(E2b)

Figure 5.27 Photos of the debonded side of E2 beams

Beams E3a and E3b had less tension reinforcement. The load deflection curves
and crack patterns are shown in Figure 5.28 and Figure 5.29, respectively. The failure
occurred in a similar way to that of E1 beams.

155
CHAPTER 5

70
60

Shear load (kN)


50
40 C1a
30 C1b
20 E3a
10 E3b
0
0 5 10 15 20 25 30
Midspan deflection (mm)

Figure 5.28 Load-deflection behaviour of beams E3a and E3b

Right load
point

(E3a)

Right load
point

(E3b)

Figure 5.29 Photos of the debonded side of E3 beams

The concrete covers in beams E4a and E4b were deeper than the other beams.
Their load deflection responses and crack patterns are plotted in Figure 5.30 and Figure
5.31, respectively. The ultimate capacities were not consistent. The maximum loads
recorded for E4a and E4b were 79kN and 61kN, respectively. The reason for this
inconsistency is unknown.

156
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

80
70

Shear load (kN)


60
50
C1a
40
30 C1b
20 E4a
10 E4b
0
0 5 10 15 20 25 30
Midspan deflection (mm)

Figure 5.30 Load-deflection behaviour of beams E4a and E4b

Right load
point

(E4a)

Right load
point

(E4b)

Figure 5.31 Photos of the debonded side of E4 beams

Beams E5a and E5b were bonded with a very thick layer of CFRP. As shown in
Figure 5.32, the stiffness of the beams after cracking was much higher that that of the
control beams. Less cracking was seen in the beam body due to a high amount of tension
reinforcement (Figure 5.33). These two beams failed at slightly lower shear loads
compared to those of E1 beams.

157
CHAPTER 5

70
60

Shear load (kN)


50
40 C1a
30 C1b
20 E5a
10 E5b
0
0 5 10 15 20 25 30
Midspan deflection (mm)

Figure 5.32 Load-deflection behaviour of beams E5a and E5b

Right load
point

(E5a)

Right load
point

(E5b)

Figure 5.33 Photos of the debonded side of E5 beams

5.4.5.1.3 S beams
The load-deflection responses for S beams (the beams retrofitted with two plies of
fibres) are plotted in Figure 5.34. All of S beams failed by yielding of steel followed by
debonding of CFRP. Along the loading curves, three regions can be clearly identified
corresponding to when the concrete was not cracked, when the concrete had cracked and
when the steel reinforcement had yielded. After the concrete had cracked, the retrofitted

158
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

beams showed slightly stiffer behaviour compared to the control beams. After the
yielding of steel, the major increase in tension was transferred to the CFRP and the load
still increased further with increasing deflection.

90
C1a
80
70 C1b

Shear load (kN)


60 S1a
50 S1b
40 S2a
30
S2b
20
S3a
10
0 S3b
0 5 10 15 20 25 30 35 40
Midspan deflection (mm)

Figure 5.34 Load-deflection behaviour of S beams

Beams S1a and S1b failed by a combination of intermediate span debond and end
debond. The debond sequences in beam S1a are illustrated in Figure 5.35. The photos of
the cracking patterns after failure are shown in Figure 5.36. The two beams experienced
very similar cracking behaviour prior to failure. Visible fine flexural cracks were seen at
the load level of approximately 30 kN. These cracks propagated gradually as the load
increased. They were spaced more evenly and closer compared to those in the control
beams. At around 40 kN load level, diagonal flexure-shear cracks became visible. After
the yielding of the steel (at around 61 kN), these cracks appeared more clearly in the
shear spans. They widened as the load increased. Cracking noises could be heard before
failure. Concrete fractured simultaneously from the unclamped end of the CFRP and from
the tip of a wide flexure-shear crack under the right load point. Stress transferred and
cracking propagated toward the beam centre along the tension reinforcement level. The
failure was very brittle.

159
CHAPTER 5

Right load
point

(1) (2)

(3) (4)

(5) (6)
Figure 5.35 High-speed video captures of beam S1a at failure (combination of end
and intermediate span debond) (rate: 500 frames per second)

160
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

Right load
point Initiation locations

(S1a)

Right load
point
Initiation locations

(S1b)

Figure 5.36 Photos of the debonded side of beams S1a and S1b

Beams S2a and S2b failed by intermediate span debond. The full loading curves
of these two beams are plotted in Figure 5.37. The failures were slightly different in the
two beams. In beam S2a, debonding occurred on the clamped side. A wide flexure-shear
crack was observed under the load point and delamination of the laminate from concrete
was initiated at its tip as the shear load reached 78 kN. This was followed by gradual
delamination of the composite from the concrete beam along the bond surface toward the
clamp location (crack 1 in Figure 5.39). The crack path was within a thin layer of
concrete near the bond surface (Figure 5.40). The load continued to increase to a peak of
80.4 kN until the concrete cover near the tip of the flexure-shear crack broke from the
beam leading to complete separation of the composite from the crack location to the steel
clamp. The load level stayed at around 70 kN as the CFRP was still held by the clamp. As
the prescribed displacement increased further, severe crushing of concrete occurred under
the load points. A crack along the tension reinforcement layer also formed (crack 2 in
Figure 5.39). As the load increased further, more cracks formed in the concrete cover
(cracks 3 in Figure 5.39). Eventually, the entire concrete cover on the right side was
sheared off and the load decreased to approximately 50 kN. In beam S2b, delamination
occurred on the unclamped side. The debonding sequences of this beam were captured

161
CHAPTER 5

with the high-speed camera and are illustrated in Figure 5.38. As the composite was
peeled from the beam towards the right side of a flexure-shear crack, the unbalanced
force in the CFRP transferred to the concrete cover on the left and led to the fracture of
the cover. The failure was brittle. The ultimate load of beam S2b was slightly less than
that of beam S2a.

Intermediate Total
100 span debond separation
of CFRP
80
Total load (kN)

60

40 S2a
S2b
20

0
0 20 40 60 80 100
Midspan deflection (mm)

Figure 5.37 Full loading curves of beams S2a and S2b

162
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

Right load
point

(1) (2)

(3) (4)

(5) (6)
Figure 5.38 High-speed video captures of beam S2b at failure (intermediate span
debond) (rate: 500 frames per second)

163
CHAPTER 5

Left load
point
(S2a) Debond sequence
1 2 3

Right load
point

(S2b)
Initiation location

Figure 5.39 Photos of the debonded side of beams S2a and S2b

Initiation location

Figure 5.40 Failure surface in beam S2a

Beams S3a and S3b had less tension reinforcement, which was indicated by their
lower stiffness as seen in Figure 5.34. Beam S3a failed by intermediate span debond,
which first appeared on the clamped side. Beam S3b failed due to a combination of
intermediate span and end debond on the unclamped side (Figure 5.41).

164
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

Left load
Initiation point point
(S3a)

Right
load point
Initiation points
(S3b)

Figure 5.41 Photos of the debonded side of beams S3a and S3b

5.4.5.1.4 Beam strength after total debonding of CFRP


For a number of beams, after the CFRP was delaminated from the RC beam, the
loading was continued until secondary compression failure of concrete occurred. It was
found that after detachment of the strengthening material, the beams kept their original
capacity as without retrofitting (Figure 5.42).
80
70
C1a
60
Shear load (kN)

C1b
50 E2a (retested)
40 E2b
30 E4a
20 E4b
10 E5b
0
0 5 10 15 20 25 30
Midspan deflection (mm)
Figure 5.42 Loading curves for all beams tested until concrete crushed

5.4.5.1.5 Summary
A summary of the failure modes and load capacities of the beams is presented in
Table 5.4. The crack patterns of all beams are compared in Figure 5.43 to Figure 5.45.

165
CHAPTER 5

Except for the beams strengthened with a very short length of CFRP, the ultimate
capacities of retrofitted beams were increased compared to the control beams. For E
beams, which were strengthened with 6 or more layers of CFRP, the ultimate load
capacities were increased up to 38 %. For S beams, which were strengthened with 2
layers of FRP, the maximum increase in the capacity was 49 %.

Table 5.4 Summary of beam test results


Load at 1st Ultimate load Load capacity
Label Failure mode debond crack* capacity after debond
(kN) (kN) (kN)
C1a Concrete crushing N/A 54.4 N/A
C1b Concrete crushing N/A 53.9 N/A
E1a End cover peeling 55.0 70.7 N/A
E1b End cover peeling 59.0 74.6 N/A
E2a End cover peeling 26 51.4 52.5
E2b End cover peeling N/A 53.4 N/A
E3a End cover peeling 60.0 66.0 N/A
E3b End cover peeling 58.0 65.2 N/A
E4a End cover peeling 59.5 79.0 56.5
E4b End cover peeling N/A 61.2 54.5
E5a End cover peeling 57.5 63.3 N/A
E5b End cover peeling 48.5 63.2 54.0
Flexure-shear crack debond
S1a & end cover peeling 47.0 73.8 N/A
Flexure-shear crack debond
S1b & end cover peeling N/A 74.5 N/A
S2a Flexure-shear crack debond 74.0 (after peak) 80.4 N/A
S2b Flexure-shear crack debond N/A 74.5 N/A
S3a Flexure-shear crack debond 60.3 60.3 N/A
Flexure-shear crack debond
S3b & end cover peeling N/A 60.2 N/A

* Load level where the strain level at 50mm from the CFRP end reached its peak.

166
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

It is visually clear that the flexural crack widths were reduced when the CFRP
composite were bonded to a beam. The thicker CFRP, the narrower the crack widths
were. The crack spacing in constant moment region was also more uniformly distributed
in retrofitted beams.

C1a

C1b

Figure 5.43 Crack patterns in control beams

S1a S1b

S2a S2b

S3a S3b
Figure 5.44 Crack patterns in S beams

167
CHAPTER 5

E1a E1b

E2a E2b

E3a E3b

E4a E4b

E5a E5b
Figure 5.45 Crack patterns in E beams

5.4.5.2 Reinforcement strains and stresses


The tension reinforcement strain distributions and stirrup strain development in
the control beam C1a are plotted in Figure 5.46 and Figure 5.47. The tension
reinforcement strain distributions of beam C1a are typical for under-reinforced beams.
The maximum strains are in the constant moment region and reduced gradually over the
shear span. At the low load levels, the slope of the strain distributions in the shear span is
close to being linear. At the higher loads, with the presence of flexure-shear cracks in the
shear span, the curves become more nonlinear. The steel near the beam centre yielded
first and after that, the steel strain increased rapidly. The readings from the strain gauges

168
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

on the stirrup indicate no yielding of steel. The strains at different stirrup locations along
the shear span also have comparably similar values.

The CFRP strain distributions, bond stress distributions, CFRP strain


developments, tension reinforcement strain distributions and stirrup strain developments
in three example retrofitted beams E1a, S1a and S2a are plotted in Figure 5.48 to Figure
5.62. The plots for other beams are shown in Appendix D.2. In the stirrup strain
distribution plots, the strains on both sides of the stirrup are shown and labelled as ‘F’ for
the front leg and ‘B’ for the back leg. The average bond stress between the two strain
gauge locations was calculated by dividing the force difference in the CFRP by the bond
area as given by Equation 4.2.

The CFRP strain in all retrofitted beams is found to decrease from a maximum
value under the load point to a minimum value at the end of the composite (Figures 5.48,
5.53 and 5.58). The descending slopes in the shear span are close to being linear for
beams S1a and S2a. However, for beam E1a, the curves have a nonlinear distribution
with the slope reducing gradually from the CFRP end to around zero under the load point,
especially at higher loads. As expected, the maximum strain level in beams S1a and S2a
is much higher than that in beam E1a.

With the exception at the peak load, the average bond stress distributions of
beams S1a and S2a appear to have two peaks: one near the CFRP end and one under the
load point; whereas the distributions of beam E1a show only one clear peak near the FRP
end (Figures 5.49, 5.54 and 5.59). These are the indication of the mixed mode of end and
intermediate span debond in S beams and end debond in E beams. For beam S2a, since
the failure happened to be on the clamped side, where the end debond was prevented, this
beam failed by intermediate span debond.

The effects of concrete cracking, steel yielding and debonding are illustrated
clearly in the CFRP strain development plots (Figures 5.50, 5.55 and 5.60). Concrete
cracking and steel yielding influence the tension in the composite noticeably. Before
concrete cracking, the strain level in the composite is very small. In beams S1a and S2a,
after concrete cracking at approximately 10 kN, the CFRP strain gradually increases with
the load. After yielding of steel at around 65 kN, there is a steep increase in the CFRP

169
CHAPTER 5

strain indicating that the major increase in tension was taken by the composite. In beam
E1a, except at the ultimate load, the longitudinal steel did not yield and the CFRP strain
increased gradually until the ultimate load was reached. The CFRP strain development
also reveals early debonding near the CFRP end at a load level much lower than the
ultimate load. Early sign of debonding is indicated by the change of the curve slope
recorded by the gauge at 50 mm from the composite end. This change occurred at 47 kN
for both beam S1a and beam S2a, and at 41 kN for beam E1a. The composite is debonded
when the slope becomes negative. The load levels when that occurred for all beams are
listed in Table 5.4

The tension reinforcement strain distributions were changed when CFRP


composites were bonded on the beam (Figures 5.51, 5.56 and 5.61). At the low loads (less
than 30 kN), the strain distributions in the retrofitted beams are similar as those in the
control beams. However, at the higher loads, higher steel strains are observed along the
shear span in all retrofitted beams. This effect is most clear in beam E1a, where the steel
strains are almost constant along the bonded length of the CFRP at the peak load. This is
an important indication for designing retrofitted beams. Certain amount of steel tension
reinforcement is needed near the CFRP end even though the applied moment can be low
there.

The strain levels in the stirrup closer to the composite end also appear to be
slightly higher than those under the load point. However, this is not clearly demonstrated
considering the variation in the measurement and the fact that the strain depends largely
on the position of major cracks. Except for the stirrup near the CFRP end in beam S1a,
yielding of the steel was not seen.

170
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

Load point Load (kN)


5000
10.1
4000 15.1
20.0

Microstrain
3000 25.0
30.0
2000 35.1
40.0
1000 45.1
50.0
0 53.6
1000 750 500 250 0
Distance from left support (mm)

Figure 5.46 Tensile steel strain distributions in beam C1a

1400 Location
1200 from left
support (mm)
1000
Microstrain

800 250 F
600 250 B
500 F
400
500 B
200
625 F
0 625 B
0 10 20 30 40 50
Shear load (kN)

Figure 5.47 Stirrup strain development in beam C1a

171
CHAPTER 5

Load point Load (kN)


4000
10.0
3500
20.1
3000
Microstrain 2500 30.0
2000 40.0
1500 50.1
1000 60.0
500 65.1
0 70.7
1000 750 500 250 0
Distance from end of FRP (mm)

Figure 5.48 CFRP strain distributions in beam E1a on unclamped side

Load point Load (kN)


1
10.0
0.8
20.1
0.6
Stress (MPa)

30.0
0.4 40.0
0.2 50.1
0 60.0
-0.21000 750 500 250 0 65.1
-0.4 70.7
Distance from end
of FRP (mm)

Figure 5.49 Bond stress distributions in beam E1a on unclamped side

4000 Location from


3500 CFRP end (mm)
3000 50
Microstrain

2500
175
2000
1500 300
1000 425
500 550
0 1000
0 20 40 60 80
Shear load (kN)

Figure 5.50 CFRP strain development in beam E1a on unclamped side

172
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

Load point Load (kN)


3500
10.0
3000 20.1
2500 30.0

Microstrain
2000 40.0
1500 50.1
55.1
1000
60.0
500 65.1
0 70.7
1000 750 500 250 0
Distance from end of FRP (mm)

Figure 5.51 Tensile steel strain distributions in beam E1a on unclamped side

1400 Location from


1200 CFRP end (mm)
1000 100 F
Microstrain

800 100 B
600 350 F
400 350 B
200 475 F
0 475 B
0 20 40 60 80
Shear load (kN)

F: front leg; B: back leg

Figure 5.52 Stirrup strain development in beam E1a on unclamped side

173
CHAPTER 5

Load point Load (kN)


10000
10.0
8000 20.0
30.0

Microstrain
6000 40.1
50.0
4000
60.0
2000 65.0
70.0
0 73.7
1000 750 500 250 0
Distance from end of FRP (mm)

Figure 5.53 CFRP strain distributions in beam S1a on unclamped side

Load point Load (kN)


2.5
10.0
2 20.0
Stress (MPa)

1.5 30.0
40.1
1 50.0
0.5 60.0
65.0
0
70.0
-0.51000 750 500 250 0
73.7
Distance from end
of FRP (mm)

Figure 5.54 Bond stress distributions in beam S1a on unclamped side

10000 Location from


CFRP end (mm)
8000
50
Microstrain

6000 175
4000 300
425
2000
550
0 1000
0 20 40 60 80
Shear load (kN)

Figure 5.55 CFRP strain development in beam S1a on unclamped side

174
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

Load point Load (kN)


10000
10.0
8000 20.0
30.0

Microstrain
6000 40.1
50.0
4000
60.0
2000 65.0
70.0
0 73.7
1000 750 500 250 0
Distance from end of FRP (mm)

Figure 5.56 Tensile steel strain distributions in beam S1a on unclamped side

4500 Location from


4000 CFRP end (mm)
3500 100 F
3000
Microstrain

100 B
2500
2000 350 F
1500 350 B
1000
475 F
500
0 475 B
0 20 40 60 80
Shear load (kN)

F: front leg; B: back leg

Figure 5.57 Stirrup strain development in beam S1a on unclamped side

175
CHAPTER 5

Load point Load (kN)


10000
10.0
8000 20.1
30.0
Microstrain 6000 40.0
50.0
4000
60.0
2000 70.0
75.0
0 80.2
1000 750 500 250 0
Distance from end of FRP (mm)

Figure 5.58 CFRP strain distributions in beam S2a on unclamped side

Load point Load (kN)


2
10.0
1.5 20.1
30.0
Stress (MPa)

1 40.0
50.0
0.5 60.0
70.0
0
75.0
1000 750 500 250 0
-0.5 80.2
Distance from end
of FRP (mm)

Figure 5.59 Bond stress distributions in beam S2a on unclamped side

10000 Location from


CFRP end (mm)
8000
50
Microstrain

6000 175
4000 300
425
2000
550
0 1000
0 20 40 60 80
Shear load (kN)

Figure 5.60 CFRP strain development in beam S2a on unclamped side

176
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

Load point Load (kN)


10000
10.0
8000 20.1

Microstrain
30.0
6000 40.0
50.0
4000
60.0
2000 70.0
75.0
0 80.2
1000 750 500 250 0
Distance from end of FRP (mm)

Figure 5.61 Tensile steel distributions in beam S2a on unclamped side

2000 Location from


CFRP end (mm)
1500 85 F
Microstrain

85 B
1000 445 F
445B
500 535 F
535 B
0
0 20 40 60 80
Shear load (kN)

F: front leg; B: back leg

Figure 5.62 Stirrup strain development in beam S2a on unclamped side

The maximum CFRP strain and maximum bond stress for all retrofitted beams are
listed in Table 5.5. Since the bond stress was calculated based on the strain reading at two
locations, the calculated result was not accurate once debonding had occurred at a strain
gauge location. For those cases, the maximum bond stress values were taken from the
previous load level. The bond stress values are only available for the beams instrumented
with enough strain gauges.

It can be seen that for S beams, the maximum CFRP strains varied around an
average value of 8615 microstrain with the variation being less than 12 % for all cases.
For E beams, it can be seen that end debond occurred at shear bond stress levels varying
around 1 MPa. However, the calculated bond stresses near the CFRP end were less

177
CHAPTER 5

accurate due to the fact that they depended on the CFRP strains near the CFRP end where
the readings were more susceptible to noise.

Table 5.5 Maximum CFRP strains and bond stresses


Maximum FRP Maximum FRP Maximum bond Maximum bond
Beam strain under strain near stress near stress near load
label load point beam's centre CFRP end point
(microstrain) (microstrain) (MPa) (MPa)
E1a 3036 3577 0.88 -
E1b 3417 3912 1.38 -
E2a 1740 2032 0.63 -
E2b 2142 2080 - -
E3a 3681 3502 1.48 -
E3b 3996 3624 1.26 -
E4a 3386 3492 0.82 -
E4b 2286 2685 - -
E5a 2329 2909 0.84 -
E5b 2350 2496 - -
S1a 8571 9445 1.19 0.79
S1b 9437 9760 - -
S2a 9014 9176 0.95 1.04
S2b 7614 9232 - -
S3a 8707 8080 1.4 1.2
S3b 8345 9007 - -

5.4.5.3 Parametric study


In this section, the influences of several parameters are investigated. In addition to
the loading curves, the effectiveness of CFRP is also compared using a dimensionless
ratio, which is the ratio of the debond capacity Vdebond to the maximum capacity of the
retrofitted section Vmax as calculated by the beam theory assuming full composite action.
The larger the ratio, the more efficiently the CFRP is used. When the ratio is equal to 1,
full composite action is achieved.

178
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

As shown in Figure 5.63a, the amount of the CFRP and tension steel
reinforcement clearly affects the beam behaviour. It is apparent that retrofitting RC beams
with a thicker layer of CFRP does not always lead to higher capacity. The average
ultimate load for the beams retrofitted with 2, 6 and 9 layers of FRP were 148.3 kN, 145.3
kN and 126.4 kN respectively. Figure 5.63b clearly shows that the effectiveness of the
CFRP is greatly reduced as the relative stiffness of CFRP to steel increases.

6 plies 2 plies
3 bars 3 bars
80 9 plies 1
70 3 bars
0.8
60
Shear load (kN)

Vdebond / Vmax
50 0.6
40
2 plies 0.4
30
2 bars
20 0.2
6 plies
10 2 bars
0 0
0 5 10 15 20 25 30 35 0 0.1 0.2 0.3 0.4 0.5 0.6
Midspan deflection (mm) A f Ef / A s Es

(a) (b)
Figure 5.63 Effect of CFRP and tension steel stiffness ratio

The bond length in the shear span, Lf, also has a significant influence on the
beam’s capacity. As illustrated in Figure 5.64, the beam capacity reduces by 39 % as the
bond length is shortened by 36 %.

80 1
70 Lf = 550 mm
0.8
60
Shear load (kN)

Vdebond / Vmax

50 0.6
40 Lf = 350 mm
30 0.4
20 0.2
10
0 0
0 5 10 15 20 25 30 0 0.25 0.5 0.75 1
Midspan deflection (mm) Lf / a

(a) (b)
Figure 5.64 Effect of CFRP bond length

179
CHAPTER 5

Concrete cover and stirrup spacing seem to have an insignificant effect on the beam
capacity.

These remarks will be investigated further in Chapters 7 and 9 by comparing with the
results from numerical and theoretical modelling.

5.4.5.4 Effect of steel clamping


The crack patterns on the clamped side of two example beams E1a and S1a are
shown in Figure 5.65. Steel clamping proved to be very effective in preventing end
debond. Fewer cracks were present on the clamped side and the cracks were generally
much less widely open compared to those in the unclamped side. The clamp also limited
the formation of horizontal inclined debonding cracks because of a higher concrete shear
strength in the cover resulted from the confinement by the bottom steel plate. As a result,
all E beams failed on the unclamped side.

End steel clamping was not able to prevent intermediate span debond. Two of S
beams, S2a and S3a, failed by intermediate span debond on the clamped side as described
previously. However, end steel clamping improved ductility of the beams significantly.
After intermediate span debond occurred, the CFRP was still tightened to concrete by the
clamp and the beam continued to carry a significant portion of the peak load. This was
clearly observed in beam S2a as described previously.

180
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

Left load
(a) S1a: point
Clamped
side

Left load
point
(b) E1a:
Clamped
side

Figure 5.65 Crack patterns on the clamped side of E1a and S1a

From the above observation, it is clear that steel clamping is a simple and
effective anchorage method. However, steel poses corrosion problems. Another possible
set back of steel clamping is that it needs to connect to the top of the beam. It means that
the top flange of T-beams and the top slab need to be drilled through. This might not be
practical for some cases. A better anchorage method is using composite U-straps. This is
investigated in the next section.

181
CHAPTER 5

5.5 Testing of beams in three-point bending


5.5.1 Beam details
In this experimental program, a total of eight tests were carried out on retrofitted
beams to investigate the performance of CFRP U-strap systems for anchorage. The tested
beams had similar configurations to those described in section 5.4. All the beams were
tested in three-point bending until failure with a shorter shear span of 700 mm. The same
reaction frame, testing machine and support systems as in the first experimental program
were used with an exception that the left support was moved to under the left load point
of the first test. The test variables are summarised in Table 5.6. Four tests were done on
the less damaged side of the four beams tested in the first experimental program (beams
E1a, E3a, E3b and E5a). These four beams were selected since they were subjected to a
minimum damage on the clamped side from the first test (the load was stopped
immediately after end debond occurred on the other side). Four other tests were also
performed on two newly manufactured beams, A1 and A2, of similar configurations as
beam E1.

Out of these eight tests, one was carried out on the beam without anchorage. In the
other seven tests, the beams were anchored with non-prestressed and prestressed CFRP
U-straps placed either at the CFRP end only or at a spacing of 180 mm in the shear span.
In the non-prestressed anchorage system, two plies of CFRP fabrics of 50 mm width were
wrapped and bonded around the sides and the soffit of the concrete beam near the end of
the longitudinal CFRP (Figure 5.66). In the prestressed system, a gap was introduced
between the strap and concrete soffit. Prestressing was introduced into the sides of the
CFRP strap by inserting a wedge into the gap (Figure 5.67). More description of this new
prestressed system is followed.

The strap had at a slope of 1:20 to the vertical, which was also the wedge surface
slope. This slope was chosen so that a prestressing strain of 500×10-6 could be introduced
to the strap sides. The prestressing strain was monitored by two strain gauges placed on
the two sides of the strap over the un-bonded area close to the beam soffit. To avoid
slipping of the wedge over the longitudinal composite, grooves were introduced to the
wedge bond surface. The grooves were designed for one-directional slippage only. To
further prevent slippage, epoxy resin was also injected between the grooves the wedge.

182
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

Slipping was however allowed between the wedge and the bottom side of the U strap.
This method of prestressing was designed specifically for CFRP fabrics bonded using wet
lay-up method since the fibres can not be stressed prior the formation of the composite
due to breaking of the individual strands. Another prestressing method is also available
for CFRP thermoplastic tape and has been reported elsewhere (Lees et al., 2002).

Since no stirrup was provided between the load points, additional steel clamps
were used there to improve the beam shear strength. The steel clamps should have an
insignificant effect on the failure in the shear span.

Table 5.6 Variables in experimental program No. 2

1 1
Cover1
Label ns x db db,sv – s nf x tf,01 L0 (mm) 1 U-strap
(mm)
E3b2 2 - 12 10 - 125 25 6 - 0.176 150 None
E3a2 2 - 12 10 - 125 25 6 - 0.176 150 1-N2
E1a2 3 - 12 10 - 125 25 6 - 0.176 150 1-P26
E5a2 3 - 12 10 - 125 25 9 - 0.176 150 3-P5
A1a 3 - 12 10 - 125 25 6 - 0.176 150 1-N2
A1b 3 - 12 10 - 125 25 6 - 0.176 150 3-N3
A2a 3 - 12 10 - 125 25 6 - 0.176 150 1-P4
A2b 3 - 12 10 - 125 25 6 - 0.176 150 3-P5

Notes: 1) Refer to the notes of Table 5.2.


2) 1-N: one end non-prestressed U-strap
3) 3-N: three non-prestressed U-straps @ 180 mm spacing
4) 1-P: one end prestressed U-strap
5) 3-P: three prestressed U-straps @ 180 mm spacing
6) 1-P2: the end strap was moved closer to the beam middle by 60 mm due to unsuccessful
prestressing of the first end strap.

183
CHAPTER 5

Load point

Support Support

U-strap

Longitudinal
CFRP

Figure 5.66 Non-prestressed U strap system

Load point

Support Support

CFRP concrete
U-strap
steel reinforcement
duct tape
CFRP
PERSPEX grooves
Longitudinal wedge no bonding
CFRP

Figure 5.67 Prestressed U strap system

184
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

5.5.2 Materials
The material properties are described in Section 5.4.2.

5.5.3 Load set-up, specimen preparation and instrumentation


The beams were retrofitted in two stages. In the first stage, the longitudinal fibres
were bonded on the beam soffit in a similar way as for the beams in experimental
program No. 1. In the second stage, the U-straps were installed.

The installation of non-prestressed straps involved surface preparation and


bonding of fibres. Surface preparation included rounding off the soffit edges, roughening
the bond area, cleaning the bond area and applying the primer. The first two steps were
carried using a hand disc sander. For a small bonding area, this method proved to be
effective in providing an adequately rough surface suitable for bonding operation. A
bonding operation included resin under-coating, carbon fibre sheet application and resin
over-coating as for the longitudinal fibres. Photos of this system are shown in Figure
5.68.

For the prestressed straps, the installation involved surface preparation, bonding
of the U-strap and prestressing of the U-strap. To create a gap between the U-strap and
concrete soffit, a PERSPEX ‘forming’ wedge was attached to the soffit. The wedge was
wrapped with a plastic sheet to prevent it from bonding to the fibres. Two peaces of duct
tape were also adhered on the sides of the concrete beam within the concrete cover
portion in order to leave two un-bonded areas for monitoring prestressed strain level. The
fibre strips were applied on the concrete as for the non-prestressed system. After that, it
was left to cure. During that time, strain gauges were installed. After seven days of
curing, the ‘forming’ wedge was then removed and a prestressing wedge was inserted to
the gap. The prestressing wedge was forced into the gap using a plastic hammer until the
strain level at the control area reached the desire level (using the monitoring strain gauges
on the two legs). Epoxy was also injected to the contact surface between the wedge and
the longitudinal CFRP after the wedge had been positioned. Together with the grooves on
the wedge surface, a strong bond was formed between the wedge and the longitudinal
CFRP. Photos of this system are presented in Figure 5.69.

185
CHAPTER 5

After surface preparation After bonding of U-straps


Note: The beam is upside down in the photos.
Figure 5.68 Non-prestressed U-strap system

After surface preparation After bonding of U-straps


Note: The beam is upside down in the photos.
Figure 5.69 Prestressed U-strap system

The location of LVDTs and load cells are shown in Figure 5.70. The strain gauges
were placed on the steel and the longitudinal CFRP at similar locations to the beams
tested in four-point bending. A number of strain gauges were also installed on the U-
straps. The locations of those are shown in Figure 5.71. A digital camera was also used to
capture crack propagation. The beams were loaded at a rate of 1 mm per minute for most
of the time.

186
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

900 700
450 725 425
Load cell
LVDT LVDT
LVDT

Load cell Load cell

All dimensions are in mm

Figure 5.70 Load and deflection measurement for beams tested in three-point
bending

30 30

80 80
80 80
30 30

All dimensions are in mm


(a) Non-prestressed system (b) Prestressed system
Figure 5.71 Strain gauge locations on U-straps

5.5.4 Experimental results


5.5.4.1 Load-deflection curves, crack patterns and failure modes
The loading curves are shown in Figure 5.72 and Figure 5.73 for all tests. The
curves show that with the addition of U-straps, the ultimate capacities increased

187
CHAPTER 5

significantly compared with those of the beams without U-straps. Detailed descriptions
for each beam are presented in the following paragraphs.

120
E5a2
100 E1a2 E1a load capacity
E3a2
Shear force (kN)
80
E3a load capacity
E3b load capacity
60
E5a load capacity
40
E3b2
20

0
0 5 10 15 20 25 30 35
Deflection under load (mm)

Figure 5.72 Load-deflection behaviour of retested beams

A2b
120
A1b
100
A2a E1a load capacity
A1a
Shear force (kN)

80

60

40

20

0
0 5 10 15 20 25 30 35
Deflection under load (mm)

Figure 5.73 Load-deflection behaviour of beams A1a, A1b, A2a and A2b

Beam E3b2 was a retest of beam E3b in three-point bending. It had no U-strap. It
failed by end debond in a similar fashion as beam E3b (Figure 5.74). Its ultimate load
carrying capacity was however slightly less than that of beam E3b.

188
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

Load point

Figure 5.74 Photo of the debonded side of beam E3b2

Beams E3a2 and A1a were anchored with one end non-prestressed strap. The end
strap shifted the failure mode to intermediate span debond, which occurred at a higher
load (Figure 5.75). Similar flexural and flexure-shear cracks were observed as in the
beams without anchorage (beams E3a and E1a). The end debond crack was also seen but
its propagation was restricted significantly by the U-strap. As the load increased beyond
the capacity of the beams without anchorage, more inclined cracks formed in the concrete
cover and intermediate span debond was initiated from the tip of one of those cracks. As
this occurred, the load dropped. However, the strap held the longitudinal composite and
the beam was still able to hold a significant portion of the peak load. As the tension in the
CFRP increased further, sliding of the longitudinal plate underneath the U-strap was
observed visually. The sliding action caused bending of the strap legs near the beam soffit
and led to local debonding and rupture of one of the legs before the longitudinal CFRP
was complete separated from the beam (Figure 5.76).

189
CHAPTER 5

Load point

(E3a2)

Load point

(A1a)

Figure 5.75 Photos of the debonded side of beams anchored with a non-prestressed
U strap

rupture
end debond
crack
debonding

sliding

Figure 5.76 Sliding of the longitudinal CFRP and local debonding and rupture of a
strap leg in beam E3a

Beams E1a and E2a were anchored with one end prestressed strap and they failed
by intermediate span debond in a similar way as in beams E3a2 and A1a (Figure 5.77).
However, no evidence of end debond crack or sliding of the longitudinal composite near

190
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

its end was found. This was due to the higher concrete shear capacity resulted from the
compressive stress field caused by the prestressing force. A leg of the strap was
eventually delaminated from concrete surface (Figure 5.78).

Load point

(E1b2)

Load point

(A2a)

Figure 5.77 Photos of the debonded side of beams anchored with a prestressed U
strap

total debonding

No sign of end
debond crack

Figure 5.78 Debonding of a leg of the prestressed strap in beam A1a

191
CHAPTER 5

The beams with three U-straps (E5a2, A1b and A2b) failed by debonding of the
straps followed by intermediate span debond (Figure 5.79 and Figure 5.80). The straps in
the shear span crossed the flexure-shear cracks and limited their opening. They also
prevented the formation of inclined cracks in the concrete cover and increased the bond
strength between the concrete and longitudinal CFRP. As a result, both end debond and
intermediate span debond were limited. Debonding of the vertical strap nearest to the load
point occurred first. It was followed by intermediate span debond occurring near the
debonded strap location. As the tension transferred to other two straps, they were
subsequently subjected to either local debonding and rupture (in the beam bonded with
non-prestressed straps) or total peeling off (in the beams bonded with prestressed straps).

Load point

(A1b)

Figure 5.79 Photos of beams anchored with three non-prestressed U straps after
failure

192
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

Load point

(E5a2)

Load point

(A2b)

Figure 5.80 Photos of beams anchored with three prestressed U straps after failure

A summary of failure modes and load capacities of the beams is presented in


Table 5.7. The crack patterns of all beams are compared in Figure 5.81 to Figure 5.83. In
the beams anchored with one strap, the failure mode shifted to intermediate span debond
and the ultimate load capacities increased by 15 to 44 %. For the beams with three U
straps, the load capacities increased significantly (up to 79 %) compared to the beams
without anchorage since both end and intermediate span debond mode were limited.
Compared to the non-prestressed system, the prestressed only increased the beam strength
slightly.

193
CHAPTER 5

Table 5.7 Experimental results


Ultimate
Increase
shear
Label U-strap1 Failure mode (compared
load
to)
(kN)
E3b2 None End debond 60.0 N/A
E3a2 1-N Intermediate span debond & rupture of end strap 75.9 15 % (E3a)
E1b2 1-P2 Intermediate span debond & debond of end strap 98.1 32 % (E1b)
E5a2 3-P Debond of straps & intermediate span debond 113.1 79 % (E5a)
A1a 1-N Intermediate span debond & rupture of end strap 94.5 34 % (E1a)
Debond and rupture of straps & intermediate span
A1b 3-N 108.7 54 % (E1a)
debond
A2a 1-P Intermediate span debond & debond of end strap 101.7 44 % (E1a)
A2b 3-P Debond of straps & intermediate span debond 114.5 62 % (E1a)

Note: 1) Refer to the notes for Table 5.6

Figure 5.81 Crack patterns in beam E3b2

E3a2 E1b2

A1a A2a
Figure 5.82 Crack patterns in beams with one U-strap

194
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

E5a2

A1b A2b
Figure 5.83 Crack patterns in beams with three U-straps

5.5.4.2 Reinforcement strain distributions


The CFRP reinforcement strain distributions and development, and the tensile
steel strain development of the newly manufactured beams (A1a, A1b, A2a and A2b), are
plotted in Figure 5.84 to Figure 5.100. The results of the retested beams are included in
Appendix D.3.

The CFRP strain distributions (Figures 5.84, 5.87, 5.92 and 5.96) show high strain
values extending deep into the shear span. This was due to the fact that more severe
cracking was present in the cover, which led to more stress redistribution in the
composite. The CFRP strain development curves also indicate three clear regions
corresponding to when the concrete was uncracked, when the concrete was cracked and
when the steel yielded. Once the CFRP was debonded, its strain decreased rapidly as
expected.

The tensile steel strain development in the beams with anchorage (Figures 5.91,
5.95 and 5.100) shows clear yielding of steel. After debonding of the CFRP, the steel
rebar strain increased promptly as all tension transferred to the rebar.

The strain development in the strap legs (Figures 5.86, 5.90, 5.94 and 5.99) shows
that the straps only became effective after the load reached approximately 60 kN. The U-
strap strain increased significantly after the peak as the longitudinal CFRP was separated
from the beam. As it occurred, the tension was resisted by the end friction, which was
mainly due to the interlocking mechanism as a result of the confinement from the U-strap.

195
CHAPTER 5

The crack frictional faces dilated and therefore imposed more strain in the strap legs until
the legs ruptured or debonded.

The strain distributions along the strap height (Figures 5.89 and 5.98) near the
ultimate load levels were affected by the intersected cracks. In beam A1b, the crack
located at 100 mm height along the leftmost strap; whereas in beam A2b, the crack was at
approximately 200 mm height.

196
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

Load point
5000 Load (kN)
20.0
4000
40.0

Microstrain
3000 50.0
60.1
2000
70.0
1000 80.0
0 90.0
500 400 300 200 100 0 94.5

Distance from end of FRP (mm)

Figure 5.84 Longitudinal CFRP strain distributions in beam A1a

5000

4000 Location from


CFRP end (mm)
Microstrain

3000 137.5
2000 275
412.5
1000
550
0
0 20 40 60 80 100
Shear load (kN)

Figure 5.85 Longitudinal CFRP strain development in beam A1a

3500
3000 Gauge
2500 location
Microstrain

2000 Front
1500 Back
1000
500
0
-500 0 20 40 60 80 100
Shear load (kN)

Figure 5.86 Strain development in the U-strap legs in beam A1a

197
CHAPTER 5

Load point Load (kN)


6000
20.0
5000 40.0

Microstrain
4000 60.1
3000 70.1
2000 80.0
90.0
1000
100.0
0 108.6
500 400 300 200 100 0
Distance from end of FRP (mm)

Figure 5.87 Longitudinal CFRP strain distributions in beam A1b

6000

5000 Location
from CFRP
4000
Microstrain

end (mm)
3000 137.5
2000 412.5

1000 550

0
0 20 40 60 80 100
Shear load (kN)

Figure 5.88 Longitudinal CFRP strain development in beam A1b

Leftmost strap Middle strap Rightmost strap


250 250 250

200 200 200


Height (mm)

150 150 150

100 100 100

50 50 50

0 0 0
-500 500 1500 2500 -500 500 1500 2500 3500 -500 500 1500 2500
Microstrain Microstrain Microstrain

Load (kN)
Front 20.0 40.0 59.9 80.0 90.0 100.0 108.7
Back 20.0 40.0 59.9 80.0 90.0 100.0 108.7

Figure 5.89 Strain distributions along U strap height in beam A1b at different load
levels

198
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

5000 Leftmost strap 5000 Middle strap 5000 Rightmost strap


5000 Leftmost strap 5000 Middle strap 5000 Rightmost strap
4000 4000 4000
4000 4000 4000
Microstrain

3000 3000 3000


Microstrain

3000 3000 3000


2000 2000 2000
2000 2000 2000
1000 1000 1000
1000 1000 1000
0 0 0
0 0 20 40 60 80 100 0 0 0 0
20 40 60 80 100 20 40 60 80 100
0 20 40 60 80 100 0 20 40 60 80 100 0 20 40 60 80 100
Shear load (kN) Shear load (kN) Shear load (kN)
Shear load (kN) Shear load (kN) Shear load (kN)

Front Back

Figure 5.90 Strain development in the U-strap legs in beam A1b

5000

4000
Microstrain

3000

2000

1000

0
0 20 40 60 80 100 120
Shear load (kN)

Figure 5.91 Strain development in the tension reinforcement in beam A1b

199
CHAPTER 5

Load point Load (kN)


5000 19.9

4000 40.0
Microstrain 3000
60.0
70.1
2000
80.0
1000 89.9
0 101.7
500 400 300 200 100 0
Distance from end of FRP (mm)

Figure 5.92 Longitudinal CFRP strain distributions in beam A2a

5000

4000 Location from


CFRP end (mm)
Microstrain

3000 137.5
2000 275
412.5
1000
550
0
0 20 40 60 80 100
Shear load (kN)

Figure 5.93 Longitudinal CFRP strain development in beam A2a

5000
Gauge
4000
location
Microstrain

3000 Front
Back
2000

1000

0
0 20 40 60 80 100
Shear load (kN)

Figure 5.94 Strain development in the U-strap legs in beam A2a

200
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

10000

8000

Microstrain
6000

4000

2000

0
0 20 40 60 80 100 120
Shear load (kN)

Figure 5.95 Strain development in the tension reinforcement in beam A2a

201
CHAPTER 5

Load point Load (kN)


6000
19.9
5000 40.0

Microstrain
4000 60.0
3000 80.0
90.1
2000
100.0
1000
109.9
0 114.5
500 400 300 200 100 0
Distance from end of FRP (mm)

Figure 5.96 Longitudinal CFRP strain distributions in beam A2b

6000
5000 Location
from CFRP
Microstrain

4000 end (mm)


3000 137.5
2000 320
1000 550
0
0 20 40 60 80 100 120
Shear load (kN)

Figure 5.97 Longitudinal CFRP strain development in beam A2b

Leftmost strap Middle strap Rightmost strap


250 250 250

200 200 200


Height (mm)

150 150 150

100 100 100

50 50 50

0 0 0
-500 500 1500 2500 -500 500 1500 2500 -500 500 1500 2500 3500
Microstrain Microstrain Microstrain

Load (kN)
Front 20.1 40.0 60.0 80.0 90.0 100.0 109.9 114.5
Back 20.1 40.0 60.0 80.0 90.0 100.0 109.9 114.5

Figure 5.98 Strain distributions along U strap height in beam A2b at different load
levels

202
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

6000 Leftmost strap 6000 Middle strap 6000 Rightmost strap

5000 5000 5000


4000 4000
Microstrain
4000
3000 3000 3000
2000 2000 2000
1000 1000 1000
0 0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120 0 20 40 60 80 100 120
Shear load (kN) Shear load (kN) Shear load (kN)

Front Back

Figure 5.99 Strain development in the U-strap legs in beam A2b

10000

8000
Microstrain

6000

4000

2000

0
0 20 40 60 80 100 120
Shear load (kN)

Figure 5.100 Strain development in the tension reinforcement in beam A2b

Table 5.8 summarises the maximum strains recorded for the U-strap legs and the
failure types. It can be seen from the table that the non-prestressed systems failed by
CFRP rupture for most cases. The average maximum strain was 3270 microstrain, which
was much lower than the tensile rupture strain. This was due to bending of the strips and
therefore non-uniform stress distributions. The prestressed systems failed by CFRP
debonding for all cases with the average debonding strain in the straps of 4540
microstrain. This strain is approximately 90 % of the average debonding strain measured
in the shear-lap tests with the same composite width.

203
CHAPTER 5

Table 5.8 Maximum U-strap leg strains


Leftmost strap Right most
Middle
(near load point) (near CFRP end)
Beam Face
Maximum Failure Maximum Failure Maximum Failure
strain type strain type strain type
Front N/A N/A N/A N/A N/M2 Local1
E3a2
Back N/A N/A N/A N/A 3284 Rupture
Front N/A N/A N/A N/A 3580 Debond
E1a2
Back N/A N/A N/A N/A 4696 None
Front N/M2 None N/M2 None 3302 None
E5a2
Back N/M2 Debond N/M2 Debond 3149 Debond
Front N/A N/A N/A N/A 2994 Local1
A1a
Back N/A N/A N/A N/A 3178 Rupture
Front 3432 Debond 3635 Rupture 2994 Rupture
A1b
Back 4062 Local1 3965 Local1 3178 Local1
Front N/A N/A N/A N/A 4214 None
A2a
Back N/A N/A N/A N/A 4165 Debond
Front 3168 Debond 6067 Debond 3558 None
A2b
Back 2978 None 6448 None 8233 Debond

Notes: 1) Local failure: the strap was not ruptured or completely debonded but some localised debonding
and minor damage were present.
2) Not measured.

5.6 Summary of findings


This chapter has presented the procedures and results of an experimental
investigation into the debonding behaviour of retrofitted RC beams without and with
anchorage. The investigation consisted of two main programs. The first program involved
18 retrofitted beams tested to failure under four-point loading. All of the beams failed by
debonding. The findings from this experimental program are:

• Two failure modes were observed: intermediate span and end debond. They were
the result of the high shear stress level in concrete. The former mode appeared to occur at
a fixed CFRP strain level of approximately 8600 microstrain. The later appeared to occur
when the shear bond stress reached approximately 1 MPa.

204
EXPERIMENTAL INVESTIGATION OF RETROFITTED BEAMS UNDER BENDING

• The performance of CFRP retrofitting was influenced mainly by two important


parameters: the CFRP bond length in the shear span and the ratio of laminate stiffness to
tension reinforcement stiffness. The efficiency of the retrofitted composite increased with
the bond length and decreased with the stiffness ratio.
• The concrete cover and amount of shear reinforcement had an insignificant
influence on debonding.
• Steel clamping provided a good method to avoid end debond. It did not prevent
intermediate span debond but it improved the beam ductility by holding the delaminated
composite to concrete by friction.
• After debonding of the retrofitting material, the beams still had their original
strength (as without bonding FRP).
• Bonding CFRP to a RC beam led to a higher tensile stress in the tension
reinforcement near the support.

The second experimental program involved eight tests of beams in three-point


bending. One test was carried out on the beam without anchorage and seven tests were
done on the beams with either one end U-strap or three U-straps at 180 mm spacing. The
findings from this experimental program are:

• U-straps placed over longitudinal composite ends prevented end debond. In these
cases, the failure mode shifted to intermediate span debond.
• U-straps placed in shear span limited intermediate span debond.
• Prestressed U-strap anchorage system only improved the beam stiffness and ultimate
capacity slightly compared with the system with non-prestressed straps.
• Non-prestressed systems failed by rupture of the side strips at a strain of
approximately 3200 microstrain.
• Prestressed systems failed by debonding of side strips at a strain of approximately
4500 microstrain.

205
SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS

CHAPTER 6 - SMEARED CRACK MODELLING OF


SHEAR-LAP SPECIMENS

6.1 Introduction and scope


Chapters 4 and 5 have presented the testing of the shear-lap specimens and
retrofitted RC beams. In the present study, numerical modelling is utilised to simulate the
testing. The simulations are carried out to predict the peak loads and failure mechanisms
in order to gain further understanding of the debonding behaviour of the shear-lap
specimens and retrofitted beams.

Two methods of finite element modelling are used in the study. The first method
is based on smeared cracking, whereas the second method uses a combination of smeared
and discrete cracking.

In this chapter, the smeared crack modelling of the shear-lap specimens under
testing is presented. The simulation was implemented using a commercially available
non-linear finite element code, DIsplacement method ANAlyser or DIANA (version
8.1.2), developed by TNO Building and Construction Research in the Netherlands (de
Witte and Kikstra, 2003). The chapter reports the finite element idealisation, material
models, solution procedure and verification with the shear-lap test results. A study of the
model sensitivity to mesh size, mesh type, concrete tensile properties and adhesive
stiffness is also included. In Chapter 7, modelling of the beam tests using smeared cracks
is presented, whereas modelling of the shear-lap and beam tests using discrete/smeared
cracks is reported in Chapter 8.

6.2 Finite element idealisation


There have been limited attempts to model the bond test using smeared crack
modelling as reviewed in Chapter 2. In these studies, relatively coarse meshes were used
with bond-slip interface elements inserted between the FRP composite and concrete.
Specific constitutive relations relating the tractions to the relative displacements were
assumed for these interface elements. The major drawback of this method was that those
relations had to be known in advance.

207
CHAPTER 6

A different approach was used in the finite element model presented in this
chapter. Instead of lumping the bond failure into an interface slip, the failure was
simulated as cracking in concrete near the bond surface using a fine mesh near the bond
line. This approach was successfully used to model the bond between steel reinforcement
and concrete and has the potential of explaining the fundamentals of the bond (Rots,
1989).

A typical finite element mesh and boundary conditions of the model are shown in
Figure 6.1. Since a fine mesh was required near the bond line, the concrete block was
modelled using three layers: a fine layer with an element size of approximately 2 mm, a
transition layer with transition elements and a coarse layer with an element size of
approximately 10 mm. The CFRP was connected to the top concrete surface via an
adhesive layer. The adhesive layer was assumed to be 1 mm thick and the CFRP
composite formed by one ply of fibres was taken to be 0.7 mm thick (Figure 4.8b).
Horizontal and vertical roller supports were placed appropriately to simulate the actual
physical restraints. A prescribed displacement was applied on the CFRP end.

Lf Horizontal
Adhesive CFRP
roller

H Fine layer

Y
Transition
layer
X
Coarse
Vertical Horizontal roller layer
roller

Figure 6.1 Two-dimensional mesh of a typical specimen (T3)

The concrete in the fine and transition layers was modelled using six-node
quadrilateral isoparametric plane stress elements (CQ12M) (Figure 6.2a). Each element
has twelve degrees of freedom (dof) with two displacements, ux and uy, at each node. A
three-point Gaussian integration scheme was used. The CFRP, the adhesive and the

208
SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS

concrete in the coarse layer were modelled as eight-node quadrilateral isoparametric


plane stress elements (CQ16M). Each element has sixteen degrees of freedom (Figure
6.2b). The element was numerically integrated using a 2 x 2 Gaussian integration scheme.

5
6 5
7
6 η 4
η
8 4
ξ
1 3 ξ
2
1 3
2
(a) (b)
Figure 6.2 CQ12M element (a) and CQ16M element (b)

6.3 Material models


6.3.1 Concrete
Concrete was modelled using a total strain crack model. The model was
developed along the lines of the Modified Compression Field Theory, initially proposed
by Vecchio and Collins (1986). Detailed descriptions of this model have been reported
elsewhere (Rots et al., 1985; Rots, 1989). The model implementation is summarized
below as adapted from de Witte and Kikstra (2003):

1) The stress is evaluated as a function of the strain in the directions of the cracks
and this function follows an explicit curve as in the following formula:
t + ∆t
i +1 σ nst = σ( t +i +∆1t ε nst ) (6.1)
where nst is the crack coordinate system. The strain and stress in the element
coordinate system are related to the strain and stress in the crack direction by a
strain transformation matrix T, as follows
t + ∆t
t + ∆t
i +1 ε nst =T t +i ∆+1t ε xyz ; i +1 σ nst = T T t +i +∆1t σ nst (6.2)

The strain is updated with the strain increment as


t + ∆t
i +1 ε xyz = t ε xyz + t +i +∆1t∆ε xyz (6.3)

where ∆t is the time increment.


2) The deterioration of the material due to cracking or crushing depends on different
strain levels and is monitored by six internal damaged variables αk (the first three

209
CHAPTER 6

variables are for tension and the last three are for compression). Since no damage
recovery is assumed, to ensure that these internal damaged variables are not
decreasing in case of unloading, the unloading constraints rk are used. They are
active (rk = 1) if unloading occurs and inactive (rk = 0) otherwise (Figure 6.3). The
internal variables are updated as
t + ∆t
i +1 α = t α + (1 − rk )∆ε (6.4)
3) The stress is related to the strain with the following equation:
σ = f (α, ε).g(α, ε nst ) (6.5)
where the reduction due to the loading-unloading effect is taken into account by
the following function:
αk − ε j
g =1− (6.6)
αk

σj

αj+3
εj
αj

rj = 1 rj = 0
rj+3 = 1 rj+3 = 0

Figure 6.3 Loading-unloading for total strain crack models (de Witte and Kikstra,
2003)

4) The tensile behaviour of concrete is described with a softening curve. For the
fixed crack model, the shear behaviour after cracking can also be reduced by a
shear retention factor, β.
5) The stress strain curve for the compressive behaviour of concrete depends on the
effect of lateral confinement (increasing strength and ductility) or lateral cracking
(reducing strength and ductility), which can be modelled using Thorenfeldt’s

210
SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS

curve (Thorenfeldt et al., 1987) and Vecchio and Collin’s reduction factors
(Vecchio and Collin, 1986), respectively.

Two total strain crack models are available in DIANA: a fixed crack model and a
rotating crack model. In the former, the stress-strain relationship is evaluated in a
coordinate system which is fixed upon cracking. For the latter, the coaxial stress-strain
concept is used, in which the stress-strain relationship is evaluated in the principal
directions of the strain vector.

To avoid stress-locking problems associated with the fixed crack model when
changes in crack orientation occur during loading (Rots, 1989), the rotating crack model
was used in this study to model the concrete block.

The input for the model comprises the elastic and nonlinear properties in tension,
shear and compression. The elastic properties consist of the elastic modulus and Poisson’s
ratio. In this study, the elastic modulus was determined using the expression given by
ACI Committee 363 (1992), as follows

E c = 3320 f c' + 6900 (MPa) (6.7)

This expression can be applied to concrete of all strengths (Warner, 1998). The Poisson’s
ratio, νc, was taken to be 0.2.

The behaviour of concrete in tension was described using a nonlinear tension


softening stress-strain relationship proposed by Hordijk (1991). The relationship is as
illustrated in Figure 6.4a. The peak tensile strength of the concrete, fct, was determined
using the expression provided by Comite Euro-International du Beton (1991) as given in
Equation 2.52d. The area under the descending branch of the stress-strain curve is given
by Gf / h, where Gf is the mode-I fracture energy and h is the crack band width. The crack
band width is a discretisation factor, which was determined in DIANA as the square root
of the total area of the element. The fracture energy was calculated using the expression
by Trunk and Wittman (1998) which is given in Equation 2.56.

211
CHAPTER 6

The behaviour of the concrete in compression was modelled using the function
proposed by Thorenfeldt et al. (1987), of which the general shape is illustrated in Figure
6.4b. The function described both the hardening and softening behaviour of concrete.

σ σ
fct ε

Gf/h
fc
ε

(a) (b)
Figure 6.4 Concrete stress-strain curve in tension (a) and compression (b)

A summary of the concrete properties used in the present FEA is provided in


Table 6.1.

Table 6.1 Concrete properties adopted in the finite element model


Property Value
Elastic modulus (MPa) 29400
Poisson’s ratio 0.2
Compressive strength (MPa) 53.7
Tensile strength (MPa) 3.86
Fracture energy (N/mm) 0.188

6.3.2 CFRP composite and adhesive


The CFRP composite and adhesive were assumed to be isotropic linear elastic
materials. The CFRP elastic modulus based on the nominal fibre thickness of 0.176 mm
per ply was taken to be 209000 MPa. The corresponding CFRP elastic modulus based on
the true thickness of 0.7 mm per ply was 53000 MPa. The adhesive elastic modulus was
taken as 8500 MPa. The Poisson’s ratios of the CFRP composite and the adhesive were
assumed to be 0.30.

212
SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS

6.4 Solution procedure


In a nonlinear finite element analysis, the relation between a force vector and
displacement vector is not linear and a solution cannot be found directly. To determine
the state of equilibrium, the load is ‘applied’ incrementally in steps of ∆u. To achieve
equilibrium at the end of each increment, an iterative solution algorithm is used. In the
iterative algorithm, the total displacement increment ∆u is adapted iteratively by an
increment δu until equilibrium is reached. The method, called incremental-iterative
method, was used in the present study. In the models, the load was ‘applied’ in
increments of 0.0025 mm initially. At each load step, the modified Newton-Raphson
iterative scheme was used to bring the internal forces to an acceptable level of
equilibrium. This scheme is illustrated in Figure 6.5, in which the stiffness matrix was
evaluated at the start of the increment. The convergence criterion adopted was based on
the energy norm composed of internal forces and relative displacements given by the
following equation:

δu iT (f int,i +1 + f int,i )
Energy Norm Ratio = (6.8)
∆u T0 (f int,1 + f int,0 )

where δui is the ith iterative displacement increment, ∆u0 is the displacement increment
and fint is the internal force. A new reference norm was determined at the start of each
step. In all of the models, the tolerance for convergence was set to 0.0005 to ensure
accurate and good results. The maximum number of iterations for each load step or
increment was set to 100.

Figure 6.5 Modified Newton-Raphson iteration (de Witte and Kikstra, 2003)

213
CHAPTER 6

To improve the convergence, the line search algorithm was also used in the study.
The algorithm scaled the prediction of the iterative displacement increment, δu, by a
value to minimize the energy potential.

6.5 Verification of finite element model


To verify the validity of the FE model, the crack patterns at failure, the peak
loads, the strain distributions and the bond-slip relations along the CFRP, and the load-
slip curves predicted by the model were compared with the corresponding experimental
measurements.

6.5.1 Comparison of crack patterns


The FEA indicated two failure modes similar to those observed in the
experiments. The failure sequences are illustrated in Figure 6.6 and Figure 6.7 for two
representative cases, T5 and T1, respectively.

Model T5 predicted an interfacial debond failure. As the load was applied, the
tensile force transferred to the concrete near the loaded end. The first clear transverse
shear-tension crack formed normal to the bond at approximately 10 mm from the loaded
end and propagated at an inclined angle into the concrete block. As the load increased, the
second transverse crack formed at 20 mm from the loaded end. When this crack opened,
the stress was released around the first crack and the crack opening was reduced there. As
the load increased further, a debond crack was initiated at the tip of the second normal
crack and propagated parallel to the bond surface. The shear stress transferred along and
subsequent cracking occurred. The transverse cracks formed at an equal interval of
approximately 10 mm with the two dominant cracks located at 20 and 60 mm from the
loaded edge.

Model T1 predicted a shear-tension failure. In this model, as the composite was


relatively short, the stress transferred well towards the unloaded end leading to a shear-
tension crack originated from the cut-off point. The applied load reached its peak when
the crack started to propagate into the concrete block.

214
SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS

24 kN

27 kN
Transverse cracks Debond cracks

28 kN (peak)
Figure 6.6 Deformed shapes of model T5 at different load levels (magnification
factor = 60)

215
CHAPTER 6

10 kN

15 kN
Shear-tension crack

18 kN (peak)
Figure 6.7 Deformed shapes of model T1 at different load levels (magnification
factor = 60)

The crack patterns at the peak load were illustrated in Figure 6.8 to Figure 6.11 for
all specimens. For clarity, only the cracks, of which the crack strain is large enough so
that the crack tensile stress has reduced by 50 % from the maximum tensile strength, are
shown. The predicted peak loads and failure modes are compared in Table 6.2

It can also be seen that for some specimens the number of cracks, the pulled-out
concrete prism sizes and the failure modes predicted by the FEA were different to those
observed in the experiments. A possible reason for the discrepancy was that in the
experiment, stress transferring occurred in a more brittle manner. The propagation, which
led to total debonding of the CFRP composite, might occur from the tip of a transverse
crack closer to the loaded edge as a result of the dynamic effect from a sudden release of
energy once the crack opened. It was not captured with the static FE model.

216
SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS

(T1) (T2)

(T3) (T4)

(T5) (T6)
Figure 6.8 Crack patterns in models T1 to T6

(T7) (T8)
Figure 6.9 Crack patterns in models T7 and T8

(T9) (T10)

(T11)
Figure 6.10 Crack patterns in models T9, T10 and T11

(T12) (T13)
Figure 6.11 Crack patterns in models T12 and T13

217
CHAPTER 6

Table 6.2 Comparison between the experimental and the FEA results
Experiment Finite element analysis
Spec- Var- Peak Vari- Peak Devia-
imen iable load ation1 Failure mode load tion2 Failure mode
(kN) (%) (kN) (%)
T1a 20.0 Shear-tension 18.3 8.3 Shear-tension
6.2
T1b 18.8 Interfacial debond 18.3 2.5 Shear-tension
T2a 25.8 Interfacial debond 22.1 14.3 Shear-tension
2.4
T2b 25.2 Interfacial debond 22.1 12.2 Shear-tension
Shear-tension/
T3a 25.8 Interfacial debond 26.9 4.1
Interfacial debond
5.6
Shear-tension/
T3b Lf 27.3 Interfacial debond 26.9 1.6
Interfacial debond
T4a 26.7 Interfacial debond 28.7 14.0 Interfacial debond
3.0
T4b 25.9 Interfacial debond 28.7 17.5 Interfacial debond
T5a 27.8 Interfacial debond 29.4 5.9 Interfacial debond
13.1
T5b 31.7 Interfacial debond 29.4 7.1 Interfacial debond
T6a 31.7 Interfacial debond 29.4 7.3 Interfacial debond
10.3
T6b 28.6 Interfacial debond 29.4 2.8 Interfacial debond
T7a 33.0 Interfacial debond 26.7 19.2 Interfacial debond
20.4
T7b 26.9 Interfacial debond 26.7 0.9 Interfacial debond
H
T8a 28.5 Interfacial debond 27.3 4.1 Shear-tension
4.5
T8b 29.8 Interfacial debond 27.3 8.3 Shear-tension
T9a 28.4 Shear-tension 27.3 4.0 Shear-tension
4.8
T9b 29.8 Shear-tension 27.3 8.5 Shear-tension
T10a 37.4 Interfacial debond 36.7 1.9 Shear-tension
11.6
T10b 33.3 Shear-tension 36.7 10.2 Shear-tension
tf
Shear-tension/
T11a 42.8 Interfacial debond 46.0 7.5
Interfacial debond
9.3
Shear-tension/
T11b 39.0 Shear-tension 46.0 18.0
Interfacial debond
T12a 18.8 Interfacial debond 20.5 9.2 Interfacial debond
2.6
T12b 19.3 Interfacial debond 20.5 6.4 Interfacial debond
bf
T13a 21.2 Interfacial debond 24.0 13.1 Interfacial debond
13.2
T13b 24.2 Interfacial debond 24.0 0.9 Interfacial debond

Notes: 1) Variation = |Fa – Fb| / (Fa/2+Fb/2), where Fa and Fb are the peak loads for specimens ‘a’ and ‘b’,
respectively.
2) Deviation = |Fexp-FFE| / Fexp, where Fexp and FFE are the experimental and predicted load
capacities of the bond.

218
SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS

6.5.2 Comparison of peak loads and behavioural trends


The correlation of the tensile load capacities as calculated by the FEA and as
measured in the experiments for all specimens is illustrated in Figure 6.12. The deviations
from the experimental values ranged from 0.9 to 19.2 %. The predictions are reasonably
accurate given the fact that experimental variations were from 2.4 % to 20.4 %. The
average of the ratio of the predicted to measured capacity Ppredicted/Pexp was 1.00 with the
coefficient of variation of 9.2 %.

50 Varying H tf bf
Lf
45 parameter
40
Maximum load (kN

35
30
25
20
15
10 Experiment FEA
5
0

T10a

T11a

T12a

T13a
T10b

T11b

T12b

T13b
T1a

T2a

T3a

T4a

T5a

T6a

T7a

T8a

T9a
T1b

T2b

T3b

T4b

T5b

T6b

T7b

T8b

T9b

(a)
50

40
Predicted max. load (kN

30

20

10

0
0 10 20 30 40 50
Measured max. load (kN)

(b)
Figure 6.12 Correlation of peak loads predicted by the FE model and measured in
the experiment

The behavioural trends of the specimens with different CFRP bond lengths,
support block clearances, CFRP thicknesses and CFRP widths as observed in the
experiments have been previously presented in section 4.4.3. The trends are depicted

219
CHAPTER 6

again in Figure 6.13 to Figure 6.16 where the predictions by the FEA are also included. It
is clear that the trends are captured well by the FE model. For the specimens bonded with
2 plies of CFRP, the effective bond length is found to be 100 mm. The effective bond
stress decreases almost linearly with the CFRP width. It is also found that the support
clearance has an insignificant effect on the bond capacity.

Experiment 2 plies FEA 2 plies Experiment 2 plies FEA 2 plies


Experiment 6 plies FEA 6 plies Experiment 6 plies FEA 6 plies

Average bond stress (MPa)


50 3.5
3.0
40
2.5
Load (kN)

30 2.0
20 1.5
1.0
10 0.5
0 0.0
0 50 100 150 200 250 0 50 100 150 200 250
Lf (mm) Lf (mm)

Figure 6.13 Effect of varying CFRP bond length (Experiment and FE)

Experiment FEA
35
30
25
Load (kN)

20
15
10
5
0
0 25 50 75 100 125
H (mm)
Figure 6.14 Effect of varying support block height H (Experiment and FE)

220
SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS

Lf = 100 mm Experiment FEA Lf = 100 mm Experiment FEA


Lf = 140 mm Experiment FEA Lf = 140 mm Experiment FEA
Lf = 180 mm Experiment FEA Lf = 180 mm Experiment FEA
50 1020
45

CFRP stress (MPa)


820
Load (kN)

40
620
35
30 420
25 220
20
20
0 0.5 1 1.5
0 0.5 1 1.5
tf (mm) tf (mm)

Figure 6.15 Effect of varying CFRP thickness (Experiment and FE)

Experiment FEA Experiment FEA

Average bond stress (MPa)


30 5
25 4
Load (kN)

20 3
15
2
10
5 1
0 0
20 40 60 80 100 120 20 40 60 80 100 120
b f (mm) b f (mm)

Figure 6.16 Effect of varying CFRP width (Experiment and FE)

6.5.3 Comparison of strain distributions and bond-slip curves


The strain distributions along the bonded joint are compared in Figure 6.17 to
Figure 6.21. The comparison was made at several load levels. The experimental and
numerical load levels were selected to be as close as possible to each other. Since the
strain gauges used in the experiments measured the average strain over a distance, which
was the gauge length of 6 mm, the numerical strain readings were taken to be the average
strain over three elements of 2 mm in size. The element strains were calculated at the
Gauss points.

In general, the numerical models simulated the strain distributions well. There are
exceptions for specimens T9, T10 and T11, for which, as discussed in Chapter 4, the
experimental strain readings were not reliable. Comparable trends can be observed in the
distribution plots for most specimens. At low load levels, the predicted strain distributions

221
CHAPTER 6

exhibit an exponential trend with the highest strain near the loaded end. As the load
increases, the strain distribution curves displace further along the joint similar to the
observation in the experiments. This phenomenon is most clearly seen in the specimens
with a long bond length, for example specimens T5a and T6a. When the failure mode is
shear-tension, the shifting of the high stress region is not observed clearly because once
the end crack opens, the CFRP strain level drops and the peak load is reached. This is
typically observed in the specimens with a very short bond length or with a very thick
CFRP layer, for example specimens T1a, T9a and T10a.

The strain readings were utilised to plot load-slip curves for all specimens in the
same way as for the experiments. They are compared in Figure 6.22 and Figure 6.24. The
load-slip plots from the experiments and the numerical models are comparable. For most
cases, the FE results show slightly more brittle behaviour. As expected, the initial
stiffness of the specimens with different bond lengths is similar but higher peak loads and
more ductile behaviour are observed in those with longer bond lengths.

222
SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS

4000 Exp. (kN) FEA (kN)


3500 10.0 9.8
3000 15.0 15.1

Microstrain
2500 18.0 18.0
T1a 2000
1500
1000
500
0
0 50 100 150 200
Distance from loaded edge (mm)

4000 Exp. (kN) FEA (kN)


3500 10.0 9.8
15.0 15.2
3000
20.0 20.1
Microstrain

2500 22.0 22.1


T2a 2000
1500
1000
500
0
0 50 100 150 200
Distance from loaded edge (mm)

3500 Exp. (kN) FEA (kN)


10.0 9.9
3000
15.0 15.2
2500 20.0 20.2
Microstrain

2000 25.7 25.9


T3a 1500
1000
500
0
0 50 100 150 200
Distance from loaded edge (mm)

Figure 6.17 Comparison of CFRP strain distributions in specimens T1a, T2a and
T3a

223
CHAPTER 6

5000 Exp. (kN) FEA (kN)


10.0 10.0
15.0 14.8
4000
Microstrain 20.0 20.1
24.0 24.0
3000 26.0 26.0

T4a 2000

1000

0
0 50 100 150 200
Distance from loaded edge (mm)

5000
10.0 9.9
4000 15.0 15.3
20.0 20.2
Microstrain

3000 26.0 26.0


27.8 27.4
T5a 2000

1000

0
0 50 100 150 200
Distance from loaded edge (mm)

5000 Exp. (kN) FEA (kN)


10.0 9.9
4000 15.0 14.8
20.0 19.9
Microstrain

24.0 24.1
3000 26.0 26.1
28.0 28.0
2000 26.9 26.9
T6a
1000

0
0 50 100 150 200
Distance from loaded edge (mm)

Figure 6.18 Comparison of CFRP strain distributions in specimens T4a, T5a and
T6a

224
SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS

3500 Exp. (kN) FEA (kN)

3000 10.0 10.7


15.0 15.5
2500 20.0 19.9

Microstrain
2000 25.0 25.0
1500
T7a
1000
500
0
0 50 100 150 200
Distance from loaded edge (mm)

3500 Exp. (kN) FEA (kN)


3000 10.0 9.7
15.0 15.1
2500 20.0 20.2
Microstrain

25.0 25.0
2000 26.0 27.1
1500
T8a
1000
500
0
0 50 100 150 200
Distance from loaded edge (mm)

Figure 6.19 Comparison of CFRP strain distributions in specimens T7a and T8a

225
CHAPTER 6

1500 Exp. (kN) FEA (kN)


10.0 10.0
1000 15.0 15.4
Microstrain
500 20.0 20.1
25.0 27.3
0
T9a
0 50 100 150 200
-500

-1000

-1500
Distance from loaded edge (mm)

1600 Exp. (kN) FEA (kN)


1400 10.0 10.2
1200 20.0 20.4
Microstrain

1000 25.0 25.3


30.0 30.3
800
35.0 36.7
T10a 600
400
200
0
0 50 100 150 200
Distance from loaded edge (mm)

2000 Exp. (kN) FEA (kN)


10.0 10.2
1500 20.0 19.7
Microstrain

30.0 29.9
40.0 40.0
1000
42.8 46.0
T11a
500

0
0 50 100 150 200
Distance from loaded edge (mm)

Figure 6.20 Comparison of CFRP strain distributions in specimens T9a, T10a and
T11a

226
SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS

5000 Exp. (kN) FEA (kN)


10.0 10.1
4000 14.0 14.2
16.0 16.0

Microstrain
3000 18.0 18.3

T12a 2000

1000

0
0 50 100 150 200
Distance from loaded edge (mm)

4000 Exp. (kN) FEA (kN)


3500 10.0 10.0
15.0 15.1
3000 18.0 18.0
Microstrain

2500 20.0 21.0


20.1 20.6
2000
T13a 1500
1000
500
0
0 50 100 150 200
Distance from loaded edge (mm)

Figure 6.21 Comparison of CFRP strain distributions in specimens T12a and T13a

227
CHAPTER 6

25 30

Total load (kN)


Total load (kN)

20 25
20
15 T1a (Exp.) T2a (Exp.)
15
10 10
T1 (FE) T2 (FE)
5 5
0 0
0.00 0.01 0.02 0.03 0.00 0.02 0.04 0.06 0.08
Slip at 42 mm from loaded edge Slip at 42 mm from loaded edge

T1a T2a
30 40

Total load (kN)


Total load (kN)

25
30
20
15 T3a (Exp.) 20
T4a (Exp.)
10 T3 (FE) 10 T4 (FE)
5
0 0
0.00 0.05 0.10 0.15 0.20 0.00 0.10 0.20 0.30
Slip at 42 mm from loaded edge Slip at 42 mm from loaded edge

T3a T4a
40 40
Total load (kN)

Total load (kN)

30 30

20 T5a (Exp.) 20
T6a (Exp.)
10 T5(FE) 10
T6 (FE)
0 0
0.00 0.20 0.40 0.60 0.00 0.10 0.20 0.30 0.40
Slip at 42 mm from loaded edge Slip at 42 mm from loaded edge

T5a T6a

Figure 6.22 Comparison of load-slip curves for specimens T1a to T6a

228
SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS

30 40
25

Total load (kN)


Total load (kN)
30
20
15 20
T9a (Exp.) T10a (Exp.)
10
T9 (FE) 10 T10 (FE)
5
0 0
-0.01 0.00 0.01 0.02 0.03 0.00 0.02 0.04 0.06
Slip at 42 mm from loaded edge Slip at 42 mm from loaded edge

T9a T10a
50
40
Total load (kN)

30
20 T11a

10 T11 (FE)
0
0.00 0.05 0.10 0.15
Slip at 42 mm from loaded edge

T11a
Figure 6.23 Comparison of load-slip curves for specimens T9a, T10a and T11a

40 30
Total load (kN)

Total load (kN)

25
30
20
20 T7a (Exp.) 15 T8a (Exp.)
10
10 T7 (FE) T8 (FE)
5
0 0
0.00 0.05 0.10 0.15 0.20 0.00 0.05 0.10 0.15 0.20
Slip at 42 mm from loaded edge Slip at 42 mm from loaded edge

T7a T8a
25 30
Total load (kN)
Total load (kN)

20 25
20
15
T12a (Exp.) 15
10 T13a (Exp.)
10
5 T12 (FE) 5 T13 (FE)
0 0
0.00 0.05 0.10 0.15 0.20 0.25 0.00 0.05 0.10 0.15 0.20
Slip at 42 mm from loaded edge Slip at 42 mm from loaded edge

T12a T13a
Figure 6.24 Comparison of load-slip curves for specimens T7a, T8a, T12a and T13a

229
CHAPTER 6

The local bond-slip curves are compared in Figure 6.25 and Figure 6.26 for two
specimens with a 100 mm wide CFRP composite bonded over a relatively long distance
(180 and 220 mm). The bond-slip curves for two specimens bonded with narrower
composites (50 and 70 mm wide), are also plotted in Figure 6.27 and Figure 6.28. The
plotted stresses and slips are the average over the distance between strain gauges and
calculated in the same way as in the experiments. In the plots, the locations are measured
from the loaded edge.

The bond-slip curves vary significantly from one location to the other along the
CFRP length. This was also observed in the experiments. The main reason for the
variation is due to the influence from the transverse cracks. The bond-slip curves at
locations further from the loaded edge appear to have stiffer behaviour because they are
less influenced by transverse cracks formed near the edge. It is not possible to identify a
curve which would describe the bond behaviour accurately for all specimens at any
position along the joint. However, generally, the bond-slip curves have a non-linear
ascending and descending portion. The maximum bond stress is reached at a slip of
approximately 0.05 mm. The values for the maximum bond stress are approximately 4
MPa for specimens with CFRP width of 100 mm, 5 MPa for specimens with CFRP width
of 70 mm and 6 MPa for specimens with CFRP width of 50 mm. The maximum slip is
approximately 0.2 mm. These values are close to those observed in the experiments.

7 Location (mm) 7 Location (mm)


24 42 24 42
6 6
60 78
Bond stress (MPa)

Bond stress (MPa)

5 60 78 5
96 96
4 4
3 3
2 2
1 1
0 0
0.00 0.05 0.10 0.15 0.20 0.25 0.00 0.05 0.10 0.15 0.20 0.25
Slip (mm) Slip (mm)

(a) FE (b) Experiment


Figure 6.25 Comparison of local bond-slip relationships in specimen T5a (bf = 100
mm, Lf = 180 mm)

230
SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS

6 Location (mm) 6 Location (mm)


24 42 24 42
5 5

Bond stress (MPa)

Bond stress (MPa)


60 78 60 78
4 4
96 96
3 3

2 2

1 1

0 0
0.00 0.05 0.10 0.15 0.20 0.25 0.00 0.05 0.10 0.15 0.20 0.25
Slip (mm) Slip (mm)

(a) FE (b) Experiment


Figure 6.26 Comparison of local bond-slip relationships in specimen T6a (bf = 100
mm, Lf = 220 mm)

12 Location (mm) 9 Location (mm)

24 42 8
10 24 42
7
Bond stress (MPa)

Bond stress (MPa)

8 60 78 6
60 78
5
6
4
4 3
2
2
1
0 0
0.00 0.05 0.10 0.15 0.20 0.25 0.00 0.05 0.10 0.15 0.20 0.25
Slip (mm) Slip (mm)

(a) FE (b) Experiment


Figure 6.27 Comparison of local bond-slip relationships in specimen T12a (bf = 50
mm, Lf = 100 mm)

231
CHAPTER 6

9 Location (mm) 9 Location (mm)


8 8
24 42 24 42
7 7
Bond stress (MPa)

Bond stress (MPa)


6 60 78 6
60 78
5 5
4 4
3 3
2 2
1 1
0 0
0.00 0.05 0.10 0.15 0.20 0.25 0.00 0.05 0.10 0.15 0.20 0.25
Slip (mm) Slip (mm)

(a) (b)
Figure 6.28 Comparison of local bond-slip relationships in specimen T13a (bf = 70
mm, Lf = 180 mm)

6.6 Local strain and stress distributions


The above FE model produced comparable results to those of the experiments in
terms of the failure mechanisms, the peak loads, the CFRP strain distributions and the
load-slip curves. The CFRP strains considered were the average strains over the gauge
length. These average strains are adequate to describe the global behaviour of the joint.
However, as pointed out before by Chen et al. (2001), the strain varies not only along the
composite plate and the adhesive but also across their sections. It is very difficult to
measure these variations. They can however be studied using the finite element model.
The following paragraphs are devoted to the prediction of the local CFRP strain
distributions using the model described above.

The analysis is an approximation of the actual situation and there are limitations
inherent in the model. Firstly, the composite itself consists of a number of different
layers. Its prosperities could depend on the thickness and the properties of the layers, and
the bond between the layers. However, in the analysis, the composite was assumed to be
homogenous and have uniform properties across its cross section. Secondly, the spew
fillets at the composite ends formed from excess adhesive were not avoidable during
preparation of the specimens. These fillets could affect the local stress distributions near
the CFRP end. Nevertheless, the spew fillets were not considered in the analysis. Thirdly,
there are stress singularities at two-material wedges in the problem. The stress will never

232
SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS

converge at the singular points on the bond lines near the composite ends and therefore
the stress magnitude near the composite ends might be sensitive to the mesh size. In this
analysis, the mesh size of 2 mm was used. It will be proved later (in section 6.7.1) that
this size is adequate to estimate the stresses near the composite ends.

It is also worth noting that, in this analysis, the stress and strain outputs were
taken not at the nodes but rather at the Gaussian integration points in the element (Figure
6.29). As a result, the finite element outputs were only the indication of the stress/strain
levels in the upper and lower halves of the layer.

Points representing
the top half
CFRP elements
Points representing
the bottom half
Adhesive elements

Figure 6.29 Gaussian integration points in the CFRP and adhesive elements

Figure 6.30 shows the CFRP strain distributions in the top and bottom halves of
the composite at three load levels for specimen T5. This specimen was bonded with two
plies of CFRP over a length of 180 mm. It can be seen clearly that at the loaded end, the
CFRP composite was bended and the strain in the top half is much lower than that in the
bottom half. The difference diminishes further away from the loaded edge. At or below
service load levels, the difference is insignificant beyond a distance of 10 mm from the
loaded edge. At the unloaded end, similar bending action also existed but the difference is
less noticeable than that at the loaded edge. It diminishes beyond a distance of around
7 mm from the end. Near the ultimate load, the influence of the major normal cracks on
the CFRP strain distribution can also be seen. Over the tip of these cracks, the strain on
the CFRP top reduces from a high crest to a low trough.

233
CHAPTER 6

6000 Load (kN)


Major normal 10.4(Top)
5000 crack location 10.4(Bottom)

CFRP microstrain
4000 20.2(Top)
20.2(Bottom)
(a) The whole 3000 29.4(Top)
29.4(Bottom)
CFRP strain 2000
1000
distribution
0
-1000 0 50 100 150 200

Distance from loaded end (mm)

6000 Load (kN)


10.4(Top)
5000
10.4(Bottom)
CFRP microstrain

4000 20.2(Top)
20.2(Bottom)
(b) CFRP strain 3000 29.4(Top)
29.4(Bottom)
distributions near 2000
1000
the loaded end
0
-1000 0 10 20 30 40

Distance from loaded end (mm)

Load (kN)
200
10.4(Top)
10.4(Bottom)
150
CFRP microstrain

20.2(Top)
20.2(Bottom)
(b) CFRP strain 100 29.4(Top)
29.4(Bottom)
distributions near 50
the un-loaded end
0
160 165 170 175 180
-50
Distance from loaded end (mm)

Figure 6.30 CFRP strain distributions in model T5

The CFRP strain distributions in the specimen T11 bonded with six plies of CFRP
over the same bond length of 180 mm are shown in Figure 6.31. The distributions are
similar to those of model T5. However, the lengths, over which significant differences
between the top and bottom layers exit, are longer compared to those in model T5.

234
SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS

3500 Load (kN)


10.2(Top layer)
3000 Major normal 10.2(Bottom layer)
crack location 23.4(Top layer)

CFRP microstrain
2500 23.4(Bottom layer)
2000 46.0(Top layer)
(a) The whole 46.0(Bottom layer)
1500
CFRP strain 1000
distribution 500
0
-500 0 50 100 150 200
Distance from loaded end (mm)

3500 Load (kN)


10.2(Top layer)
3000 10.2(Bottom layer)
23.4(Top layer)
2500
CFRP microstrain

23.4(Bottom layer)
2000 46.0(Top layer)
(b) CFRP strain 46.0(Bottom layer)
1500
distributions near 1000
the loaded end 500
0
-500 0 10 20 30 40
Distance from loaded end (mm)

200 Load (kN)


10.2(Top layer)
10.2(Bottom layer)
150 23.4(Top layer)
CFRP microstrain

23.4(Bottom layer)
46.0(Top layer)
(b) CFRP strain 100 46.0(Bottom layer)

distributions near 50
the un-loaded end
0
160 165 170 175 180
-50
Distance from loaded end (mm)

Figure 6.31 CFRP strain distributions in model T11

235
CHAPTER 6

6.7 Sensitivity study


The FE models described above produced comparable results to the experiments.
The models can be stated to be valid. However, the mesh and material models were
selected from a range of possible options so that the best fitted results could be found.
Therefore, in the sections to follow, a parametric study is carried out to investigate the
sensitivity of the numerical results to a number of parameters. The comparison is done for
two specimens, T5 and T1, only since they experienced two typical failure modes:
interfacial debond and shear-tension failure, respectively.

6.7.1 Sensitivity to mesh size


6.7.1.1 Introduction
To study the model sensitivity to mesh sizes, three meshes were generated and
they are listed in Table 6.3.
Table 6.3 Mesh size variations
Model Mesh size (mm)
Mesh5 5
Mesh2 2
Mesh2 1

6.7.1.2 Comparison
The predictions of the top and bottom halves of the CFRP composite near the
loaded edge are illustrated in Figure 6.32. Due to the singularity near the edge, the
smaller the mesh size, the more accurate prediction can be obtained. It can be seen that
the differences between the strain distributions near the loaded edge are not significant
once the mesh size is equal to or lower than 2 mm. The predictions of the CFRP strain are
however very close beyond a distance of 10 mm from the loaded edge.

236
SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS

Top - Mesh5 Top - Mesh5


160 2500
Bottom - Mesh5 Bottom - Mesh5
140 Top - Mesh2
Top - Mesh2 2000
120 Bottom - Mesh2

CFRP microstrain
CFRP microstrain
Bottom - Mesh2
100 Top - Mesh1 1500 Top - Mesh1
Bottom - Mesh1
80 Bottom - Mesh1
1000
60
40 500
20
0
0
0 5 10 15 20
-20 0 5 10 15 20 -500

Distance from loaded end (mm) Distance from loaded end (mm)

(a) In linear elastic range (b) At half of the ultimate load

6000
5000
Top - Mesh5
CFRP microstrain

4000 Bottom - Mesh5


3000 Top - Mesh2
Bottom - Mesh2
2000
Top - Mesh1
1000 Bottom - Mesh1
0
-1000 0 5 10 15 20

Distance from loaded end (mm)

(c) At ultimate
Figure 6.32 Prediction of CFRP strains near the loaded end at different load levels
using different mesh sizes

The crack patterns after peak in the models of T5 and T1 are compared in Figure
6.33. All models of T5 indicate interfacial debond failure mode. While the models Mesh2
and Mesh1 show similar crack patterns, model Mesh5 predicts debonding initiated from
an inclined crack closer to the loaded edge. The crack patterns are very similar for the
models of T1.

The load-slip curves are compared in Figure 6.34. There is little difference
between the curves.

237
CHAPTER 6

Mesh5

Mesh2

Mesh1

Figure 6.33 Comparison of crack patterns after peak in specimens T5 (left) and T1
(right) predicted by the models with different mesh sizes

T5a (Exp.) T1a (Exp.)


T5 (Mesh5) T1 (Mesh5)
T5 (Mesh2) T1 (Mesh2)
35 T5 (Mesh1) 25 T1 (Mesh1)
Total load (kN)

30 20
Total load (kN)

25 15
20
15 10
10 5
5
0
0
0.00 0.01 0.02 0.03
0.00 0.10 0.20 0.30 0.40 0.50
Slip at 42 mm from loaded edge
Slip at 42 mm from loaded edge

T5 T1
Figure 6.34 Comparison of load-slip curves predicted by the models with different
mesh sizes

238
SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS

6.7.2 Sensitivity to mesh type


6.7.2.1 Introduction
The mesh used previously was triangle elements generated using Delaunay
algorithm (de Witte and Kikstra, 2003). It was to produce a mesh with no bias in mesh
orientation. Several other mesh types can also be used. They are listed in Table 6.4 and
illustrated in Figure 6.35. In the models, all material properties remained the same as the
base model which is labelled as ‘D6’ in Table 6.4.

Table 6.4 Mesh type variations


Model Model description
D6 6-node elements generated by Delaunay algorithm
M6 6-node elements generated by mapped algorithm
M8 8-node elements generated by mapped algorithm

(a) D6 (b) M6 (c) M8


Figure 6.35 Different meshes used

6.7.2.2 Comparison
The crack patterns after peak in the models of T5 and T1 are compared in Figure
6.36. All models of T5 show two major inclined cracks at around 20 and 60 mm from the
loaded edge. There is a slight difference in the inclined crack orientation between model
D6 and the two models M6 and M8. The interfacial cracks appear to propagate more
easily in the last two models due to their more regular mesh orientation. In the models of
T1, the crack patterns are very similar, showing a shear-tension failure.

239
CHAPTER 6

The load-slip curves are compared in Figure 6.37. The models with a triangular
mesh predict slightly higher ultimate load capacities than the ones with a rectangular
mesh.

(D6)

(M6)

(M8)

Figure 6.36 Comparison of crack patterns after peak in specimens T5 (left) and T1
(right) predicted by the models with different mesh types

T5a (Exp.) T1a (Exp.)


T5 (D6) T1 (D6)
T5 (M6) T1 (M6)
40 T5 (M8) 25 T1 (M8)
Total load (kN)
Total load (kN)

20
30
15
20
10
10 5
0 0
0.00 0.10 0.20 0.30 0.40 0.50 0.00 0.01 0.02 0.03
Slip at 42 mm from loaded edge Slip at 42 mm from loaded edge

(a) T5 (b) T1
Figure 6.37 Comparison of load-slip curves predicted by the models with different
mesh types

240
SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS

6.7.3 Sensitivity to concrete tensile strength and fracture energy


6.7.3.1 Introduction
The bond behaviour is strongly dependent on concrete tensile properties, namely
concrete tensile strength and fracture energy. As discussed in section 2.4.1, different
values of these properties are recommended by different researchers. The values used in
the based model were selected such that the best fitted results can be found. To study the
sensitivity of the finite element model to concrete tensile properties, different values of
the tensile strength , fct, and the fracture energy, Gf, were adopted. They are listed in
Table 6.5, where the base model is labelled as ‘F3.86-G0.188’.

Table 6.5 Concrete tensile property variations


Tensile strength fct Fracture energy Gf
Designation (MPa) (N/mm)
Source Value Source Value
F3.86 – G0.188 CEB1 3.86 Trunk4 0.188
F3.86 – G0.124 CEB1 3.86 Bazant5 0.124
F3.86 – G0.091 CEB1 3.86 CEB1 0.091
F4.40 – G0.188 AS-F2 4.40 Trunk4 0.188
F2.93 – G0.188 AS-D3 2.93 Trunk4 0.188

Notes: 1) CEB-FIB Model Code 1990


2) AS3600, flexural tensile strength
3) AS3600, direct tensile strength
4) Trunk and Wittmann (1998)
5) Bazant and Becq-Giraudon (2002)

6.7.3.2 Comparison
The crack patterns after peak for the specimens with different Gf values are
compared in Figure 6.38. The load-slip curves are compared in Figure 6.39. It is found
that Gf has a significant influence in the behaviour of the shear-lap tests. In T5 models,
the location, at which the debond crack starts to propagate, shifts closer to the loaded
edge as Gf reduces. This is due to the fact that with a lower value of fracture energy, the
interfacial debond crack propagates more easily and the transverse cracks can open wider.
As a result, the stress does not transfer further and the predicted ultimate load drops. In
T1 models, there is little difference in the crack patterns. The predicted ultimate load also
reduces as the fracture energy value decreases.

241
CHAPTER 6

The crack patterns after peak for specimens with different fct values are compared
in Figure 6.40. The load-slip curves are compared in Figure 6.41. The crack patterns in
T5 models indicate that with a lower tensile strength, the transverse cracks tend to be
more inclined towards the loaded edge. The transverse cracks also form at lower load
levels and therefore the load-slip curves have lower stiffness. The predicted ultimate
loads are however only slightly different. For T1 models, the crack propagation is very
rapid and no stress redistribution is possible. Therefore, the predicted ultimate load drops
as fct decreases.

(F3.86 –
G0.188)

(F3.86 –
G0.124)

(F3.86 –
G0.091)

Figure 6.38 Comparison of crack patterns after peak in specimens T5 (left) and T1
(right) predicted by the models with different Gf values

T5a (Exp.) T1a (Exp.)


T5 (F3.86-G0.188) T1 (F3.86-G0.188)
T5 (F3.86-G0.124) T1 (F3.86-G0.124)
35 T5 (F3.86-G0.091) 25 T1 (F3.86-G0.091)
Total load (kN)

30
Total load (kN)

20
25
15
20
15 10
10 5
5
0 0
0.00 0.10 0.20 0.30 0.40 0.50 0.00 0.01 0.02 0.03
Slip at 42 mm from loaded edge Slip at 42 mm from loaded edge

(a) T5 (b) T1
Figure 6.39 Comparison of load-slip curves predicted by the models with different
Gf values

242
SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS

(F4.40 –
G0.188)

(F3.86 –
G0.188)

(F2.93 –
G0.188)

Figure 6.40 Comparison of crack patterns after peak in specimens T5 (left) and T1
(right) predicted by the models with different fct values

T5a (Exp.) T1a (Exp.)


T5 (F4.40-G0.188) T1 (F4.40-G0.188)
T5 (F3.86-G0.188) T1 (F3.86-G0.188)
T5 (F2.93-G0.188) 25 T1 (F2.93-G0.188)
40
Total load (kN) 20
Total load (kN)

30
15
20 10
10 5
0
0
0.00 0.01 0.02 0.03
0.00 0.10 0.20 0.30 0.40 0.50
Slip at 42 mm from loaded edge
Slip at 42 mm from loaded edge

(a) T5 (b) T1
Figure 6.41 Comparison of load-slip curves predicted by the models with different
fct values

243
CHAPTER 6

6.7.4 Sensitivity to adhesive stiffness


6.7.4.1 Introduction
In this section, the effects of varying the adhesive stiffness on the crack patterns
and the load-slip behaviour are investigated. The adhesive layer is defined as the saturant
layer directly above the concrete surface. The stiffness of this layer, Ea, can influence the
stress transferring between the composite and concrete as well as the opening of the
inclined cracks. To study the sensitivity of the finite element model, three different values
of Ea were adopted. They are listed in Table 6.6.

Table 6.6 Adhesive modulus variations


Designation Adhesive elastic modulus Ea (MPa)
A12800 12800
A8500 8500
A3500 3500

6.7.4.2 Comparison
The crack patterns after peak for specimens with different Ea values are compared
in Figure 6.42. The load-slip curves are compared in Figure 6.43. There is little difference
in the behaviour of T1 models except the fact that the load-slip curve stiffness is slightly
higher with a higher adhesive modulus. For T5 models, the models with the adhesive
stiffness of 8500 and 12800 MPa behave very similarly in terms of crack patterns. In
model A3500, the major inclined crack, from which the interfacial debond crack initiates,
shifts slightly closer to the loaded edge. The ultimate loads predicted by all models are
however similar even through model A3500 predicts slightly more ductile behaviour.

244
SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS

(A12800)

(A8500)

(A3500)

Figure 6.42 Comparison of crack patterns after peak in specimens T5 (left) and T1
(right) predicted by the models with different Ea values

T5a (Exp.) T1a (Exp.)


T5 (A12800) T1 (A12800)
T5 (A8500) T1 (A8500)
T5 (A3500) 25 T1 (A3500)
40
Total load (kN) 20
Total load (kN)

30
15
20 10
10 5

0 0
0.00 0.10 0.20 0.30 0.40 0.50 0.00 0.01 0.02 0.03
Slip at 42 mm from loaded edge Slip at 42 mm from loaded edge

(a) T5 (b) T1
Figure 6.43 Comparison of load-slip curves predicted by the models with different
Ea values

6.8 Bond-slip based model


6.8.1 Finite element idealisation
The FE analysis described above was able to simulate both the interfacial debond
and shear-tension failure. The analysis confirmed the finding from the experimental study
that even though the bond slip curves can vary significantly due to the transverse cracks,
they tend to follow a nonlinear curve with a steep ascent followed by a gradual descent,
which can be described by a nonlinear curve. This finding is applicable for the specimens
with two plies of CFRP fabrics, in which the failure mode is dominantly interfacial
debond. Therefore, despite the complexity of the failure mechanism, the global behaviour
of a bond between CFRP and concrete might be described using a simple bond-slip based
model.

245
CHAPTER 6

In this section, an approximate FE analysis of the bond between the CFRP and
concrete is presented. A relatively coarse mesh was used with the bond behaviour
idealised by a nonlinear bond-slip relationship. The analysis was carried out on the
specimens with only two plies of CFRP. The typical finite element mesh and boundary
conditions are shown in Figure 6.44. The element size was maintained at approximately 9
mm.
Horizontal
Lf CFRP
roller

X
Vertical
roller Horizontal roller

Figure 6.44 Two-dimensional mesh of a typical specimen (T3)

The concrete was modelled using eight-node quadrilateral isoparametric plane


stress elements (CQ16M). The CFRP was modelled using three-node, two-dimensional
beam elements (CL9BE). The basic variables of these beam elements include the
translations ux and uy and the rotation φz in the nodes (Figure 6.45b). The beam elements
were numerically integrated over the cross-section and along their axis. Two-point and
three-point Gauss integration schemes were used along the element axis and in the
element cross-section respectively. The beam elements were connected to concrete
elements through 3+3 node structural interface elements (CL12I). The normal and shear
tractions, and the normal and shear relative displacements across the interface are
illustrated in Figure 6.46. A three-point Newton-Cotes integration scheme and linear
interpolation were used for the interface elements. The nodes for the CFRP were offset
from the concrete surface by a distance equal to the sum of the adhesive thickness and
half of the composite thickness.

246
SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS

η
1

2
ξ 3

Figure 6.45 L7BEN element

Interface element
CFRP element

Concrete element

(a) Location

5 uy tn
4 6 ux
y
1 x 3 tt
z 2

(b) Topology (c) Displacements (d) Tractions


Figure 6.46 3+3 node structural interface element (CL12I)

6.8.2 Material models


The only nonlinearity in the model is from the bond-slip relationship adopted for
the interface elements between the CFRP and concrete. Popovics’ equation (Equation 2.1)
was used to describe the bond-slip relationship (Figure 6.47). The parameters defining the
relationship for three different CFRP widths are listed in Table 6.7. The parameters were
selected to obtain the best fitted results.

τmax

s1

Figure 6.47 Idealised bond-slip relationships

247
CHAPTER 6

Table 6.7 The parameters defining Popovics’ bond-slip curve


CFRP width τmax s1
a
(mm) (MPa) (mm)
100 4.2 0.05 3
70 4.8 0.05 3
50 5.5 0.05 3

6.8.3 Results
The predicted ultimate load capacities are compared with the actual values in
Figure 6.48. The predicted load-slip curves and the CFRP strain distributions in two
example specimens, T5 and T1, are compared with the experimental ones in Figure 6.18
and Figure 6.20, respectively.

The load capacities predicted are quite close to the experimental values. The
bond-slip based models have lower initial stiffness compared with the testing. However,
in general, the curves are reasonably similar.

Thus the bond-slip based model can be used as an approximate solution to


describe the general behaviour of the bond between CFRP and concrete.

40 Varying H bf
parameters L
35
Maximum load (kN

30
25
20
15
10
Exp. Bond-slip
5
0
T12a

T13a
T12b

T13b
T1a

T2a

T3a

T4a

T5a

T6a

T7a

T8a
T1b

T2b

T3b

T4b

T5b

T6b

T7b

T8b

Figure 6.48 Correlation of peak loads predicted by the bond-slip based model and
measured in the experiment

248
SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS

30 25
Total load (kN)

Total load (kN)


25 20
20 T1 (Bond-slip)
15
15 T5a (Exp.) 10 T1a (Exp.)
10
5 T5 (Bond-slip) 5
0 0
0.00 0.10 0.20 0.30 0.40 0.50 0.00 0.02 0.04 0.06 0.08 0.10
Slip at 42 mm from loaded edge Slip at 42 mm from loaded edge

(a) T5 (b) T1
Figure 6.49 Comparison of the load-slip curves of specimens T5a and T1a

5000 Exp. (kN) Bond-slip (kN)


10.0 10.8
4000 15.0 15.7
20.0 20.1
Microstrain

3000 26.0 26.3


27.8 28.3

(a) T5 2000

1000

0
0 50 100 150 200
Distance from loaded edge (mm)

2500 Exp. (kN) Bond-slip (kN)


10.0 10.0
2000 15.0 14.7
18.0 18.1
Microstrain

1500 20.0 20.1

(b) T1 1000

500

0
0 50 100 150 200
Distance from loaded edge (mm)

Figure 6.50 Comparison of CFRP strain distributions of specimens T5a and T1a

6.9 Summary of findings


In this chapter, a nonlinear finite element analysis of the shear-lap tests has been
presented. Concrete mechanical properties and concrete modelling were reviewed before
the numerical model and the numerical results were reported. The influence of meshing,

249
CHAPTER 6

and inputs for concrete tensile properties and adhesive stiffness was also discussed. The
findings from this work are:

• The nonlinear finite element model was able to simulate the test behaviour relatively
well by predicting the peak load, strain distributions and load-slip curves with reasonable
accuracy. The predicted crack patterns were also quite comparable to those observed in
the experiments indicating two typical failure modes: interfacial debond and shear-
tension failure. The behavioural trends were predicted well when the CFRP bond length,
CFRP thickness, CFRP width and support block height were varied.
• The numerical results were slightly sensitive to mesh sizes, mesh types and mesh
orientation. Adhesive stiffness had little influence on the model behaviour. The major
influence was from the concrete tensile properties. The effect was different in the
specimens failing by interfacial debond and the ones failing by shear-tension.
• The local bond-slips, which were often assumed to be constant along the bond, were
found to be variable due to the presence of intersecting inclined cracks. This finding was
consistent with the experimental observations.
• An approximate solution to the bond between the CFRP and concrete could be
obtained using a bond-slip relationship following Popovics’ equation.

250
SMEARED CRACK MODELLING OF RETROFITTED BEAMS

CHAPTER 7 - SMEARED CRACK MODELLING OF


RETROFITTED BEAMS

7.1 Introduction
In this chapter, nonlinear finite element modelling of the retrofitted RC beams
under bending using smeared cracks is presented. The beams are modelled using DIANA
(version 8.1.2). The modelling is carried out for beams under four-point and three-point
bending. The models are verified using the data reported in Chapter 5. A parametric study
is also presented to investigate the effect of several parameters.

7.2 Modelling of beams under four-point bending


7.2.1 Finite element idealisation
A typical finite element mesh and boundary conditions are shown in Figure 7.1.
The element size was maintained at approximately 12x12 mm. The aspect ratios (length
over height) ranged from 1.0 to a 1.2. Since the beam geometry, loading and boundary
conditions were symmetrical about the centreline, only half of the beam was modelled.
The model was supported vertically at the base and horizontally along the centreline with
roller supports. Loading was applied to a single node on the top of the beam.

Load Concrete above tension


Steel plate reinforcement

Horizontal X
Vertical roller CFRP Concrete below tension roller
reinforcement
Figure 7.1 Two-dimensional mesh of beams loaded in 4-point bending

The concrete was modelled using four-node quadrilateral isoparametric plane


stress elements (Q8MEM) as shown in Figure 7.2a. Each element has eight degrees of
freedom (dof) with two displacements, ux and uy, at each node. A 2 x 2 Gaussian

251
CHAPTER 7

integration scheme was used. The steel and CFRP reinforcements in the beams were
modelled using beam elements (L7BEN). This element type is a two-node, two-
dimensional beam element with the basic variables including the translations ux and uy
and the rotation φz in the nodes (Figure 7.2b). The element was numerically integrated
over the cross-section and along their axis. Two-point and three-point Gauss integration
schemes were used along the element axis and in the element cross-section, respectively.
The reinforcement elements were connected to concrete elements through 2+2 node
structural interface elements (L8IF). The normal and shear tractions, and the normal and
shear relative displacements across the interface are illustrated in Figure 7.3. Three-point
Newton-Cotes integration scheme and linear interpolation were used. The nodes for the
steel reinforcement were superimposed on top of the concrete nodes, whereas the nodes
for the CFRP reinforcement were offset from the concrete beam soffit by a distance equal
to the sum of the adhesive thickness and half of the composite thickness.

3
4 η
η 1
ξ
ξ 2
1
2

(a) (b)
Figure 7.2 Q8MEM element (a) and L7BEN element (b)

Concrete element

Interface element

Reinforcement
element

(a) L8IF location

uy tn
4 3 ux
y
1 x 2 tt
z

(b) Topology (c) Displacements (d) Tractions


Figure 7.3 2+2 node structural interface element (L8IF)

252
SMEARED CRACK MODELLING OF RETROFITTED BEAMS

45

177

38

245 125 125 125 125 125 125 125 125 75

All dimensions are in mm


Figure 7.4 Steel and CFRP reinforcement locations in model E1

Figure 7.4 plots the location of the reinforcement elements. Since anchorage
failure of steel reinforcement was not observed, the bended portions of the tension bars
were not modelled. The longitudinal bar locations were simplified as straight lines
extending from one end of the beam to the other end. At the longitudinal tension
reinforcement level, the thickness of the concrete was reduced to account for the
reduction of concrete volume due to the presence of the bars. This was achieved by
reducing the thickness of the concrete elements just below the tensile beam elements such
that the concrete area lost was equivalent to the area of tension reinforcement. By doing
so, the weakest shear plane was assumed to be in the concrete elements just below the
tension reinforcement. In the experiments, a portion of the strength of the failure plane
was contributed from the bond between the steel bars and concrete. However, this
contribution was small and it was ignored in the present analysis.

Tension
reinforcement
(typ.)

Figure 7.5 Reduced concrete element width

To account for the influence of the stirrup extrusions when they intersected the
debond crack, the beam elements representing the vertical stirrups were extended
downward passing the longitudinal tension reinforcement by an element size. The beam
element extrusions crossed the weak layer of concrete elements.

253
CHAPTER 7

In the experiments, a steel plate was placed under the actuator to distribute the
load over a small area. The supports also had a certain contact area with the beam to
provide an even load distribution. As a result, localised crushing in the region around the
load points and supports were not observed in the experiments. In the finite element
models, 20-mm thick steel plates, modelled using Q8MEM elements, were added at the
load location. A layer of interface elements were placed between the plate and concrete to
model the friction between the two materials. The steel plate elements were assumed to
have linear elastic properties. To further overcome the problem of localised crushing in
the finite element models, selected concrete elements around the load point and support
were assumed to have linear elastic properties.

The beams strengthened with two plies of CFRP failed either by intermediate span
debond or combination of intermediate span debond and end debond. In these beams,
intermediate span debond occurred either on the clamped or unclamped side of the
beams. Therefore, for each of these beams, two FE models were built. The behaviour of
the unclamped side was studied using the model described above. To model the failure on
the clamped side of the beam, a prestressed steel clamp was added (Figure 7.6). The
clamp consisted of two steel plates connected by a steel bolt. The steel plates were
modelled using Q8MEM elements and the bolt was idealised by a beam element L7BEN.
A frictional slip was assumed between the plate and the concrete and between the plate
and the CFRP. An initial strain of 500×10-6 was applied to the bolt before loading.

Concrete element
Bolt element
Interface element
Common node at
CFRP element connection of bolt and
Steel plate steel plate

Figure 7.6 Modelling of the steel clamp

254
SMEARED CRACK MODELLING OF RETROFITTED BEAMS

7.2.2 Material models


7.2.2.1 Concrete
To avoid stress-locking problem, a rotating crack model was used for the concrete
elements below the tension reinforcement. The concrete in this region behaved like
unconfined plain concrete. The shear strength of the cracked surfaces resulted from
aggregate interlock would be very small and could be ignored. The concrete portion
above the tension steel was reinforced with steel bars and had certain shear strength after
cracking. Therefore, it was modelled with a fixed crack model.

Concrete properties were chosen in a similar manner as those used in modelling of


shear-lap specimens. The properties adopted are summarised in Table 7.1. In the fixed
crack model, an additional factor is needed to describe the shear stiffness of cracked
concrete. This factor is called shear retention factor, β, and it indicates the percentage of
elastic shear capacity that remains after cracking. The remaining capacity is due to
aggregate interlock in the reinforced concrete. In the current implementation in DIANA,
only the constant stiffness reduction model is available, which is given by Equation 7.1
and illustrated in Figure 7.7. In reality, the shear stiffness that remains after cracking is a
function of the strain normal to the crack. The factor β was assumed to be 0.05 in the
present analysis.

G cr = β G (7.1)

βG
γ
Figure 7.7 Concrete shear strength after cracking

255
CHAPTER 7

Table 7.1 Concrete properties used in the finite element model


Property S beams E beams
Elastic modulus (MPa) 27800 29400
Poisson’s ratio 0.2 0.2
Compressive strength (MPa) 47.7 53.7
Tensile strength (MPa) 3.51 3.86
Fracture energy (N/mm) 0.188 0.188

7.2.2.2 Steel
The main flexural and shear steel reinforcements in the finite element models
were assumed to be an isotropic linear elastic material up until the yield point. Yielding
of the reinforcement was based on the yield criterion of Von Mises with strain hardening.
The elastic moduli, yield stresses and hardening diagrams used in the models were
determined from the tensile testing as described in Appendix B.2. The moduli and yield
stresses are summarised in Table 7.2. The steel plates were assumed to be a linear elastic
material.

Table 7.2 Steel properties used in the finite element model


Elastic modulus Yield stress
Bar type
(MPa) (MPa)
Main flexural reinforcement N12 205000 551
Shear reinforcement Y10 204000 334
Shear reinforcement Y6 238000 423
Steel plate 200000* -

* Property not tested.

7.2.2.3 Interfaces
In reinforced concrete, the interaction between the reinforcement and the concrete
is highly complex. The interaction is governed by secondary transverse and longitudinal
cracks in the vicinity of the reinforcement. This behaviour can be modelled with a bond-
slip mechanism (de Witte and Kikstra, 2003). The bond-slip constitutive law in DIANA is
based on a ‘total deformation’ theory, which expresses the tractions as a function of the
total relative displacements. The relationship between the normal traction and the normal

256
SMEARED CRACK MODELLING OF RETROFITTED BEAMS

relative displacement is assumed to be linear elastic, whereas the relationship between the
shear traction and the slip is assumed to follow a nonlinear function. In the present
analysis, a cubic function proposed by Dorr (1980) was used to describe the bond-slip
relation between the reinforcement rebars and concrete (Figure 7.8a). The normal
stiffness was assumed to be 1000 MPa.

A different bond-slip relation was used to describe the bond between the CFRP
and concrete. The bond-slip relation followed a nonlinear curve established previously
from the shear-lap tests (Figure 7.8b). The normal stiffness was calculated using the
following equation.
D11 = E / t (7.2)
where the modulus, E, was taken as 8500 MPa and the bond thickness, t, was taken to be
1 mm.

8 Bond stress (MPa) 5


Bond stress (MPa)

6 4
3
4 2
2 1
0 0
0 0.05 0.1 0.15 0.2 0.25 0 0.05 0.1 0.15 0.2 0.25
Slip (mm) Slip (mm)

(a) (b)

Figure 7.8 Bond-slip relationships adopted for steel bar/concrete interface (a) and
CFRP/concrete interface (b)

7.2.2.4 CFRP composite


The CFRP properties were similar to those used in modelling of shear-lap tests.

7.2.3 Solution procedure


In the models, load was ‘applied’ in steps or increments varied from 0.05 to 0.1
mm so that there were about 200 steps before the peak load was reached. At each load
step, the Quasi Newton iterative scheme was used to bring the internal forces to an
acceptable level of equilibrium. This scheme is illustrated in Figure 7.9, in which it
essentially uses the information of previous solution vectors and out-of-balance force
vectors during the increment to achieve a better approximation. This scheme proved to be

257
CHAPTER 7

suitable for the beam models since it converged quickly for most steps. The convergence
criterion adopted was based on the energy norm composed of internal forces and relative
displacements. A new reference norm was determined at the start of each step. In all of
the models, the tolerance for convergence was set to 0.0005 to ensure accurate and good
results. The maximum number of iterations for each load step or increment was set to
100. To improve the convergence, the line search algorithm was also used in the study.

Figure 7.9 Quasi-Newton iteration (de Witte and Kikstra, 2003)

7.2.4 Verification
To verify that the finite element models simulated the behaviour of the beams
properly, four outputs from the experiments and numerical simulations were compared.
They were the crack patterns at failure, the load-displacement curves and the strain
distributions in the steel and CFRP reinforcements. For clarity, only the cracks, of which
the crack strain is large enough so that the crack tensile stress has reduced by 50 % from
the maximum tensile strength, are shown in the crack patterns.

7.2.4.1 Crack patterns


The crack pattern at the peak load for the model of the control beams is illustrated
in Figure 7.10. The main cracks were vertical flexural cracks in the constant moment
region. The predicted failure mode was steel yielding and secondary compression failure
of concrete. This is the typical ductile failure of under-reinforced concrete beams. This
failure mode was also observed in the experimental study.

258
SMEARED CRACK MODELLING OF RETROFITTED BEAMS

Figure 7.10 Crack pattern in model C1

The FE models of the retrofitted beams indicated three failure modes similar to
those observed in the experiments. The modes were end debond, intermediate span
debond, and combination of end span debond and intermediate span debond. Three
respective example models are models E1, S2, and S1. Their crack patterns at different
load levels are illustrated in Figure 7.11 to Figure 7.13. The deformed shapes of the
example models at or near the peak load are plotted in Figure 7.14 to Figure 7.16.

259
CHAPTER 7

40.4 kN

50.3 kN

60.6 kN

70.0 kN

76.0 kN
(peak)

67.3 kN
(post
peak)

Figure 7.11 Crack patterns in E1 (unclamped side)

260
SMEARED CRACK MODELLING OF RETROFITTED BEAMS

40.3 kN

60.3 kN

70.0 kN

75.0 kN

77.8 kN
(peak)

68.9 kN
(post
peak)

Figure 7.12 Crack patterns in model S2 (clamped side)

261
CHAPTER 7

40.2 kN

50.6 kN

60.5 kN

70.0 kN

76.1 kN
(peak)

72.7 kN
(post
peak)

Figure 7.13: Crack patterns in model S1 (unclamped side)

76.0 kN (peak) 67.3 kN (post peak)


Figure 7.14 Deformed shapes of the unclamped side of model E1 near the peak load
(magnification factor = 20)

262
SMEARED CRACK MODELLING OF RETROFITTED BEAMS

77.8 kN ( peak) 68.9 (post peak)


Figure 7.15 Deformed shapes of the clamped side of model S2 near the peak load
(magnification factor = 20)

76.1 kN (peak) 72.7 kN (post peak)


Figure 7.16 Deformed shapes of the unclamped side of model E1 near the peak load
(magnification factor = 20)

The models of the retrofitted beams showed more closely spaced and uniformly
distributed flexural cracks compared to those in model C1. More diagonal cracks were
also observed in the shear span as the load increased beyond the capacity of the control
beams.

In model E1, a crack originated from the CFRP end and propagated upward until
it reached the tensile rebar. The crack then branched into two cracks: a crack propagating
up into the concrete beam and a crack propagating horizontally along the rebar. As the
load increased further, the horizontal crack opened wider, propagated towards the beam
middle and joined other vertical cracks in the concrete cover. When the propagation path
was intersected by a stirrup extrusion, it was forced to go around the stirrup. The tensile
stress near the stirrup extrusion induced tooth-like cracks in the concrete cover (Figure
7.17). The peak load was reached when debonding of the concrete cover near the CFRP
end was clearly observed. As more displacement was applied, the horizontal debond
crack propagated deeper toward the beam middle.

263
CHAPTER 7

Longitudinal crack Stirrup Induced transverse crack

Tensile steel Tensile force

CFRP
Propagation direction

Figure 7.17 Induced tooth-like crack

In model S2, since the CFRP layer was thinner, the flexural cracks opened wider
compared to those in model E1. The tensile steel yielded at around 60 kN. After yielding,
more diagonal cracks became visible, while the flexural cracks continued to widen. The
widest flexural crack located in the shear span near the load point. Upon additional
application of loading, the strain level in the steel rebar bridging the crack increased much
beyond the yield strain. A horizontal crack was seen along the steel rebar level. (Figure
7.18). This crack propagated to a location where the steel strain reduced to a lower value
of around the yield strain. The crack then propagated across the concrete cover towards
the bond surface and eventually along the bond surface just above the CFRP. The final
propagation path was not in the interface elements but in the concrete elements above the
joint.

The situation in model S1 was similar to that in model S2. The only exception was
that as the steel clamp was not present, an end debond crack was also seen together with
the horizontal crack near the bond line. The failure mode indicated was a combination of
intermediate span and end debond.
Wide
Region of high flexural
steel strain crack Stirrup
Steel strain
around
yielding level

Tensile steel

CFRP Tensile
force

Propagation direction

Figure 7.18 Flexure-shear crack debond

264
SMEARED CRACK MODELLING OF RETROFITTED BEAMS

The crack patterns after the peak load in the models of S beams (S models) are
compared in Figure 7.19. Both unclamped and clamped side were shown. As expected, if
the failure occurred on the clamped side, the failure mode was clearly intermediate span
debond. Meanwhile, the models of the unclamped side indicated a combination of
intermediate span and end debond. In general, the patterns matched reasonably well with
the observations from the experimental study.

The models of E beams (E models) are compared in Figure 7.20. All E models
predicted end debond failure. The crack patterns predicted are also close to those
observed in the experimental study.

(S1U)

(S1C)

(S2U)

(S2C)

(S3U)

(S3C)

S#U: unclamped side; S#C: clamped side

Figure 7.19 Crack patterns in S models at the peak loads

265
CHAPTER 7

(E1)

(E2)

(E3)

(E4)

(E5)

Figure 7.20 Crack patterns in E models (unclamped side) at the peak loads

7.2.4.2 Load displacement behaviour


The peak loads are compared in Figure 7.21 which shows that the predicted values
by the FEA agree relatively well with the experimental measurements.

120
C beams S beams E beams
100
Maximum load (kN)

80

60

40
Experiment
20
FE
0
C1a

E1a

E2a

E3a

E4a

E5a
C1b

E1b

E2b

E3b

E4b

E5b
S1a

S2a

S3a
S1b

S2b

S3b

Figure 7.21 Correlation of peak loads as predicted by the FE model and measured
in the experiment

266
SMEARED CRACK MODELLING OF RETROFITTED BEAMS

The loading curves are compared in Figure 7.22 and Figure 7.23. The curves
generated by the finite element analysis and measured in the experiments are in good
agreement, showing similar stiffness before and after yielding of the steel.

100 100

80 80
Total load (kN)

Total load (kN)


60 S1a 60 S2a
S1b S2b
40 40
S1N (FE) S2N (FE)
20 20
S1C (FE) S2C (FE)
0 0
0 10 20 30 0 10 20 30 40
Midspan deflection (mm) Midspan deflection (mm)

(a) S1 (b) S2

70
60
Total load (kN)

50
S3a
40
30 S3b

20 S3N (FE)
10 S3C (FE)
0
0 10 20 30
Midspan deflection (mm)

(c) S3
Figure 7.22 Comparison of load-displacement curves for S beams

267
CHAPTER 7

80 70
70 60

Shear load (kN)


Shear load (kN)

60 50
50 E1a (Exp.) 40 E2a
40 30 E2b
E1b (Exp.)
30 20
20 E2 (FE)
E1 (FE) 10
10
0
0
0 5 10 15 20
0 5 10 15 20
Midspan deflection (mm)
Midspan deflection (mm)

(a) E1 (b) E2

70 80
60 70
Shear load (kN) 60
Shear load (kN)

50
50
40 40 E4a
30 E3a 30
E4b
20 E3b 20
10 E4 (FE)
10 E3 (FE)
0
0
0 5 10 15 20
0 5 10 15 20
Midspan deflection (mm)
Midspan deflection (mm)

(c) E3 (d) E4

80
70
Shear load (kN)

60
50
40 E5a
30
E5b
20
10 E5 (FE)
0
0 5 10 15 20
Midspan deflection (mm)

(e) E5
Figure 7.23 Comparison of load-displacement curves for E beams

7.2.4.3 CFRP and steel reinforcement strain distributions


Figure 7.26 to Figure 7.28 compare the experimental and numerical CFRP strain
distributions along the bonded joint for three example beams E1a, S2a and S1a,
respectively. Comparisons for other beams are shown in Appendix D.4. The distributions
were compared at several load levels. The experimental and numerical load levels were

268
SMEARED CRACK MODELLING OF RETROFITTED BEAMS

selected to be as close as possible to each other. In the FE models, the CFRP strains were
taken at the average of the values at Gaussian integration points 1 and 4 (Figure 7.24).
These strains represent the average strain in the bottom half of the composite. It was
found that the strains were not significantly different between the top and bottom layers
of the composite (Figure 7.25).

The simulated CFRP strain distributions are, in general, close to those in the
experimental program at both high and low load levels for all beams. The fluctuation in
the simulated CFRP strains at high load levels is due to the nature of the smeared crack
model, where the composite strains are sensitive to the presence of distributed concrete
cracks transverse to the bond.

Concrete element

Interface element
3 6
2 5 FRP element
1 4
Node 1 Node 2
Gauss integration point

Figure 7.24 Gauss integration points in a CFRP beam element

12000
Load (kN)
10000 Top Bottom
20.3 20.3
8000
Microstrain

60.3 60.3
6000 70.0 70.0

4000 77.2 77.2

2000

0
1000 800 600 400 200 0
Distance from end of FRP (mm)

Figure 7.25 Variation of CFRP strain along the top and bottom halves in model S2N

269
CHAPTER 7

Load point
4000 Load (kN)

3500 Exp. FEA


10.0 11.0
3000
20.1 20.3
2500 30.0 30.3
Microstrain

40.0 40.4
2000
50.1 50.3
1500 60.0 60.6
70.7 76.0
1000

500

0
1000 750 500 250 0
Distance from end of CFRP (mm)

Figure 7.26 Comparison of CFRP strain distributions in beam E1a

Load point
12000

Load (kN)
10000
Exp. FEA

8000 20.1 20.3


40.0 40.3
Microstrain

6000 60.0 60.3


70.0 70.0

4000 75.0 75.0


80.2 77.2

2000

0
1000 750 500 250 0
Distance from end of CFRP (mm)

Figure 7.27 Comparison of CFRP strain distributions in beam S2a

270
SMEARED CRACK MODELLING OF RETROFITTED BEAMS

Load point
10000

9000

8000 Load (kN)


Exp. FEA
7000
20.0 20.6
6000
Microstrain

40.1 40.2
5000 60.0 60.5
4000 65.0 65.2

3000 70.0 70.0


73.7 76.1
2000

1000

0
1000 750 500 250 0
Distance from end of CFRP (mm)

Figure 7.28 Comparison of CFRP strain distributions in beam S1a

The strain distributions in the tensile steel reinforcement are plotted in Figure 7.29
to Figure 7.31 for three representative beams E1a, S2a and S1a, respectively, at several
load levels. Comparisons for other beams are shown in Appendix D.4.

In general, the numerical models simulate the tensile steel strain distributions
reasonably well before yielding. After yielding, the steel strain depends greatly on the
location of the main flexural cracks and therefore it is not possible to compare the
experimental measurements with the numerical results.

271
CHAPTER 7

Load point
3500
Load (kN)
3000 Exp. FEA
20.1 20.3
2500
30.0 30.3
Microstrain

2000 40.0 40.4


50.1 50.3
1500
60.0 60.6

1000 70.7 76.0

500

0
1000 750 500 250 0
Distance from end of CFRP (mm)

Figure 7.29 Comparison of tensile steel strain distributions in beam E1a

Load point
10000

9000
Load (kN)
8000 Exp. FEA
7000 20.1 20.3
40.0 40.3
6000
Microstrain

60.0 60.3
5000
70.0 70.0
4000
75.0 75.0
3000
80.2 77.2
2000

1000

0
1000 750 500 250 0
Distance from end of CFRP (mm)

Figure 7.30 Comparison of tensile steel strain distributions in beam S2a

272
SMEARED CRACK MODELLING OF RETROFITTED BEAMS

Load point
10000
Load (kN)
9000
Exp. FEA
8000
20.0 19.8
7000 40.1 40.2
6000
Microstrain

60.0 61.1
5000 65.0 64.8

4000 70.0 69.8

3000 73.7 74.9

2000

1000

0
1000 750 500 250 0
Distance from end of CFRP (mm)

Figure 7.31 Comparison of tensile steel strain distributions in beam S1a

273
CHAPTER 7

7.2.5 Sensitivity study


The above FE models produced comparable results to the experiments in terms of
the failure mechanisms, the load displacement behaviour and the strain distributions. The
models can be stated to be valid. In the following sections, a parametric study is carried
out to investigate the sensitivity of the numerical results to a number of parameters. The
comparison is done for only two specimens, E1 (unclamped side) and S2 (clamped side),
since they experienced two typical failure modes: end debond and intermediate span
debond, respectively.

7.2.5.1 Model sensitivity to concrete tensile strength and fracture energy


To study the sensitivity of the finite element model to concrete tensile properties,
different values of the fracture energy, Gf and concrete tensile strength, fct, were adopted.
They are listed in Table 7.3, where the base models are labelled as ‘F3.86-G0.188’ for
beam E1 and ‘F3.51-G0.188’ for beam S2.

Table 7.3 Concrete tensile property variations


Tensile strength fct Fracture energy Gf
Beam Designation (MPa) (N/mm)
Source Value Source Value
E1 F3.86 – G0.188 CEB1 3.86 Trunk4 0.188
E1 F3.86 – G0.124 CEB1 3.86 Bazant5 0.124
1 1
E1 F3.86 – G0.091 CEB 3.86 CEB 0.091
E1 F4.40 – G0.188 AS-F2 4.40 Trunk4 0.188
E1 F2.93 – G0.188 AS-D3 2.93 Trunk4 0.188
S2 F3.51 – G0.188 CEB1 3.51 Trunk4 0.188
S2 F3.51 – G0.117 CEB1 3.51 Bazant5 0.124
S2 F3.51 – G0.086 CEB1 3.51 CEB1 0.091
S2 F4.14 – G0.188 AS-F2 4.14 Trunk4 0.188
S2 F2.76 – G0.188 AS-D3 2.76 Trunk4 0.188

Notes: 1) CEB-FIB Model Code 1990


2) AS3600, Flexural tensile strength
3) AS3600, Direct tensile strength
4) Trunk and Wittmann (1998)
5) Bazant and Becq-Giraudon (2002)

274
SMEARED CRACK MODELLING OF RETROFITTED BEAMS

The crack patterns at the peak loads for the models with different Gf values are
compared in Figure 7.32. The load-displacement curves are compared in Figure 7.33. It
can be seen that Gf has a moderate influence on the behaviour of the beams. The models
with a lower Gf tend to experience more cracking in the concrete cover prior to failure.
The peak loads are influenced most in E1 models, where lowering Gf results in a lower
ultimate capacity.

The crack patterns at the peak loads for the models with different fct values are
compared in Figure 7.34. The load-displacement curves are compared in Figure 7.35. fct
has a minor influence on the behaviour of the beams. The peak loads predicted for beams
E1 and S2C are similar for three different values of fct.

F3.86 – G0.188 F3.51 – G0.188

F3.86 – G0.124 F3.51 – G0.117

F3.86 – G0.091 F3.51 – G0.086


(a) (b)
Figure 7.32 Comparison of crack patterns at the peak loads in beams E1 (a) and
S2C (b) predicted by models with different Gf values

275
CHAPTER 7

E1a S2a
E1b S2b
E1 (F3.86-G0.188) S2C (F3.51-G0.188)
E1 (F3.86-G0.124) S2C (F3.51-G0.117)
80 E1 (F3.86-G0.091) 100 S2C (F3.51-G0.086)
70
80
Total load (kN)

Total load (kN)


60
50 60
40
30 40
20
20
10
0 0
0 5 10 15 0 10 20 30 40
Midspan deflection (mm) Midspan deflection (mm)

(a) E1 (b) S2C


Figure 7.33 Comparison of load-deflection curves predicted by models with
different Gf values

F4.40 – G0.188 F4.14 – G0.188

F3.86 – G0.188 F3.51 – G0.188

F2.93 – G0.188 F2.76 – G0.188


(a) (b)
Figure 7.34 Comparison of crack patterns at the peak loads in beams E1 (a) and
S2C (b) predicted by models with different fct values

276
SMEARED CRACK MODELLING OF RETROFITTED BEAMS

E1a S2a
E1b S2b
E1 (F4.40-G0.188) S2C (F4.14-G0.188)
E1 (F3.86-G0.188) S2C (F3.51-G0.188)
100 E1 (F2.93-G0.188) 100 S2C (F2.76-G0.188)
80 80
Total load (kN)

Total load (kN)


60 60

40 40

20 20

0 0
0 5 10 15 0 10 20 30 40
Midspan deflection (mm) Midspan deflection (mm)

(a) E1 (b) S2C


Figure 7.35 Comparison of load-deflection curves predicted by models with
different fct values

7.2.5.2 Model sensitivity to concrete shear retention


The shear retention factor, β, is used to account for the effect of aggregate
interlock in concrete. The main beam body above the tension reinforcement is reinforced
and its behaviour can be influenced significantly by β. To study the sensitivity of the FE
model to the shear retention factor, three different values were adopted. They are listed in
Table 7.4.

Table 7.4 Shear retention factor variations


Designation β
Beta0.001 0.001
Beta0.05 0.05
Beta0.10 0.10

Comparisons of crack patterns and loading curves for two representative cases are
included in Figure 7.36 to Figure 7.37. The crack patterns are similar even though the
models with a higher shear retention capacity seem to have more cracks in the shear
spans. These models also show stiffer behaviour as expected.

277
CHAPTER 7

Beta0.001 Beta0.001

Beta0.05 Beta0.05

Beta0.10 Beta0.10
(a) (b)
Figure 7.36 Comparison of crack patterns at the peak loads in beams E1 (a) and
S2C (b) predicted by models with different β values

E1a S2a
E1b S2b
E1 (Beta0.001) S2C (Beta0.001)
E1 (Beta0.05) S2C (Beta0.05)
100 E1 (Beta0.10) 100 S2C (Beta0.10)
80 80
Total load (kN)

Total load (kN)

60 60

40 40

20 20

0 0
0 5 10 15 0 10 20 30
Midspan deflection (mm) Midspan deflection (mm)

(a) E1 (b) S2
Figure 7.37 Comparison of load-deflection curves predicted by models with
different β values

7.2.6 Parametric study


In section 5.4.5.3, a parametric study is presented using limited experimental
results. In this section, another parametric study is implemented using the previously
developed finite element model based on smeared cracking. The influence of the CFRP
thickness, tension reinforcement amount, CFRP bond length and concrete cover on
debonding loads and debonding modes is investigated. References to the experimental

278
SMEARED CRACK MODELLING OF RETROFITTED BEAMS

results are also included. The study is carried out on a base specimen with the dimensions
and material properties similar to those of beam E1a. The mode intermediate span debond
is abbreviated as IS debond in the following paragraphs.

Figure 7.38 shows the behavioural trend of retrofitted RC beams bonded with
different composite cross sectional areas, Af. As the area increases, the failure mode shifts
in the following order:
concrete crushing => IS debond => mixed IS and end debond => to end debond
It is clear that increasing the CFRP amount does not always lead to an increase in the
beam capacity. The optimal CFRP cross sectional area, when the capacity of the beam
reaches the maximum value, is around 50 mm2 (or 3 plies of CFRP of 100 mm width).
The corresponding failure mode is mixed debond. When the failure mode is concrete
crushing or FSCD, the beam capacity increases with Af. However, when the mode is end
debond, the beam capacity reduces gradually with Af.

IS debond Mixed
Crush End debond
debond
Ultimate shear load (kN)

100

80

60

40
Exp.
FE
20

0
0 50 100 150 200
2
A f (mm )

Figure 7.38 Behavioural trend of retrofitted beams with different CFRP cross
sectional areas

Figure 7.39 shows the behavioural trend of retrofitted beams reinforced with
different amounts of tensile rebars. When the tensile rebar cross-sectional area, As, is
small, flexural and flexural-shear cracks in the intermediate span can open wider.
Therefore, the failure mode is IS debond. As As increases, the failure mode shifts to a
mixed mode of IS and end debond, and then to end debond.

279
CHAPTER 7

IS debond Mixed debond End debond

100
Ultimate shear load (kN)

80

60

40 Exp
FE
20

0
0 50 100 150 200 250 300 350 400
2
A s (mm )

Figure 7.39 Behavioural trend of retrofitted beams with different tension


reinforcement cross sectional areas

Figure 7.40 presents the relationship between the ultimate shear load and the ratio
of the CFRP bond length, Lf, to the shear span, a. The failure mode is end debond with
the exception of the beam with the longest bonded length, where the mode is IS debond.
The general trend is that the beam capacity increases moderately as the bond length
increases.

End debond IS debond


Ultimate shear load (kN)

100

80

60

40
Exp.
FE
20

0
0.5 0.6 0.7 0.8 0.9 1.0
Lf / a

Figure 7.40 Behavioural trend of retrofitted beams with different CFRP bond
lengths

280
SMEARED CRACK MODELLING OF RETROFITTED BEAMS

Figure 7.41 illustrates the behavioural trend of retrofitted beams with different
concrete covers. All of the beams investigated fail by end debond. It is clear that the
cover has an insignificant influence on the beam capacity.

Ultimate shear load (kN) End debond

100

80

60

40
Exp.
FE
20

0
10 20 30 40 50
c (mm)

Figure 7.41 Behavioural trend of retrofitted beams with different concrete covers

7.3 Modelling of beams under three-point bending


7.3.1 Finite element idealisation and material models
These beams were loaded in an unsymmetrical manner. Therefore, the whole
beam needed to be modelled (Figure 7.42 and Figure 7.43). The steel clamps were
modelled in a similar way as for the beams tested in four-point bending. The CFRP
vertical strap layer was superimposed on top of the concrete layer and connected through
a layer of the interface elements. These elements were placed both longitudinally and
transversely (Figure 7.44).

The same material models as described in Section 7.2.2 were used. Orthotropic
material properties were used for the vertical CFRP straps. The CFRP stiffness in the
transverse direction was assumed to be the same as that of the adhesive. To simulate the
bond behaviour between the concrete and CFRP composite, the nonlinear bond-slip
model derived previously from the shear-lap tests was used. In those shear-lap tests, the
stress distribution in the CFRP was close to uniform. However, in the vertical straps,
bending action resulted in non-uniform stress distribution, which could lead to premature
rupture or debonding. To model the distribution accurately, a much finer mesh is
required. This was however undesirable since the computational time would increase

281
CHAPTER 7

significantly. Therefore, the current models are not expected to be able to pick up
debonding and rupture of the composite.

In the experiments, after the peak load was reached and a sudden failure was
observed, loading was continued until complete debonding of the composite. The
dynamic nature of the failure caused severe damage to the straps, concrete and bonding
between them. This action is not picked up by the current static model. Therefore, the
simulation was done only up to the peak load and the behaviour after peak was not
considered.

For the models of the beams with prestressed straps, i.e. models A2a and A2b, the
concrete just above the wedge was subjected to a compressive field and therefore behaved
more liked reinforced concrete. Therefore, the concrete portion above the wedges was
modelled using the fixed smeared cracks.

Figure 7.42 Mesh of a beam loaded in three-point bending

282
SMEARED CRACK MODELLING OF RETROFITTED BEAMS

(a) E3b2

concrete
strap elements
interface elements
main CFRP

(b) A1a, E3a2

(c) A1b

concrete
strap elements
interface elements
main CFRP

(d) A2a and E1b2 (for E1b2, end strap moved to the left by 60 mm)

(e) A2b and E5a2


Figure 7.43 Modelling of reinforcements

283
CHAPTER 7

Interface
elements Concrete
layers

Unbonded area CFRP strap

Interface
Longitudinal
elements
CFRP

Figure 7.44 Location of interface elements in the FE mesh for model A2a

7.3.2 Verification
7.3.2.1 Crack patterns
The crack patterns at the peak load of retested beams and A beams are illustrated
in Figure 7.45 and Figure 7.46, respectively. The beams with one strap failed by
intermediate span debond; whereas the beams with three straps failed by debonding of the
straps in the concrete cover followed by debonding of the longitudinal CFRP between the
straps (Figure 7.47). Slight crushing of concrete under the load point was also observed.
The patterns are similar to those observed in the experiments. However, complete
debonding of the straps was not seen.

284
SMEARED CRACK MODELLING OF RETROFITTED BEAMS

(E3b2)

(E3a2)

(E1a2)

(E5a2)
Figure 7.45 Crack patterns in retested beams at peak

285
CHAPTER 7

(A1a)

(A1b)

(A2a)

(A2b)
Figure 7.46 Crack patterns in A beams at peak

debond debond

(a) A1b (b) A2b


Figure 7.47 Deformed shapes after peak of beams with three straps (magnification
factor = 10)

7.3.2.2 Load displacement behaviour


The peak loads are compared in Figure 7.48. The loading curves are compared in
Figure 7.49 and Figure 7.50. Because of the difficulties mentioned previously, the

286
SMEARED CRACK MODELLING OF RETROFITTED BEAMS

predictions do not match as well as for the beams tested in four-point bending. However,
the predictions and the measurement are still in relatively good agreement.

140
120

Maximum load (kN)


100
80
60
40 Experiment FE
20
0 E3b2

E3a2

E1a2

E5a2

A1a

A2a
A1b

A2b
Figure 7.48 Correlation of peak loads as predicted by the FE model and measured
in the experiment

140 140
120 120
Shear load (kN)

Shear load (kN)

100 100
80 80
E3b2 (Exp.)
60 60 E3a2 (Exp.)
E3b2 (FE)
40 40
E3a2 (FE)
20 20
0 0
0 2.5 5 7.5 10 12.5 15 0 2.5 5 7.5 10 12.5 15
Midspan deflection (mm) Midspan deflection (mm)

(a) E3b2 (b) E3a2

140 140
120 120
Shear load (kN)
Shear load (kN)

100 100
80 80
60 60 E5a2 (Exp.)
E1a2 (Exp.)
40 40
E1a2 (FE) E5a2 (FE)
20 20
0 0
0 2.5 5 7.5 10 12.5 15 0 2.5 5 7.5 10 12.5 15
Midspan deflection (mm) Midspan deflection (mm)

(c) E1a2 (d) E5a2


Figure 7.49 Comparison of load-deflection curves for retested beams

287
CHAPTER 7

140 140
120 120
Shear load (kN)

Shear load (kN)


100 100
80 80
A1a (Exp.)
60 60 A1b (Exp.)
A1a (FE)
40 40 A1b (FE)
20 20
0 0
0 2.5 5 7.5 10 12.5 15 0 2.5 5 7.5 10 12.5 15
Midspan deflection (mm) Midspan deflection (mm)

(a) A1a (b) A1b

140 140
120 120
Shear load (kN)
Shear load (kN)

100 100
80 80
60 A2a (Exp.) 60 A2b (Exp.)
40 A2a (FE) 40 A2b (FE)
20 20
0 0
0 2.5 5 7.5 10 12.5 15 0 2.5 5 7.5 10 12.5 15
Midspan deflection (mm) Midspan deflection (mm)

(c) A2a (d) A2b


Figure 7.50 Comparison of load-deflection curves for A beams

7.3.2.3 CFRP and steel reinforcement strain distributions


Figure 7.51 compares the experimental and numerical CFRP strain distributions in
the shear span for an example beam A1a. The strain development in the side straps is
compared in Figure 7.52. Comparisons for other beams are shown in Appendix D.5.
Reasonable matching can be also seen.

288
SMEARED CRACK MODELLING OF RETROFITTED BEAMS

Load point
6000
Load (kN)
5000 Exp. FE

20.0 18.2
4000
40.0 40.1
Microstrain

3000 60.1 59.9

80.0 80.6
2000
94.5 93.6
1000

0
500 400 300 200 100 0
Distance from end of CFRP (mm)

Figure 7.51 Comparison of longitudinal CFRP strain distributions in beam A1a

3500
3000
2500
Exp. (Front)
Microstrain

2000
Exp. (Back)
1500
FE
1000
500
0
-500 0 25 50 75 100
Shear load (kN)

Figure 7.52 Comparison of strain development on U-strap sides in beam A1a

7.4 Summary of findings


In this chapter, a nonlinear finite element analysis of the beam tests has been
presented. All beams without and with anchorage were modelled using smeared crack
approach. The findings from this work are:

• The nonlinear finite element analysis using smeared cracks was able to simulate the
test behaviour reasonably well, predicting similar peak loads, load-deflection curves
and strain distributions.

289
CHAPTER 7

• The failure mechanisms were also captured showing two debonding modes, end and
intermediate span debond, similar to those observed in the experiments.
• The finite element analysis indicated that the behaviour of retrofitted RC beam
depended a great deal on the amount of CFRP and steel reinforcement. The failure
mode changed from intermediate span debond to end debond as the CFRP amount
increased. The beams with a low amount of tension reinforcement were more likely to
fail by IS debond.
• The CFRP bonded length also had a moderate influence on the beam end debond
capacity. The longer the bond length, the higher the capacity was. By extending the
CFRP close to the support, end debond could be suppressed and the failure mode
shifted to intermediate span debond.
• Despite some discrepancy, the effect of the anchorage straps was simulated generally
well. The straps proved to be able to prevent or delay debonding. However,
debonding and rupture of the straps were not captured with the two-dimensional
model.

290
DISCRETE/SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS AND RETROFITTED BEAMS

CHAPTER 8 - DISCRETE/SMEARED CRACK MODELLING


OF SHEAR-LAP SPECIMENS AND RETROFITTED
BEAMS

8.1 Introduction and scope


The smeared crack models described in the previous two chapters proved to be
able to describe the behaviour of the shear-lap specimens and retrofitted beams under
testing reasonably well. However, the smeared crack model has some limitations. In this
modelling method, since displacement continuity is maintained across a crack, the
distribution of strains in regions adjacent to cracks can be overlooked. The cracking
considered in the model is based mainly on mode-I fracture while mode-II fracture is only
partly accounted for by using a constant shear retention factor. Therefore, to obtain a
more realistic prediction of the strain distribution near a crack and to be able to
investigate the effect of mode-II fracture, a discrete crack model, which allows
displacement discontinuity across a crack and uses both mode-I and mode-II crack
prediction criteria, needs to be used.

One problem with the use of discrete crack modelling is the difficulties in
simulating structures of which the behaviour is influenced by a large number of cracks.
Debonding failure of CFRP in shear-lap specimens or retrofitted beams is one of those
cases, where the structural behaviour is influenced by many cracks including both local
and global cracks. The local cracks are in the proximity of the bond area whereas the
global cracks are the distributed cracks in other parts of the structure. Therefore, to
simulate the behaviour of retrofitted members, a combination of discrete and smeared
crack modelling approaches can be used, where the major local cracks are simulated by
discrete cracks and the minor distributed cracks are modelled using smeared cracks.

This chapter reports the results from a numerical investigation of the behaviour of
the shear lap and beam specimens under testing using discrete/smeared cracks. Modelling
is implemented using the computer program MERLIN II. The program is developed by
Dr Jan Cervenka and Ron Reich at the University of Colorado where Professor Victor

291
CHAPTER 8

Saouma is the principle investigator. The model of discrete cracks is based on nonlinear
fracture mechanics.

The goal of the study presented in this chapter is to investigate the possibility of
modelling debonding failures in retrofitted structures using discrete/smeared crack
modelling based on nonlinear fracture mechanics. The discrete/smeared model is also to
further validate the failure mechanisms, which are used to develop the prediction models
presented in Chapter 9. For those purposes, only six shear-lap configurations, T1 to T6,
and eight retrofitted beams (E and S beams) are modelled in this study.

8.2 Brief review on concrete crack models available in MERLIN


8.2.1 Localised failure of concrete
There are several nonlinear fracture models for concrete available in MERLIN.
These models are described in more details in Saouma (2002). The discrete crack model
used in this study is the Interface Crack Model (Cervenka, 1994). The model is a
generalization of the classical Fictitious Crack Model by Hillerborg et al. (1976). These
two fracture models are briefly described below.

The Fictitious Crack Model proposed by Hillerborg et al. (1976) is probably the
most widely used in nonlinear fracture mechanics finite element analysis (Saouma, 2002).
Its basic concept is illustrated in Figure 8.1. It is related well to the true physical nature of
cracks in concrete. The model is based on the softening curve of the tensile stress versus
crack opening displacement or COD.

292
DISCRETE/SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS AND RETROFITTED BEAMS

COD

Microcracks Spalling
Fracture Process Zone True crack

Effective crack

f’t

s1 GF Thickness

w1 w2

Figure 8.1 Hillerborg’s Fictitious Crack Model (Saouma, 2002)

In Hillerborg’s model, it is assumed that there is no sliding movement along the


fracture process zone. However, this movement might be significant during crack
propagation. This effect has been considered in the more complete model, the Interface
Crack Model (ICM). The model defines the strength of an interface by a failure function.
For two-dimensional cases, the function is given as

F = τ 2 − 2c tan(φ f )(σ t − σ) − tan 2 (φ f )(σ 2 − σ 2t ) = 0 (8.1)

where φf is the angle of friction; τ is the interface traction vector; and σ is the normal
traction component (Figure 8.2). The shape of the failure function in a two-dimensional
case is shown in Figure 8.3.

Material 1
Material 1
ux, τ
uy, σ

Material 2 Material 2

Interface Interface model

Figure 8.2 Interface idealization and notations (Saouma, 2002)

293
CHAPTER 8

1
tan (φf)
Initial failure
1 fucntion
tan (φf)
c

Final failure
fucntion
σt σ

Figure 8.3 Failure surface (Saouma, 2002)

The cohesion, c, and the tensile strength, σt, are reduced by softening laws based
on a softening parameter uieff (as opposed to COD in Hillerborg’s model). The bi-linear
softening laws available in MERLIN are illustrated in Figure 8.4. As uieff increases, the
cohesion and tensile strength soften and the failure surface changes from its initial state to
its final state as shown in Figure 8.3.

σt c

σt0 c0

GIF GIIF
s1σ s1c

w1σ wσ w1c wc
uieff uieff

Figure 8.4 Bi-linear softening laws (Saouma, 2002)

To obtain uieff, the displacement vector u is decomposed into an elastic part ue and
an inelastic part ui:
u = ue + ui (8.2a)
uieff is defined as the norm of the inelastic displacement and is given by:
2 2
u ieff = u ix + u iy (8.2b)

ui is composed of the plastic up and fracturing displacement uf, and is given by:
ui = up + uf (8.2c)

294
DISCRETE/SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS AND RETROFITTED BEAMS

up = γui (8.2d)
For concrete, γ is usually assumed to be equal to 0.2 or 0.3 (Saouma, 2002).

The stress at the interface depends on the displacement and is given as:
σ = αE(u – up) (8.2e)
where
σ = {τ, σ}T (8.2f)
⎡K 0 ⎤
E = ⎢ to (8.2g)
⎣ 0 K no ⎥⎦

Kno and Kto are the initial normal and tangential stiffness.

σ σ

σi σi
Kno

Kns GIF
u ui = uieff
up ui up ui

Figure 8.5 Stiffness degradation in the equivalent uniaxial case (Saouma, 2002)

In the elastic range, no dilatancy is considered and the off-diagonal terms of the
matrix E are all equal to zero. The dilatancy is introduced after the failure limit has been
reached. The dilatancy angle is also assumed to be a function of uieff.
⎛ u ieff ⎞
φ d (u ieff ) = φ d 0 ⎜⎜1 − ⎟⎟ when u ieff ≤ u dil
⎝ u dil ⎠ (8.2h)
φ d (u ieff ) = 0 when u ieff > u dil

where φd0 is the initial dilatancy angle and udil is the maximum displacement after which
the dilatancy vanishes.

α is called integrity parameter. In compression, the crack is closed and α equals 1.


In tension, α is reduced to 1-D, where D is the damage parameter. D defines the stiffness
degradation (Figure 8.5) and is given as:

295
CHAPTER 8

Fracture area K ns σ t (u ieff )


D= = 1− = 1− (8.2i)
Total interface area K no σ t (u ieff ) + (1 − γ )u ieff K no
where Kns is the secant of the normal stiffness.

8.2.2 Distributed failure of concrete


Distributed cracks can be modelled in MERLIN using the smeared crack model
based on Rankine’s failure criterion and exponential softening. Concrete behaviour in
compression can be modelled using Menetrey-Willam’s failure criteria (Menetrey and
Willam, 1995). More detailed description of the models can be found in MERLIN User's
Manual (2001).

8.3 Modelling of shear-lap specimens under testing


This section presents the discrete/smeared crack models of twelve shear-lap
specimens, T1a to T6b, under tension. The experimental results from these tests are
described in Chapter 4.

8.3.1 Finite element idealisation


As established in Chapters 4 and 6, the shear-lap specimens failed either by
interfacial debond or shear-tension. The failure of a specimen is therefore influenced by
the interfacial crack along the bond surface and the transverse shear-tension cracks
normal to the bond surface. The smeared crack model presented in Chapter 6 showed the
locations of the critical transverse cracks in each specimen (Figure 6.8). The tips of these
cracks were located at approximately 20 and 60 mm from the loaded edge and at the
unloaded end of the CFRP. The propagation paths of these cracks were however sensitive
to the mesh type. In the model presented herein, the crack paths were generated
automatically using the CRACKER program. This tool, embedded in the mesh generation
program determines the crack propagation trajectory (on the basis of a strength criterion),
automatically remeshes to accommodate the new crack tip location, and finally launches
the analyses. In this study, CRACKER was only used to generate transverse crack paths
and the concrete properties were assumed to be linear elastic during generation. The
initial transverse cracks of 2 mm length were inserted just below the adhesive layer. Their
locations are shown in Figure 8.6. The final generated crack paths are presented in Figure
8.7.

296
DISCRETE/SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS AND RETROFITTED BEAMS

CFRP Initial cracks

Concrete

(T1) (T2) (T3)

(T4) (T5) (T6)


Figure 8.6 Initial crack locations in specimens T1-T6

(T1) (T2) (T3)

(T4) (T5) (T6)


Figure 8.7 Final transverse crack paths in specimens T1-T6

The typical final finite element mesh and boundary conditions are shown in
Figure 8.8. The concrete, CFRP and adhesive were modelled using three-node triangle
isoparametric plane stress elements. Line-to-line four-node interface elements were used
to model the crack surface (Figure 8.9). The thicknesses of CFRP and adhesive were
assumed to be the same as those used in the smeared crack model (0.7 mm per ply of
CFRP and 1 mm for adhesive).

297
CHAPTER 8

(a) T1 (b) T5
Figure 8.8 Two typical meshes of the shear-lap specimens

4 3

1 2
1 2
(a) (b)
Figure 8.9 Three-node plane stress element (a) and four-node interface element (b)

8.3.2 Material models


The CFRP and adhesive were modelled as linear elastic materials (Material 1).
Discrete cracks were represented by Interface Crack Model (Material 23: ICM). Other
non-critical cracks in concrete were simulated using smeared cracks based on Rankine-
fracturing model (Material 17). The material parameters used are summarised in Table
8.1 to Table 8.3.

The material parameters were chosen to be as close to those used for the smeared
crack model presented in Chapter 6 as possible. For the interface crack models, there are
a number of extra parameters which were not required in the smeared crack model. The
values used for these parameters were obtained through calibration with experimental
results. The normal and tangential stiffness were taken to be 1.5 times of concrete elastic
modulus. The friction and dilatancy angles were taken to be 45 degrees. The cohesion
was assumed to be 10 % higher the tensile strength.

The specific mode-II fracture energy, GfII of concrete is very hard to be


determined. Very limited experimental data is available. The only experimental study was

298
DISCRETE/SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS AND RETROFITTED BEAMS

by Karbhari and Engineer (1996), in which the interfacial facture energy of the bond was
investigated. However, the failure surface observed was within both the concrete and
bonded materials. In MERLIN User’s Manual, it is suggested that GfII can be taken as 10
times of GfI. In Niu and Wu (2001), GfII was varied from 0.5GfI to 1.0GfI. In this study,
the specific mode II fracture energy was taken to be twice the specific mode-I fracture
energy.

Table 8.1 CFRP and adhesive material model parameters


Description Unit CFRP Adhesive
Thickness mm 100 100
Elastic modulus MPa 53000 8500
Poisson's ratio 0.30 0.30

Table 8.2 Concrete smeared crack model parameters for shear-lap blocks
Description Unit Value
Thickness mm 140
Mass density kg/mm3 2.40E-06
Elastic modulus MPa 29344
Poisson's ratio 0.2
Tensile strength MPa 3.86
Fracture energy N/mm 0.188
Compressive strength MPa -53.7
Compressive critical displacement mm -0.5
Factor for return direction in Haigh-Westergaard space 0
Factor for roundness of Menetrey-Westergaard failure surface 0.55
Onset of nonlinearity in compression MPa -43.0
Plastic strain at compressive strength -1.05E-03

299
CHAPTER 8

Table 8.3 Interface crack model parameters for shear-lap blocks


Description Unit Value
Thickness mm 140
Tangential stiffness MPa/mm 44000
Normal stiffness MPa/mm 44000
Tensile strength MPa 3.86
Cohesion MPa 4.24
Friction angle degree 45
Dilatancy angle degree 45
Specific mode-I fracture energy release rate N/mm 0.188
Specific mode-II fracture energy release rate N/mm 0.376
Ratio of irreversible deformation 0.3
Maximum displacement for dilatancy mm 10
Tensile stress at the break-point for bilinear softening law MPa 0.96
Crack opening displacement at the break-point for bilinear
softening law mm 3.66E-02
Cohesion at the break-point for bilinear softening law MPa 1.06
Crack sliding displacement at the break-point for bilinear
softening law mm 6.65E-02

8.3.3 Solution procedure


The load was ‘applied’ in steps or increments of 0.003 mm in all of the models. At
each load step, the TangentStiff iterative scheme was used. This option invokes the
assembly of the global stiffness matrix using tangent material stiffness matrices at the
element level. For iteration control, RelResidErr, AbsResidErr and DispError options
were used. They specify the convergence tolerance for the relative residual error (of the
load vectors), the absolute residual error and the displacement error, respectively. The
tolerance for convergence was set to 0.01 for the two load error criteria and 0.01 for the
displacement error criterion. The maximum number of iterations for each load increment
was set to 40. These values were chosen based on the recommendations in MERLIN
User’s Manual. To improve the convergence, the line search algorithm was also used in
the study.

300
DISCRETE/SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS AND RETROFITTED BEAMS

8.3.4 Verification of finite element model


The finite element models were verified by comparing the crack patterns, the
loading curves and the CFRP strain distributions predicted by the discrete/smeared crack
models with the results from the experiments and from the smeared crack modelling.

8.3.4.1 Crack patterns


The crack opening sequences of two typical models, T5 and T1, are illustrated in
Figure 8.10 and Figure 8.11, respectively, for several load levels. In model T5, the first
transverse crack formed at 20 mm from the loaded edge. When the opening was wide
enough, at the tip of this crack, the interfacial crack started to propagate. As the load
increased further, the propagation continued gradually towards the other end. In model
T1, as the load increased, the first transverse crack opened to a certain extend before the
end transverse crack opened wide and propagated deep into the concrete block. It is clear
that these failures were similar to those observed in the experiments and predicted by the
smeared crack model.

301
CHAPTER 8

24.0 kN

27.2 kN

28.29 kN
Figure 8.10 Deformed shapes of model T5 at different load levels (magnification
factor = 60)

302
DISCRETE/SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS AND RETROFITTED BEAMS

10.1 kN

15.0 kN

17.7 kN
Figure 8.11 Deformed shapes of model T1 at different load levels (magnification
factor = 60)

The deformed shapes of all six models T1 to T6 are illustrated in Figure 8.12. It
can be seen that as the bond length increased beyond 80 mm (specimen T2), the failure
mode changed from shear-tension to interfacial debond. This trend is similar to the
observation from the experimental study and numerical investigation using smeared
cracks.

303
CHAPTER 8

(T1) (T2)

(T3) (T4)

(T5) (T6)
Figure 8.12 Discrete crack opening in models T1 to T6 at the peak loads
(magnification factor = 60)

8.3.4.2 Loading curves and CFRP strain distributions


The load-slip curves are compared in Figure 8.13 between the experimental, the
smeared crack modelling and the discrete/smeared crack modelling results. The curves
predicted by the discrete/smeared crack model are slightly stiffer than those predicted by
smeared crack models. But in general, the curves predicted by the two model types are
very similar and in reasonable agreement with the experimental results.

The CFRP strain distributions are compared in Figure 8.14 and Figure 8.15
between the experimental and the discrete/smeared crack modelling results. Since the
numerical model behaviour tends to be stiffer than the actual behaviour, the CFRP strains

304
DISCRETE/SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS AND RETROFITTED BEAMS

are generally lower compared with the measurements. However, the shapes of the
distributions are similar.

25 30
25
Total load (kN)

Total load (kN)


20
20
15
15 T2a (Exp.)
10 T1a (Exp.)
10 T2 (Smeared)
T1 (Smeared)
5 5 T2 (Discrete/smeared)
T1 (Discrete/smeared)
0 0
0.00 0.01 0.02 0.03 0.00 0.02 0.04 0.06 0.08
Slip at 42 mm from loaded edge Slip at 42 mm from loaded edge

(T1) (T2)

30 35
25 30
Total load (kN)

20 Total load (kN) 25


20
15 T3a (Exp.) 15 T4a (Exp.)
10 T3 (Smeared) T4 (Smeared)
10
5 T3 (Discrete/Smeared) 5 T4 (Discrete/smeared)
0 0
0.00 0.05 0.10 0.15 0.20 0.00 0.05 0.10 0.15 0.20 0.25 0.30
Slip at 42 mm from loaded edge Slip at 42 mm from loaded edge

(T3) (T4)

35 35
30 30
Total load (kN)

Total load (kN)

25 25
20 20
15 T5a (Exp.) 15 T6a (Exp.)
10 T5 (Smeared) 10 T6 (Smeared)
5 T5 (Discrete/smeared) 5 T6 (Discrete/smeared)
0 0
0.00 0.10 0.20 0.30 0.40 0.50 0.00 0.10 0.20 0.30 0.40 0.50
Slip at 42 mm from loaded edge Slip at 42 mm from loaded edge

(T5) (T6)
Figure 8.13 Comparison of load-slip behaviours between the experiments, FEA
using smeared crack modelling and FEA using discrete/smeared crack modelling

305
CHAPTER 8

4000 Exp. (kN) FEA (kN)


3500 10.0 10.1
3000 15.0 15.0
Microstrain 2500 18.0 17.7
2000
(T1a) 1500
1000
500
0
0 50 100 150 200
Distance from loaded edge (mm)

2500 Exp. (kN) FEA (kN)


10.0 9.8
2000 15.2 15.2
Microstrain

1500 20.0 20.2


22.0 22.4
(T2a) 1000
500

0
0 50 100 150 200
Distance from loaded edge (mm)

3500 Exp. (kN) FEA (kN)


3000 10.0 9.8
2500 15.0 15.3
Microstrain

2000 20.0 20.4


1500 25.7 25.8
(T3a)
1000
500
0
0 50 100 150 200
Distance from loaded edge (mm)

Figure 8.14 Comparison of CFRP strain distributions in specimens T1a, T2a and
T3a (exp. vs discrete/smeared crack model)

306
DISCRETE/SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS AND RETROFITTED BEAMS

3500 Exp. (kN) FEA (kN)

3000 10.0 10.5


15.0 15.3
2500 20.0 20.3

Microstrain
24.0 24.1
2000 26.0 27.8
1500
(T4a)
1000
500
0
0 50 100 150 200
Distance from loaded edge (mm)

5000 Exp. (kN) FEA (kN)


10.0 10.5
4000 15.0 15.3
20.0 20.3
Microstrain

3000 26.0 29.1


27.8 28.5
(T5a) 2000

1000

0
0 50 100 150 200
Distance from loaded edge (mm)

5000 Exp. (kN) FEA (kN)


10.0 10.5
4000 15.0 15.3
20.0 20.3
Microstrain

3000 24.0 24.1


26.0 26.7
2000 28.0 30.1
(T6a) 26.9 29.0
1000

0
0 50 100 150 200
Distance from loaded edge (mm)

Figure 8.15 Comparison of CFRP strain distributions in specimens T4a, T5a and
T6a (exp. vs discrete/smeared crack model)

8.3.5 Sensitivity study


8.3.5.1 Model sensitivity to concrete tensile strength and mode-I fracture energy
A sensitivity study similar to that for smeared crack modelling described
previously in Section 6.7.3 is presented here for the discrete/smeared crack model of the
shear-lap tests. Different values of the concrete tensile strength, fct, and mode-I fracture
energy, GIf, were adopted and are listed in Table 6.5, where the base model is labelled as

307
CHAPTER 8

‘F3.86-G0.188’. The crack patterns of these models are similar and therefore not shown
herein.

The load-slip curves are compared in Figure 8.16 and Figure 8.17. Similar to the
smeared crack models of the shear-lap tests, the mode-I fracture energy has a significant
influence on the specimen behaviour. The predicted ultimate load capacity reduces as the
fracture energy decreases for both models T5 and T1. The concrete tensile strength seems
to have a less significant effect on the predicted ultimate load model T5 than on model
T1. However, the model with a lower tensile strength shows less stiff behaviour.

T5a (Exp.) T1a (Exp.)


T5 (F3.86-G0.188) T1 (F3.86-G0.188)
T5 (F3.86-G0.124) T1 (F3.86-G0.124)
40 T5 (F3.86-G0.091) 25 T1 (F3.86-G0.091)
Total load (kN)
Total load (kN)

20
30
15
20
10
10 5
0 0
0.00 0.10 0.20 0.30 0.40 0.50 0.00 0.01 0.02 0.03
Slip at 42 mm from loaded edge Slip at 42 mm from loaded edge

(a) T5 (b) T1
Figure 8.16 Comparison of load-slip curves predicted by the models with different
GIf values

T5a (Exp.) T1a (Exp.)


T5 (F4.40-G0.188) T1 (F4.40-G0.188)
T5 (F3.86-G0.188) T1 (F3.86-G0.188)
40 T5 (F2.93-G0.188) 25 T1 (F2.93-G0.188)
Total load (kN)

Total load (kN)

30 20
15
20
10
10 5
0 0
0.00 0.10 0.20 0.30 0.40 0.50 0.00 0.01 0.02 0.03
Slip at 42 mm from loaded edge Slip at 42 mm from loaded edge

(c) T5 (d) T1
Figure 8.17 Comparison of load-slip curves predicted by the models with different
fct values

308
DISCRETE/SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS AND RETROFITTED BEAMS

8.3.5.2 Model sensitivity to mode-II fracture energy


As pointed out before in Section 8.3.2, concrete mode-II fracture energy, GIIf, is
difficult to be determined. In the based model, GIIf was taken to be twice GIf. To study the
effect of this material parameter, different values of GIIf, were adopted and they are listed
in Table 8.4, where the base model is labelled as ‘GII-2’.

The load-slip curves are compared in Figure 8.18. It is clear that since model T1 is
subjected to a shear-tension failure where mode-I is possibly more dominant, the mode-II
facture energy, GIIf, does not influence the load-slip curve significantly. Meanwhile, GIIf
has a significant effect in model T5, especially after the interfacial debond crack starts to
propagate. The model with a lower value for GIIf shows less stiff behaviour.

Table 8.4 Mode-II fracture energy variations


Mode-II fracture energy, GIIf
Designation
( N/mm)
GII-3 0.564
GII-2 0.282
GII-1 0.188

T5a (Exp.) T1a (Exp.)


T5 (GII-3) T1 (GII-3)
T5 (GII-2) T1 (GII-2)
40 T5 (GII-1) 25 T1 (GII-1)
Total load (kN)

Total load (kN)

30 20
15
20
10
10 5
0 0
0.00 0.10 0.20 0.30 0.40 0.50 0.00 0.01 0.01 0.02 0.02 0.03
Slip at 42 mm from loaded edge Slip at 42 mm from loaded edge

(a) T5 (b) T1
Figure 8.18 Comparison of load-slip curves predicted by the models with different
GIIf values

309
CHAPTER 8

8.4 Modelling of retrofitted beams under four-point bending


This section presents the discrete/smeared crack models of the retrofitted beams
tested under four-point bending. The experimental results from these tests are described
in Chapter 5.

8.4.1 Finite element idealisation


As shown in Chapters 5 and 7, two typical failure modes observed in these beams
are end debond and intermediate crack debond. The first mode occurs along the tension
reinforcement level, whereas the second mode initiates from the tip of a flexure-shear
crack under the load point. Based on the previous findings, predefined discrete cracks
were inserted in the FE models as shown in Figure 8.19. They included a peeling crack
from the CFRP end, a flexure-shear crack under the load point and an interface crack
from the tip of the flexure-shear crack. The flexure-shear crack was simplified as
comprising two straight lines and placed between two adjacent stirrups in the shear span
near the load point.

Discrete cracks

Figure 8.19 Discrete crack locations in model E1

The typical finite element mesh and boundary conditions are shown in Figure
8.20. The element size varied from 25x25 mm to 5x5 mm. Due to symmetry, only half a
beam was modelled. The model was supported vertically at the base and horizontally at
the beam’s centreline. The load was applied to a single node on the top of the beam.

310
DISCRETE/SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS AND RETROFITTED BEAMS

Tensile
rebar

Discrete
cracks
Adhesive

CFRP
near the flexure-shear crack tip near the CFRP end
Figure 8.20 A typical mesh of retrofitted beams

The concrete, CFRP, adhesive and steel load plates were modelled using three-
node triangle isoparametric plane stress elements (Element Type 8). The top longitudinal
rebar and stirrups were modelled as embedded reinforcement (ReinfRods). MERLIN
internally determines which continuum elements are crossed by the reinforcement and
properly adjusts the stiffness matrix. The tensile steel reinforcement was modelled using
two-node bar elements (Element Type 2). This allowed the reinforcement to cross the
flexure-shear crack. A perfect bond was assumed between the reinforcement and the
concrete

Similar to the smeared crack models of S beams, separate discrete/smeared crack


models were built for unclamped and clamped side of S beams (for example, two models
‘S1N’ and ‘S1C’ were built corresponding to the unclamped and clamped side of beam
S1a). Figure 8.21 shows a model of the clamped side. However, the clamp was not
prestressed in the discrete/smeared model. It was expected that the prestressing would not
affect the behaviour of the beams up to the peak load significantly.

311
CHAPTER 8

Steel plate Steel bolt

Figure 8.21 Steel clamp location in model S2C

8.4.2 Material models


The CFRP and adhesive material models used in the models were the same as
those of the shear-lap models described in Section 8.3. The concrete material parameters
for E beams were also the same as those used for the concrete blocks (Table 8.2 and
Table 8.3). The concrete material parameters for S beams were chosen in a similar way
and summarised in Table 8.5 and Table 8.6. The steel material was modelled with J2
Plasticity (Material 4) with no hardening (Table 8.7).

Table 8.5 Concrete smeared crack model parameters for S beams


Smeared
Description Unit
crack
Thickness mm 140
Mass density kg/mm3 2.40E-06
Elastic modulus MPa 27300
Poisson's ratio 0.2
Tensile strength MPa 3.51
Fracture energy N/mm 0.188
Compressive strength MPa -57.7
Compressive critical displacement mm -0.5
Factor for return direction in Haigh-Westergaard space 0
Factor for roundness of Menetrey-Westergaard failure surface 0.55
Onset of nonlinearity in compression MPa -38.2
Plastic strain at compressive strength -1.05E-03

312
DISCRETE/SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS AND RETROFITTED BEAMS

Table 8.6 Interface crack model parameters for S beams


Description Unit ICM
Thickness mm 140
Tangential stiffness MPa/mm 44000
Normal stiffness MPa/mm 44000
Tensile strength MPa 3.50
Cohesion MPa 3.85
Friction angle degree 45
Dilatancy angle degree 45
Specific mode-I fracture energy release rate N/mm 0.188
Specific mode-II fracture energy release rate N/mm 0.376
Ratio of irreversible deformation 0.3
Maximum displacement for dilatancy mm 10
Tensile stress at the break-point for bilinear softening law MPa 0.88
Crack opening displacement at the break-point for bilinear
softening law mm 4.03E-02
Cohesion at the break-point for bilinear softening law MPa 0.96
Crack sliding displacement at the break-point for bilinear
softening law mm 7.32E-02

Table 8.7 Steel model parameters


Description Unit N12 Y10 Y6
Elastic modulus MPa 205000 204000 238000
Poisson's ratio 0.3 0.3 0.3
Yield stress MPa 551 334 423

8.4.3 Solution procedure


A similar solution procedure was adopted for the beam FE models. The load was
‘applied’ in increments of 0.1 mm in all models.

313
CHAPTER 8

8.4.4 Verification of finite element model


Verification includes comparing the crack patterns, the loading curves and the
CFRP strain distributions between the experimental measurements with the results from
the smeared crack and discrete/smeared crack models.

8.4.4.1 Crack patterns


The deformed shapes of three typical beams, E1, S2C and S1N, are illustrated
below. It is clear that the failure modes are similar to those observed in the experiments
and predicted by the smeared crack model. Model E1 predicted an end debond failure
(Figure 8.22); model S2C predicted an IS debond failure (Figure 8.23); and model S1N
indicated a combination of both end and IS debond failure (Figure 8.24).

Figure 8.22 Deformed shape of the unclamped side of beam E1 at the peak load
(magnification factor = 40)

314
DISCRETE/SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS AND RETROFITTED BEAMS

Figure 8.23 Deformed shape of the clamped side of beam S2C at the peak load
(magnification factor = 30)

315
CHAPTER 8

Figure 8.24 Deformed shape of the unclamped side of beam S1N at the peak load
(magnification factor = 30)

For completeness, the deformed shapes of other modelled beams are plotted in
Figure 8.25 and Figure 8.26.

316
DISCRETE/SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS AND RETROFITTED BEAMS

(E2)

(E3)

(E4)

(E5)

Figure 8.25 Deformed shapes of the unclamped side of other E models at the peak
loads (magnification factor = 30)

317
CHAPTER 8

(S1C)

(S2N)

(S3N)

(S3C)

Figure 8.26 Deformed shapes of other S models at the peak loads (magnification
factor = 30)

318
DISCRETE/SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS AND RETROFITTED BEAMS

8.4.4.2 Load-deflection curves


The load-deflection curves are compared in Figure 8.27 and Figure 8.28. The
curves predicted by the discrete/smeared crack model are stiffer compared with those
observed in the experiments and predicted by the smeared crack model. It is due to the
fact that a perfect bond was assumed in the discrete/smeared crack model. However, in
general, the curves predicted by the discrete/smeared crack model are in good agreement
with the experiments and the smeared crack model.

319
CHAPTER 8

80 70
70 60
Shear load (kN)

Shear load (kN)


60 50
50
40
40 E2a (Exp.)
E1a (Exp.) 30
30 E2b (Exp.)
E1b (Exp.) 20 E2 (Smeared)
20 E1 (Smeared) E2 (Discrete/smeared)
10 E1 (Discrete/smeared) 10
0 0
0 5 10 15 20 0 5 10 15 20
Midspan deflection (mm) Midspan deflection (mm)

(a) E1 (b) E2

70 80
60 70
Shear load (kN)

Shear load (kN)

50 60
50
40
40
30 E3a (Exp.) E4a (Exp.)
30
20 E3b (Exp.) E4b (Exp.)
E3 (Smeared) 20
10 E4 (Smeared)
E3 (Discrete/smeared) 10
E4 (Discrete/smeared)
0 0
0 5 10 15 20 0 5 10 15 20
Midspan deflection (mm) Midspan deflection (mm)

(c) E3 (d) E4

80
70
Shear load (kN)

60
50
40
30 E5a (Exp.)
E5b (Exp.)
20
E5 (Smeared)
10 E5 (Discrete/smeared)
0
0 5 10 15 20
Midspan deflection (mm)

(e) E5
Figure 8.27 Comparison of load-displacement curves for E beams

320
DISCRETE/SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS AND RETROFITTED BEAMS

90 90
80 80
70 70
Total load (kN)

Total load (kN)


60 60
50 50
S1a (Exp.) S2a (Exp.)
40 S1b (Exp.) 40 S2b (Exp.)
30 S1N (Smeared) 30 S2N (Smeared)
20 S1C (Smeared) 20 S2C (Smeared)
S1N (Discrete/smeared) S2N (Discrete/smeared)
10 S1C (Discrete/smeared) 10 S2C (Discrete/smeared)
0 0
0 10 20 30 40 0 10 20 30 40
Midspan deflection (mm) Midspan deflection (mm)

(a) S1 (b) S2

70
60
Total load (kN)

50
40
S3a (Exp.)
30 S3b (Exp.)
S3N (FE)
20 S3C (FE)
10 S3N (Discrete/smeared)
S3C (Discrete/smeared)
0
0 10 20 30
Midspan deflection (mm)

(c) S3
Figure 8.28 Comparison of load-displacement curves for S beams

8.4.4.3 CFRP and steel reinforcement strain distributions


The predicted strain distributions in the CFRP and tensile steel reinforcement are
compared in Figure 8.29 to Figure 8.34. Since the models predict stiffer behaviour, the
strain levels output from the FE models are lower than the actual values. However, the
general shapes of the curves are comparable to the experimental ones.

321
CHAPTER 8

Load point
4000

3500 Load (kN)


Exp. FEA
3000
20.1 20.4
2500
Microstrain

30.0 30.5
2000 40.0 40.5

1500 50.1 50.0

1000 60.0 60.0


70.7 72.5
500

0
1000 750 500 250 0
Distance from end of CFRP (mm)

Figure 8.29 Comparison of CFRP strain distributions in beam E1a (exp. vs


discrete/smeared crack model)
Load point
12000

Load (kN)
10000
Exp. FEA
20.1 20.9
8000
40.0 39.8
Microstrain

6000 60.0 60.8


70.0 70.1
4000 75.0 75.0
80.2 82.0
2000

0
1000 750 500 250 0
Distance from end of CFRP (mm)

Figure 8.30 Comparison of CFRP strain distributions in beam S2a (exp. vs


discrete/smeared crack model)

322
DISCRETE/SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS AND RETROFITTED BEAMS

Load point
10000
9000
Load (kN)
8000 Exp. FEA
7000 20.0 20.9
Microstrain
6000 40.1 39.7
5000 60.0 60.5
4000 65.0 70.2
3000 70.0 75.6
2000 73.7 80.6
1000
0
1000 750 500 250 0
Distance from end of CFRP (mm)

Figure 8.31 Comparison of CFRP strain distributions in beam S1a (exp. vs


discrete/smeared crack model)

Load point
3500
Load (kN)
3000
Exp. FEA
2500 20.1 20.4
Microstrain

30.0 30.5
2000
40.0 40.5
1500
50.1 50.0
1000 60.0 60.0

500 70.7 72.5

0
1000 750 500 250 0
Distance from end of CFRP (mm)

Figure 8.32 Comparison of tensile steel strain distributions in beam E1a (exp. vs
discrete/smeared crack model)

323
CHAPTER 8

Load point
10000
9000
Load (kN)
8000 Exp. FEA
7000 20.1 20.9
Microstrain

6000 40.0 39.8


5000 60.0 60.8
4000 70.0 70.1
3000 75.0 75.0
2000
80.2 82.0
1000
0
1000 750 500 250 0
Distance from end of FRP (mm)

Figure 8.33 Comparison of tensile steel strain distributions in beam S2a (exp. vs
discrete/smeared crack model)

Load point
10000 Load (kN)
9000 Exp. FEA
8000 20.0 20.9
7000 40.1 39.7
Microstrain

6000 60.0 60.5


5000 65.0 70.2
4000 70.0 75.6
3000 73.7 80.6
2000
1000
0
1000 750 500 250 0
Distance from end of CFRP (mm)

Figure 8.34 Comparison of tensile steel strain distributions in beam S1a (exp. vs
discrete/smeared crack model)

324
DISCRETE/SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS AND RETROFITTED BEAMS

8.4.5 Sensitivity study


8.4.5.1 Model sensitivity to concrete tensile strength and mode-I fracture energy
A sensitivity study similar to that for smeared crack modelling reported previously
in Section 7.2.5.1 is presented here for the discrete/smeared crack model of the beam
tests. Different values of the concrete tensile strength, ft, and mode-I fracture energy, GIf,
were adopted and they are listed in Table 7.3.

The load-displacement curves are compared in Figure 8.35 and Figure 8.36. The
mode-I fracture energy has a more significant effect on model E1 than model S2C. As
expected, the higher the energy, the higher the predicted load capacity. A similar
observation can also be made for the effect of concrete tensile strength, ft.

E1a (Exp.) S2a (Exp.)


E1b (Exp.) S2b (Exp.)
E1 (F3.86-G0.188) S2C (F3.51-G0.188)
E1 (F3.86-G0.124) S2C (F3.51-G0.117)
100
80 E1 (F3.86-G0.091) S2C (F3.51-G0.086)
70 80
Total load (kN)
Total load (kN)

60
50 60
40 40
30
20 20
10
0 0
0 5 10 15 0 10 20 30 40
Midspan deflection (mm) Midspan deflection (mm)

(a) E1 (b) S2C


Figure 8.35 Comparison of load-displacement curves predicted by the models with
different GIf values

325
CHAPTER 8

E1a (Exp.) S2a (Exp.)


E1b (Exp.) S2b (Exp.)
E1 (F4.40-G0.188) S2C (F4.14-G0.188)
E1 (F3.86-G0.188) S2C (F3.51-G0.188)
100
80 E1 (F2.93-G0.188) S2C (F2.76-G0.188)
70 80

Total load (kN)


Total load (kN)

60
50 60
40
40
30
20 20
10
0 0
0 5 10 15 0 10 20 30 40
Midspan deflection (mm) Midspan deflection (mm)

(a) E1 (b) S2C


Figure 8.36 Comparison of load-displacement curves predicted by the models with
different fct values

8.4.5.2 Model sensitivity to mode-II fracture energy


Concrete mode-II fracture energy, GIIf, was varied in a similar manner as
described in section 8.3.5.2. The values adopted for GIIf are listed in Table 8.4, where the
base model is labelled as ‘GII-2’.

The load-displacement curves are compared in Figure 8.37. It is clear that the
mode-II fracture energy has a more significant influence on the behaviour of model S2C
than on model E1. The predicted beam capacity increases with the mode-II fracture
energy.

326
DISCRETE/SMEARED CRACK MODELLING OF SHEAR-LAP SPECIMENS AND RETROFITTED BEAMS

E1a (Exp.) S2a (Exp.)


E1b (Exp.) S2b (Exp.)
E1 (GII-3) S2C (GII-3)
E1 (GII-2) S2C (GII-2)
100
80 E1 (GII-1) S2C (GII-1)
70 80

Total load (kN)


Total load (kN)

60
50 60
40 40
30
20 20
10
0 0
0 5 10 15 0 10 20 30 40
Midspan deflection (mm) Midspan deflection (mm)

(a) E1 (b) S2C


Figure 8.37 Comparison of load-displacement curves predicted by the models with
different GIIf values

8.5 Summary of findings


In this chapter, nonlinear finite element analyses using discrete/smeared cracking
of a number of the shear-lap and beam tests have been presented. The findings from this
work are:

• The discrete/smeared crack models were able to simulate the behaviour of shear-lap
tests reasonably well. The predicted peak load, load-slip curves and CFRP strain
distributions were similar to those obtained from the smeared crack models. The
failure mechanisms were simulated relatively well showing two main failure modes:
interfacial debond and shear-tension failure.
• The discrete/smeared crack models were also able to simulate the behaviour of beam
tests relatively well. The simulation results were similar to those from the
corresponding smeared crack models. The modelling results confirmed the failure
mechanism of end and intermediate span debond. End debond was the shearing of
concrete along the tension reinforcement level; whereas intermediate span debond
was the shearing of concrete along the bond surface from the tip of a diagonal crack.

327
DEBOND STRENGTH MODELS AND RELIABLITY ANALYSIS FOR RETROFITTED BEAMS

CHAPTER 9 - DEBOND STRENGTH MODELS AND


RELIABLITY ANALYSIS FOR RETROFITTED BEAMS

9.1 Introduction and scope


There are a number of theoretical models available to predict the capacity of
shear-lap specimens and retrofitted beams and these models have been assessed in
Chapter 3. It was found that while some prediction models for the bond capacity of shear-
lap specimens show reasonably good results, most prediction models for the debonding
capacity of retrofitted beams produce relatively inaccurate and scattered results. Most of
the models for retrofitted beams are also relatively complex and not suitable for design
practice.

Therefore, in this chapter, an attempt is carried out to develop and verify design
models to calculate the strength of retrofitted beams. Three most frequently observed
failure modes are to be addressed: flexure failure, intermediate span debond and end
debond. The models are developed based on the observations from the main experimental
investigations and from the numerical analyses. Even though the models are derived
based on the data of beams retrofitted with CFRP, they are expected to be valid for beams
retrofitted with other available composites. A reliability study is also carried out for these
models to recommend appropriate capacity reduction factors for design.

9.2 Analysis of retrofitted sections


9.2.1 Implementation of beam theory
To calculate the strains and stresses in the component materials of a retrofitted
section subjected to an applied moment, the beam theory was utilised. The basic
assumption of the theory is that the strain varies linearly along the section from the top
fibre to the bottom fibre. The strain distribution is illustrated in Figure 9.1. The theory
and the adopted constitutive properties of component materials are described in Appendix
A.1.

329
CHAPTER 9

εc0 εc
ε’s0 ε’s

εs0 εs

εfrp

(a) (b)
Figure 9.1 Strain distribution before (a) and after installation (b) of FRP

The analysis algorithm is illustrated in Figure 9.2. Two loops were used to
estimate the location of the neutral axis, d, and the concrete strain at the top fibre, εc. The
implementation was executed using VBA in Microsoft Excel 2002.

330
DEBOND STRENGTH MODELS AND RELIABLITY ANALYSIS FOR RETROFITTED BEAMS

START

Input section dimensions,


material properties and M

No
M > Mcr
Yes

Initial estimate of εc

Initial estimate of d

Calculate unbalanced force ∆F


Uncracked
section analysis
Estimate of d to minimise ∆F

No
∆F < ∆F tolerance

Yes
Calculate unbalanced moment ∆M

Estimate of εc to minimise ∆M

No
∆M < ∆M tolerance

Yes

Output d, εc, εf, εs

STOP

Figure 9.2 Flow chart for implementation of sectional analysis

331
CHAPTER 9

9.2.2 Average strain in steel and FRP reinforcement


In the sectional analysis described in section 9.2.1, the strain levels in the steel and
FRP are calculated with the assumption that if the applied moment, M, is greater than the
cracking moment, Mr, the section is cracked and concrete is effective only in
compression. This is only true at the crack locations. A better indication of the strain
distribution along the reinforcement of an RC beam is the average strain, of which the
concept is illustrated in Figure 9.3. The model to compute the average steel strain has
been reported in the CEB Manual – Cracking and Deformation (Comite Euro-
International du Beton, 1985). It is summarised again below for a member under pure
bending. It is assumed that the model is also applicable for FRP reinforcement.

average
actual steel εs2 steel strain
strain εsm

bond stress

Figure 9.3 Local variations in steel strain and bond stress

Figure 9.4 illustrates the calculation model for a member subjected to a constant
moment M. To find the average strain, the actual element is replaced by a model
composed of two parts. Part I works as an uncracked section. Part II works as a cracked
section where only the reinforcement and the concrete in compression are taken into
account. The average strain is defined by:

332
DEBOND STRENGTH MODELS AND RELIABLITY ANALYSIS FOR RETROFITTED BEAMS

∆l ∆l1 + ∆l 2 l1 ε s1 + l 2 ε s2
ε sm = = = = (1 − ζ)ε s1 + ζε s2 (9.1)
l l l

where εs1 is the strain in the reinforcement calculated assuming the section is uncracked.
εs2 is the strain in the reinforcement calculated assuming the section is cracked and
neglecting the contribution of concrete in tension. ζ is a distribution coefficient which can
be calculated using the following formula:

2
⎛M ⎞
ζ = 1 − β1β 2 ⎜ cr ⎟ (9.2)
⎝ M ⎠

β1 is a coefficient characterising the bond quality of the reinforcement, which can be


taken as 1 for high bond bars. β2 is a coefficient representing the influence of the duration
of application or of repetition of loading, which can be taken as 1 for the first static
loading.

II II
I
M M
εs2 εs1 εs2

l2/2 l1 l2/2
l

(a) Actual element (b) Model

Figure 9.4 Calculation model for a RC member under pure flexure (Comite Euro-
International du Beton, 1985)

9.2.3 Verification and recommendation


The beam theory has been verified previously in Chapter 3 and it has been found
to be able to predict the ultimate load bearing capacity of retrofitted beams failing by
concrete crushing or FRP rupture with reasonable accuracy. The aim of this section is to
verify the application of the beam theory to compute the reinforcement strain levels at
any applied moment level, which is essential for the models proposed later for debonding

333
CHAPTER 9

failure. This verification study was based on the measured strain values in the
experiments described in Chapter 5.

To plot the strain distributions along the beam, sectional analyses were carried out
at 50 locations along the beams at several load levels. The average strain levels were
calculated according to Equation 9.1.

The calculation results for two typical beams, E1a and S1a, at the ultimate load
are shown in Figure 9.5 and Figure 9.6, respectively. In these figures, M is the applied
moment, Fs is the force in the tension reinforcement, Ff is the force in the FRP and dn is
the neutral axis depth from the section top. The figures clearly demonstrate the influence
of steel yielding on the Ff distribution. In beam S1a, once the steel has yielded, there is a
steep increase in the Ff slope and a gradual decrease in the neutral axis depth.

End of FRP Start of


concrete
Load point cracking
200
Fs (kN)
150
dn (mm)
100 F (kN)
f

50
M (kN.m)
0
1000 750 500 250 0
Distance from left support (mm)

Figure 9.5 Sectional analysis results along beam E1a at maximum load level

334
DEBOND STRENGTH MODELS AND RELIABLITY ANALYSIS FOR RETROFITTED BEAMS

Start of
steel Start of
yielding End of FRP concrete
Load point cracking
200
Fs (kN)
150

100 dn (mm)
Ff (kN)
50
M (kN.m)
0
1000 750 500 250 0
Distance from left support (mm)

Figure 9.6 Sectional analysis results along beam S1a at maximum load level

The CFRP strain distributions predicted by the beam theory at different load
levels along the whole beam for E1a and S1a are compared with the corresponding
experimental curves in Figure 9.7 and Figure 9.8, respectively. Comparisons for other
beams are presented in Appendix D.4. The plots show that the beam theory can predict
the peak CFRP strain level (in the constant moment region) with good accuracy at both
low and high load levels. However, the predicted CFRP strains in the shear span deviate
significantly from the measured values at high loads. Deviation occurs most clearly when
the steel reinforcement undergoes yielding. For those cases, the actual CFRP strain shows
a more gradual decline in the shear span (Figure 9.8). This finding was also observed in
the numerical models presented previously in Chapters 7 and 8. This was due to presence
of diagonal cracks which led to strain redistribution in the concrete cover.

335
CHAPTER 9

Load point
4000
3500 Load (kN)
Exp. Theory
3000
20.1 20.0
2500
Microstrain

40.0 40.0
2000 60.0 60.0
1500 70.7 70.7

1000
500
0
1000 750 500 250 0
Distance from end of FRP (mm)

Figure 9.7 Comparison of FRP strain distributions between experimental results


and predictions by beam theory in beam E1a

Load point
10000 Load (kN)
9000 Exp. Theory
8000
20.0 20.0
7000
40.1 40.0
Microstrain

6000
5000 60.0 60.0
4000 65.0 65.0
3000 70.0 70.0
2000
73.7 73.8
1000
0
1000 750 500 250 0
Distance from end of FRP (mm)

Figure 9.8: Comparison of FRP strain distributions between experimental results


and predictions by beam theory in beam S1a

Since the applied moment, M, under the load point is significantly higher than the
cracking moment, Mr, at ultimate, the distribution coefficient, ζ, could be taken to be
unity at ultimate and the maximum CFRP strain could be calculated assuming a cracked
section. As illustrated in Figure 9.9, the CFRP strain levels at ultimate are predicted
accurately using this method.

336
DEBOND STRENGTH MODELS AND RELIABLITY ANALYSIS FOR RETROFITTED BEAMS

12000 S beams E beams

10000
Theory (no reduction)
CFRP microstrain
8000 Exp. (at midspan)
Exp. (under load point)
6000
Exp. (between midspan and load point)
4000

2000

E1a

E2a

E3a

E4a

E5a
E1b

E2b

E3b

E4b

E5b
S1a

S2a

S3a
S1b

S2b

S3b
Figure 9.9 Comparison of measured and calculated CFRP strains at the ultimate
load level for retrofitted beams without anchorage

9.3 Debond strength models for retrofitted beams


9.3.1 Model derivation
Two debond models presented below are based on the failure mechanisms
depicted in Figure 9.10.

Ff,0 Ff,i Ff,i+1

Peeling crack Peeling crack

Figure 9.10 Failure mechanisms

End debond model:


End debond is the result of the pulling action of the tensile force Ff,0 on the FRP
end. The tensile force leads to a high shear stress level at the weakest layer near the

337
CHAPTER 9

tension reinforcement level, inducing a longitudinal debonding crack there. Therefore, it


can be envisaged that end debond occurs when the shear stress in the weakest layer
reaches a limiting value. This concept was first introduced in Mukhopadhyaya and
Swamy (2001). The researchers stated that the interfacial shear stress value between the
composite and concrete could be used as a simple and practical design criterion to predict
the failure at the adhesive-concrete interface. However, it was pointed out before that the
critical failure surface is not at the adhesive-concrete interface but at the concrete layer
near tension reinforcement. Nevertheless, the interfacial shear stress is closely related to
the stress along the weakest plane since the concrete cover is generally small compared to
the CFRP bonded length.

Strictly speaking, the tensile force in the composite is transferred through the
shear stress on the concrete area at the weakest plane (Figure 9.11a, surface A) and the
bond area between concrete and steel rebars (Figure 9.11a, surface B). The shear stress is
also not uniform over the failure surface. However, surface B is generally small compared
to surface A and the bond strength between the steel rebars and the concrete cover
(unconfined concrete) is also possibly small compared to the concrete shear strength.
Therefore, for simplification, the average shear stress is assumed to be distributed over
the failure surface A only. End debond can be assumed to occur when the following
inequality is satisfied:

τave ≥ fcv,d (9.3)

where τave is the average shear stress on the assumed debond surface (surface A) and fcv,d
is the shear strength of the surface. To calculate τave, the force distribution in the
composite plate can be assumed to be linear over the shear span based on the observation
from the experiments described previously (Figure 9.11b). As a result, the average shear
stress can be estimated to be:

τave = Ff,max / Lf bf (9.4)

where Ff,max is the tensile force in the FRP under the load point; Lf is the FRP bond length
in the shear span; and bf is the FRP width.

338
DEBOND STRENGTH MODELS AND RELIABLITY ANALYSIS FOR RETROFITTED BEAMS

Surface A Surface B

(a)

Lf

FRP

Composite force Ff,max


distribution

Average shear stress τave

(b)
Figure 9.11 Average shear stress concept

It is difficult to determine the shear strength at the tension reinforcement layer,


fcv,d. Limited experimental data is available on the shear strength of plain concrete when
the normal stress is negligible (which is the case here). In addition, the strength is also
affected by a number of factors including the stress state in the longitudinal steel bars, the
location of the stirrups and the location of flexural and flexure-shear cracks. However, it
can be assumed that the shear strength is proportional to the square root of the concrete
compressive strength (Standards Australia, 2001). Calibration using the strain
measurements for 10 beams tested in this study showed that the limiting average stress at
which end debond occurred, fcv,d, was 1.22 MPa, which was 0.17 f c . This is

demonstrated in Figure 9.12a.

339
CHAPTER 9

1.6 80
1.4 70
1.2 60
1.0 50
MPa

0.8

kN
40
0.6 Average bond stress 30
0.4 Average value 20 Tensile force in CFRP
0.2 10 Average value
0.0 0
E1a

E2a

E3a

E4a

E5a
E1b

E2b

E3b

E4b

E5b

S1a

S2a

S3a
S1b

S2b

S3b
(a) (b)
Figure 9.12 Variation of maximum average bond stress in beams failed by end
debond (a) and maximum tensile force in CFRP in beams failed by intermediate
span debond or mixed debond (b)

Intermediate span debond model:


Intermediate span debond is the result of the high tensile force, Ff,i, in the FRP at
the tip of a main flexural crack or flexure-shear crack (Figure 9.10). Peeling occurs in the
concrete substrate next to the bond surface and propagates away from the initial location.
When peeling occurs at the tip of a flexure-shear crack, cracking near the tension
reinforcement level also occurs as the tension is transferred to the adjacent concrete tooth
formed between the two flexure-shear cracks. For both cases, the maximum tensile stress
that the composite can take depends on the shear strength of the concrete substrate. This
strength can be determined using the results from shear-lap testing. This observation was
first made by Teng et al. (2001). They calculated the maximum pulling force in the FRP
in a shear-lap test using Chen and Teng’s formula (Chen and Teng, 2001). As also
pointed out in Chapters 3 and 4, Chen and Teng’s formula produced good predictions for
the bond capacity of shear-lap specimens. It is also recommended here that their formula
can be used to calculate the maximum tensile force in the FRP, as follows

P = αβ f β L f c b f L e (9.5)

where the factor, α, is a calibration factor to account for any difference between a beam
failing by intermediate span debond and a shear-lap specimen failing by interfacial
debond. Despite many similarities, the situation in a retrofitted beam is much more

340
DEBOND STRENGTH MODELS AND RELIABLITY ANALYSIS FOR RETROFITTED BEAMS

complex due to the presence of several factors such as bending deformation, shear
deformation and the presence of tensile steel reinforcement. Teng et al. calibrated a with 8
beams and 9 slabs and suggested a value of 0.4 for α. A similar calibration study was
carried out using the experimental results reported in Chapter 5. The study showed that
the maximum CFRP tensile force was approximately 64.4 kN, and the corresponding
value for the factor α was 1.04. This is demonstrated in Figure 9.12b.

9.3.2 Verification against experiments


The proposed models were first verified against the beams tested by the author.
They included 2 trial tests (B1 and B2), 16 main tests (S1 to S3, E1 to E5), 1 retest with
no anchorage (E3b2), 2 retests with one end anchorage (E1a2 and E3a2), and 2 tests with
one end anchorage (A1a and A2a). The models were also verified against the beam
database as described in Chapter 3. From these verification studies, it was found that
better predictions were achieved if the concrete shear strength, fcv.d, was taken as 0.2 f c .

It was also found that the factor α can be taken to be 1.0 for simplification without
altering the accuracy significantly.

With these modifications, the first verification results are shown in Figure 9.13.
As indicated by the predicted curves, beams E1a to B2 fail by end debond and beams S1
to B1 fail by intermediate span debond. These predictions are consistent with the actual
failure modes observed in the experiment. For beams S1a to S3b, the two predicted
curves are relatively close suggesting a mixed mode of end and intermediate span
debond, which was also seen the experiment. For the last 4 beams, A2a to E3a, the
anchorage improved the strength significantly by preventing end debond and therefore
forcing intermediate span debond. For these beams, the predicted failure loads are slightly
higher than those observed in the experiments.

341
CHAPTER 9

end & IS IS debond (end


end debond debond anchored)

160
140
Failure shear load (kN)

120
100
80
60
40 Exp.
Theory - No debond
20 Theory - End debond
Theory - IS debond
0
E3b2

E1a2
E3a2
E1a

E2a

E3a

E4a

E5a
E1b

E2b

E3b

E4b

E5b

B2

B1
S1a

S2a

S3a
S1b

S2b

S3b

A2a
A1a
Beam label

Note: No debond: full sectional capacity; IS debond: intermediate span debond

Figure 9.13 Prediction of end and intermediate span failure loads for the beams
tested in the present study

The results from the second verification study using the beam database are shown
in Figure 9.14, which also indicates good agreement. The averages of the ratios of
predicted failure loads on the experimental results Vexp/Vcal are 1.01 and 1.05 for
intermediate span and end debond, respectively. The coefficients of variation are 14 and
20 %, respectively. For the end debond strength model (Figure 9.14a), the predicted
values are quite conservative for a number of beams. The reason is that these beams were
reinforced with a relatively large amount of tension reinforcement (ku ≈ 0.4) and therefore
the bond area of the tensile rebars could be significant.

342
DEBOND STRENGTH MODELS AND RELIABLITY ANALYSIS FOR RETROFITTED BEAMS

160 350 FSCD


FCD
SCD
300
120
Vpredicted (kN)
250

Vcal (kN)
80 200

150
40 100
50
0
0 40 80 120 160 0
0 50 100 150 200 250 300 350
Vexp (kN) Vexp (kN)

(a) End debond (b) Intermediate span debond

Note: FSCD: flexure-shear crack debond; FCD: flexural crack debond; SCD: shear crack debond

Figure 9.14 Comparison between the theoretical predictions and experimental


results

9.3.3 Parametric study


In this section, the predicted behavioural trends of the capacity of beams with
different CFRP thicknesses, tensile steel amounts, CFRP bond length and concrete covers
on the debonding capacity are presented. Comparison is made between the experimental
results, the predictions by the FEA using smeared cracking and the predictions by the
proposed theoretical models. Similar to section 7.2.6, the study is carried out on the base
specimen similar to beam E1a.

Figure 9.15 compares the behavioural trends as the CFRP thickness varies. The
theoretical results include predictions for intermediate span debond (IS debond), end
debond, flexural failure assuming no debond (or full composite action) and flexural
failure of the original beams without CFRP. There predicted trends are very similar to
those observed from the FEA. As the CFRP area increases beyond an optimal value of
approximately 50 mm2, the predicted failure mode changes from intermediate span
debond to end debond (as the intermediate span debond capacity increases to a value
higher than the end debond capacity).

343
CHAPTER 9

Failure mode
predicted by IS debond Mixed
FEA Crush End debond
Ultimate shear load (kN) debond
100

80

60
Exp.
40 FE
Theory - IS debond
20 Theory - End debond
Theory - No debond
Theory - No FRP
0
0 50 100 150 200
2
A f (mm )

Figure 9.15 Comparison of the predicted trends with different CFRP thicknesses

Figure 9.16 compares the behavioural trends as the tension reinforcement area
varies. A good agreement can also be seen between the trends predicted by the theoretical
models and the FEA. Shifting of the failure mode from intermediate span debond to end
debond is observed as the steel reinforcement amount increases.

Failure mode
predicted by IS debond Mixed debond End debond
FEA
100
Ultimate shear load (kN)

80

60
Exp
40 FE
Theory - IS debond
20 Theory - End debond
Theory - No debond
Theory - No FRP
0
0 50 100 150 200 250 300 350 400
2
A s (mm )

Figure 9.16 Comparison of the predicted trends with different tensile steel amounts

The behavioural trends as the CFRP bond length varies are depicted in Figure
9.17. The trend predicted by the theoretical model matches well with that predicted by the
FEA.

344
DEBOND STRENGTH MODELS AND RELIABLITY ANALYSIS FOR RETROFITTED BEAMS

Critical unbonded
section near CFRP
end
Failure mode
predicted by End debond IS debond
FEA
120
Ultimate shear load (kN)
100

80

60
Exp.
FE
40 Theory - IS debond
Theory - End debond
20 Theory - No debond
Theory - No FRP
0
0.50 0.60 0.70 0.80 0.90 1.00
Lf / a

Figure 9.17 Comparison of the predicted trends with different CFRP bond lengths

As illustrated in Figure 9.18, the theoretical models also indicate that concrete
cover has an insignificant influence on the debond capacity of retrofitted beams.

Failure mode
predicted by End debond
FEA
Ultimate shear load (kN)

120

100

80

60
Exp.
40 FE
Theory - IS debond
Theory - End debond
20 Theory - No debond
Theory - No FRP
0
10 20 30 40 50
c (mm)

Figure 9.18 Comparison of the predicted trends with different concrete covers

9.4 Reliability study


9.4.1 Introduction
This section presents a reliability study of the strength of CFRP-retrofitted beams.
Reliability-based techniques have been used to account for the randomness of important

345
CHAPTER 9

variables affecting the strength of structures. The development of these techniques has
been described in Ellingwood et al. (1980).

One of the earliest studies of the reliability of concrete structures strengthened


with FRP was conducted by Plevris et al. (1995). They used the Monte Carlo method to
study the flexural reliability of RC beams strengthened with CFRP assuming full
composite action. The use of a reduction factor for CFRP material strength together with
a general resistance factor for overall member flexural strength was proposed. A large
number of design cases (1024 cases) were considered, taken from two main nominal
configurations. The nominal beams had rectangular sections. The study indicated that the
general strength reduction factor, φ, and a partial reduction factor, φCFRP, could be taken
to be 0.85 and 0.95, respectively. If a partial reduction factor was not considered (φCFRP =
1.0), the study recommended that φ could be taken as 0.8.

A more recent publication on the reliability of RC bridge girders strengthened in


flexure assuming no debonding was by Okeil et al. (2002). They focused on reliability of
RC bridge girders. Twelve design cases were considered. Monte Carlo simulation was
carried out to study the statistical properties of the resistance as calculated using a fibre
section method. Okeil et al recommended that the resistance reduction factor, φ, could be
taken as 0.85.

There are several reasons to conduct further reliability study on retrofitted beams.
Firstly, previous studies on retrofitted beams did not consider the model error or assumed
it to be the same as for normal RC beams. This was due to unavailability of verification
data for retrofitted beams. Secondly, more tests and analysis have become available
recently allowing more reliable estimation of material variability, especially for
composites. Thirdly, no study on the reliability of CFRP retrofitted RC beams has
considered debonding failures. The main reason for that was the lack of a reliable
theoretical model for these modes.

Presented below is another study on the reliability of RC members strengthened


with CFRP materials based on the models described in sections 9.2 and 9.3. In the study,
the deficiencies mentioned in the previous paragraph were addressed.

346
DEBOND STRENGTH MODELS AND RELIABLITY ANALYSIS FOR RETROFITTED BEAMS

9.4.2 Scope and procedure


The goal of this study is to study the statistics of resistance and reliability of
simply supported RC beams retrofitted with CFRP.

The following steps were carried out in this study as guided by Ellingwood et al.
(1980):
• Calculate the bias and variability of the computational procedures.
• Choose a series of representative beam configurations, defined by nominal material
strengths and nominal dimensions. Calculate the nominal resistance, Rn, using the
developed computational procedures.
• Generate a set of material strengths and dimensions randomly from the statistical
distributions of each variable. Calculate the theoretical capacities, R, using the set of
variables plus randomly generated values of the model error. Calculate the mean and
coefficient of variation of R.
• Use the statistical properties of R and select a target reliability index, β, to
recommend the resistance factor for the models.

Several beam configurations were analysed. The beams were assumed to fail in
only one of three following modes:
• Flexural failure: This mode was possible for all configurations. Other failure
mechanisms were assumed to be avoided through the use of end and intermediate span
anchorage.
• Intermediate span debond: This mode was likely when the sectional capacity is
moderately higher than the debond capacity. End debond, if it was more critical, was
assumed to be avoided by end anchorage.
• End debond: This mode was less likely compared to other two modes for large beams
bonded with a small amount of CFRP reinforcement. Intermediate span debond, if it was
more critical, was assumed to be avoided by intermediate span anchorage.

The procedures were followed separately for each failure mode, namely flexural
failure, intermediate span debond and end debond. One reason to adopt this procedure
was the complexity of the computational procedures. If the resistance could be calculated

347
CHAPTER 9

using ‘simple’ formulae, the beam variables would be able to be included in the limit-
state functions for each failure modes to compute their reliability index and a
stochastically most relevant mode would be found. Another reason was that this method
allowed determination of separate resistance factors for each case and therefore different
model errors could be incorporated.

9.4.3 Selection of variables and their variability


9.4.3.1 Design of cross sections
Three beam configurations were chosen for simulation. The dimensions were
taken from a database of bridge dimensions in Victoria, Australia. They correspond to
three spans: 4.88, 7.92 and 11.89 m. The cross sections and existing reinforcement are
shown in Figure 9.19. They were retrofitted with CFRP bonded on the soffit. The CFRP
width was taken to be 300 mm for all beams. The concrete compressive strength was
taken to be 32 MPa. The yield stress of the steel reinforcement was 300 MPa.

348
DEBOND STRENGTH MODELS AND RELIABLITY ANALYSIS FOR RETROFITTED BEAMS

1800
Cross section T1:
Tension reinforcement: 6Y22

180
692
Compression reinforcement: 2Y22
Span: 4.88 m
346

1800
Cross section T2:
Tension reinforcement: 4Y28 and 4Y26
Compression reinforcement: 2Y22
180
743

Span: 7.92 m

371

1800
Cross section T3:
Tension reinforcement: 10Y32
180

Compression reinforcement: 2Y22


870

Span: 11.89 m

435

All dimensions are in mm.


Figure 9.19 Cross sections of T beams used

The strength models developed for retrofitted RC beams were dependant mainly
on the sectional dimensions and bending moment in the section. For the debonding
models, the shear span was the only additional input for the model. For simplicity, the
beams were loaded in four-point bending only even though, in practice, bridge girders are
under more complex loading pattern mainly from dead load and traffic load.

For each beams, two loading shear spans and two applied CFRP thicknesses were
considered. This led to 12 designed retrofitted cases. For all beams, the CFRP terminated
200 mm from the support. All beams were assumed to have a sufficient shear capacity.
The beam configurations are summarised in Table 9.1.

349
CHAPTER 9

Figure 9.20 Side view of T beams and loading positions

Table 9.1 Design summary of T beam configurations


Shear shear span
Label Span (m) tf (mm) ku Mu,frp / Mu,0
span (m) / depth
T1-C0 4.88 - - 0.055 - 1.00
T1-C1a 4.88 0.352 1.63 0.080 2.35 1.38
T1-C1b 4.88 0.352 0.98 0.080 1.41 1.38
T1-C2a 4.88 0.704 1.63 0.092 2.35 1.83
T1-C2b 4.88 0.704 0.98 0.092 1.41 1.83
T2-C0 7.92 - - 0.063 - 1.00
T2-C1a 7.92 0.528 2.64 0.097 3.55 1.47
T2-C1b 7.92 0.528 1.58 0.097 2.13 1.47
T2-C2a 7.92 1.056 2.64 0.112 3.55 1.95
T2-C2b 7.92 1.056 1.58 0.112 2.13 1.95
T3-C0 11.89 - - 0.083 - 1.00
T3-C1a 11.89 0.704 3.96 0.112 4.56 1.29
T3-C1b 11.89 0.704 2.38 0.112 2.73 1.29
T3-C2a 11.89 1.408 3.96 0.129 4.56 1.62
T3-C2b 11.89 1.408 2.38 0.129 2.73 1.62

9.4.3.2 Model errors


For the analysis of reinforced concrete beam sections, Ellingwood et al. stated that
the error was 4.6 %. In Sanjayan and Candy (2004), a value of 5 % was taken. For
retrofitted RC beams, the model errors are estimated below as guided by Ellingwood et
al.

350
DEBOND STRENGTH MODELS AND RELIABLITY ANALYSIS FOR RETROFITTED BEAMS

The mean was taken as the average value of the ratio of the test strength divided
by calculated strength Rtest / Rcalc. The variability was given by

Vm = VT2 / C − Vtest
2
− Vspec
2
(9.6)

where VT/C is the coefficient of variation obtained directly from the comparison of
measured and calculated strength; Vtest represents the uncertainties in the measured loads
due to such things as the accuracies of the gages and errors in readings; and Vspec
represents the errors introduced by variations in the specimen material strengths and
dimensions. The typical values for Vtest are about 2 to 4 % and the values for Vspec are
about 4 % (Ellingwood et al., 1980). In the calculation for VT/C, it was assumed that the
verification results as described in Section 9.3.2 for mainly small to medium rectangular
beam sections were also valid for relatively large T beam sections.

The final results are shown in Table 9.2. It was clear that the variations of the
models proposed for three dominant failure modes were much higher than those of
normal RC beams. The variations are 14.8 %, 12.7 % and 18.9 % for the models to
predict the full composite action, intermediate span debond and end debond, respectively.

Table 9.2 Calculation of model errors


Full composite Intermediate
Parameter End debond
action span debond
Mean 0.972 1.010 1.053
Coefficient of variation VT/C 0.156 0.136 0.195

Standard deviation 0.152 0.138 0.206


Maximum 1.288 1.312 1.689
Min 0.650 0.620 0.628
Vtest 0.030 0.030 0.030
Vspec 0.040 0.040 0.040
Vm 0.148 0.127 0.189

351
CHAPTER 9

The model errors were assumed to follow normal distributions.

9.4.3.3 Concrete, steel yield strength and geometry


Statistical parameters for the concrete compressive strength, steel yield strength
and geometry are provided in Table 9.3 as guidance from the CONTECVET report
(CONTECVET, 2002).

Table 9.3 Statistical parameters for some random variables from CONTECVET
(2002)
CONTECVET report
Variable
Distribution COV (%)
Concrete compressive strength Log-normal 10-30
Steel yield strength Log-normal 5-10
Geometry Normal 5

Sanjayan and Candy (2004) tested six core specimens in their reliability study of a
bridge in Australia (Baandee Lakes Bridge No. 1049). They found that the variation of
concrete compressive strength was 25 %. The normal quoted range for the strength is
from 15 to 21 % (Ellingwood et al., 1980; Holicky and Markova, 2000). In this study, the
variation was chosen to be 20 %, which is the average value of the limits given in
CONTECVET report. It was assumed to have a log-normal distribution.

Some tests on the variation of steel yield strength are available in Sanjayan and
Candy (2004) and Sonnenberg and Boully (2004). In the first study, nine tensile tests
were performed. The variation found was 7.3 %. In the second study, steel reinforcement
samples were taken from seven old pre 1960 bridges in Victoria, Australia. They found
that the variations ranged from 2.2 % to 10.6 %. In this study, a value of 7.5 % was used,
which is the average value of the limits given in CONTECVET report. The steel yield
strength was assumed to have a log-normal distribution.

For geometrical dimensions, standard deviation is a more appropriate form to


express the variations. Ellingwood et al (1980) reported two values of 6.4 and 14.0 mm,
which were taken from Mirza and MacGregor’s work (Mirza and MacGregor, 1979). A

352
DEBOND STRENGTH MODELS AND RELIABLITY ANALYSIS FOR RETROFITTED BEAMS

value of 7.1 mm was recommended later in Holicky and Markova (2000). In this study,
the standard deviation was taken to be 10 mm.

9.4.3.4 CFRP
In this study, only the variability in CFRP tensile strength is considered. The
composite stiffness and geometrical dimensions are assumed to be deterministic.

CFRP strength variation is commonly described using a Weibull distribution


(Harlow and Phoenix, 1981; Batdorf and Ghaffarian, 1984; Plevris et al. 1995; Okeil et
al., 2001). The shape parameter, m, and scale parameter, σ, of the distribution can be
related to the mean, µ, and coefficient of variation, COV, through the following
expressions (Okeil et al., 2001):

µ≈σ (9.7)

COV ≈ 1.2 / m (9.8)

Okeil et al. (2002) used a value of 2.2 % for the COV and 1.10 for the bias in their
work. This based on the results from tensile testing of CFRP bars reported in Bakht et al.
(2000). Ayers and Van Erp (2002) reported an experimental program to investigate the
strength and stiffness variations of E-glass laminates formed from unidirectional fibres.
The strength reported was ‘Normalised Unit Strength’ or ultimate force per unit weight
per mass of reinforcement. They found that the variability of lamina strengths ranged
from 2.27 % to 15.34 %. A recent study on the variation of CFRP retrofitting materials
formed in the field by a wet lay-up method was by Atadero et al. (2004). The results from
the five date sets are listed in Table 9.4. The variations found were more consistent for
the last four samples with the values lying around 12 %. In this study, the COV for the
CFRP strength variability was taken to be 12 %.

353
CHAPTER 9

Table 9.4 Statistical parameters for FRP ultimate strength from Atadero et al.
Data set No. of samples Sample COV (%)
A 177 23.0
B1 49 12.1
B2 50 12.2
B3 20 13.6
C 260 12.7

9.4.3.5 Initial bottom strain


Most bridge girders in practice are under certain loading during application of
CFRP. The initial loading was taken to be equal to the dead load of the beam in this study
and assumed to be deterministic.

A summary of the variable statistical properties adopted in this study is listed in


Table 9.5.

Table 9.5 Summary of variable statistical properties


Distribution Standard
Variable Bias COV (%)
type distribution
bf Normal 1 - 10 mm
h Normal 1 - 10 mm
ds Normal 1 - 10 mm
fc Log-normal 1.10 20 -
fsy Log-normal 1.10 7.5 -
ffu Weibull 1.10 12 -
Lf Normal 1 - 10 mm
αno debond Normal 0.97 14.8 -
αintermediate span
Normal 1.01 12.7 -
debond

αend debond Normal 1.05 18.9 -

354
DEBOND STRENGTH MODELS AND RELIABLITY ANALYSIS FOR RETROFITTED BEAMS

9.4.4 Monte Carlo simulations and results


For each bridge configuration, 50000 data sets were generated. The simulation
results are shown in Table 9.6.

Table 9.6 Simulation results for the shear resistance R


Full composite Intermediate span debond End debond
Label Standard Standard Standard
Mean Bias Mean Bias Mean Bias
deviation deviation deviation
T1-C1a 533 1.051 87 441 1.098 69 N/C N/C N/C
T1-C1b 889 1.051 152 733 1.096 104 835 1.146 169
T1-C2a 699 1.042 114 465 1.092 66 592 1.140 118
T1-C2b 1146 1.025 193 775 1.091 110 826 1.142 166
T2-C1a 522 1.053 85 420 1.099 60 N/C N/C N/C
T2-C1b 871 1.054 142 700 1.101 102 838 1.141 172
T2-C2a 686 1.044 118 440 1.094 62 592 1.138 119
T2-C2b 1140 1.040 187 733 1.094 104 828 1.146 165
T3-C1a 666 1.053 108 558 1.100 80 N/C N/C N/C
T3-C1b 1109 1.052 180 931 1.101 134 1143 1.140 228
T3-C2a 832 1.041 135 576 1.102 82 796 1.133 159
T3-C2b 1387 1.042 226 959 1.100 137 1134 1.141 227
Note: 1) N/C: not critical.
2) The means and standard deviations are in kN.

9.4.5 Reliability
9.4.5.1 Reliability index
The performance of a structure can be represented by a limit-state function g,
which can be simplified as:
g = R − S = φR n (X 1 , X 2 , ..., X n ) − (γ D D + γ L L ) (9.9)
where R is the random resistance of the member and S is the random load effect acting on
the member. X1, X2…are the system random variables. D and L are dead and live loads,
respectively. The reliability index β can be defined as a measure of the probability that g
is less than zero, as follows
Pf = P(g ≤ 0) = Φ (− β ) (9.10)
where Φ is the cumulative distribution function of the standardised Normal distribution.
If g is normally distributed, β is taken as:

355
CHAPTER 9

µg
β= (9.11)
σg

where µg and σg are the mean and standard deviation of g.

For RC members, the target reliability index is about 3 (Ellingwood et al., 1980).
Allen (1992) suggested that the target reliability index should be increased by 0.25 for
components that fail suddenly with little warning but maintain their post failure capacity.
Since FRP retrofitted members maintain a capacity equal to that of the unstrengthened
member, β was taken as 3.25 in the present analysis.

9.4.5.2 Capacity reduction factor


To calculate the reduction factors for the strength of retrofitted members, the
method outline in Annex C of Eurocode 2 (2002) was used. The code provides guidance
for independently assessing safety factors for load and resistance effects. The main
advantage of this method is that the load variation and other uncertainties in its
distribution are not required. An outline of this method is presented below.

The design value can be defined as the point on the failure surface closest to the
average point in the space of normalised variables as diagrammatically indicated in
Figure 9.21.

Figure 9.21 Design point and reliability index according to the first order reliability
method for normally distributed uncorrelated variables (Eurocode 2, 2002)

356
DEBOND STRENGTH MODELS AND RELIABLITY ANALYSIS FOR RETROFITTED BEAMS

The design value of load effects, Sd, and resistance, Rd, should be defined such that the
probability of having a more unfavourable value is as follows
P(S > Sd) = Φ (+αSβ) (9.12)
P(R < Rd) = Φ (-αRβ) (9.13)
where αS and αR are the values of the FORM sensitivity factors. α is negative for
unfavourable load effects, and positive for resistances. Eurocode 2 allows independent
assessment of load and resistance if
0.16 < σS / σR < 7.6 (9.14)
where σS and σR are the standard deviations of the load effect and resistance,
respectively. Then, αR can be taken as 0.8 and Equation 9.14 becomes
P(R < Rd) = Φ (-0.8β) (9.15)

This condition was assumed to be satisfactory in the present study. With this
assumption and by reference to Figure 9.21, the capacity reduction factor was calculated
as:
R d µ R − α R βσ R ⎛ σ ⎞ µR
φ= = = ⎜⎜1 − α R β R ⎟⎟ (9.16)
Rn R ⎝ µR ⎠ Rn
µR n
µR

where σR, µR and Rn are the standard deviation, mean and nominal value of resistance,
respectively.

Table 9.7 lists the calculated reduction factors corresponding to the target
reliability index of 3.25. It can be seen from the table that the capacity reduction factor of
0.6 should be used for full composite action strength (no debond) and intermediate span
debond strength. A factor of 0.5 should be used for end debond strength. These values are
close to the lowest factor of 0.6 recommended by AS5100 (Table 9.8)

357
CHAPTER 9

Table 9.7 Capacity reduction factors corresponding to β = 3.25


Full Intermediate
Label End debond
composite span debond
T1-C1a 0.60 0.65 -
T1-C1b 0.58 0.69 0.54
T1-C2a 0.60 0.69 0.55
T1-C2b 0.58 0.69 0.55
T2-C1a 0.61 0.69 -
T2-C1b 0.61 0.68 0.53
T2-C2a 0.58 0.69 0.54
T2-C2b 0.60 0.69 0.55
T3-C1a 0.61 0.69 -
T3-C1b 0.61 0.69 0.55
T3-C2a 0.60 0.69 0.54
T3-C2b 0.60 0.69 0.55

Table 9.8 Capacity reduction factors recommended by AS5100


Type of action Capacity reduction factor φ
Bending without axial tension of
0.8
compression where ku ≤ 0.4
Concrete
Shear 0.7
members
Bending, shear and compression in plain
0.6
concrete
Members subject to bending 0.9
Composite compression members – Steel Concrete
Composite
Section capacity 0.9 0.6
members
Composite compression members –
0.9
Combined axial and flexural actions

9.5 Design recommendation


In this chapter, three theoretical models have been proposed to compute the
ultimate capacity of retrofitted beams with adequate shear strength. The adopted failure
mechanisms and the validity of the models have been proved using experimental and

358
DEBOND STRENGTH MODELS AND RELIABLITY ANALYSIS FOR RETROFITTED BEAMS

numerical results. The following steps are recommended to check the capacity of a
retrofitted beam:
1) Calculate the sectional capacity by following these steps
a. Calculate the sectional capacity using the beam theory.
b. Calculate the corresponding shear capacity and compare with applied
shear force using a reduction factor of 0.6, as follows
V* < 0.6 Vu,no debond (9.17)
2) Calculate the intermediate span debond capacity by following these
steps
a. Calculate the maximum force P that the composite can take using

P = βf βL f c b f L e (9.18)

b. Calculate the maximum FRP strain and the corresponding moment


capacity using beam theory.
c. Calculate the corresponding shear capacity and compare with applied
shear force using a reduction factor of 0.6, as follows
V* < 0.6 Vu,IS debond (9.19)
3) Calculate the end debond capacity by following these steps
a. Calculate the concrete shear strength, as follows

fcv,d = 0.2 fc (9.20)

b. Calculate the maximum force P in the composite


c. Calculate the maximum FRP strain and the corresponding moment
capacity using beam theory
d. Calculate the corresponding shear capacity and compare with applied
shear force using a reduction factor of 0.5, as follows
V* < 0.5 Vu,end debond (9.21)

359
CONCLUSIONS

CHAPTER 10 - CONCLUSIONS

10.1 Concluding remarks


Debonding of FRP composites from retrofitted concrete members is a complex
phenomenon. The aim of the research reported in this thesis was to provide greater
understanding of this phenomenon. Specifically, the aim was to establish the debond
behaviour of CFRP composites adhered to concrete members using a wet lay-up method.
The two member types considered were shear-lap specimens and retrofitted RC beams.

A literature review was conducted on the bond behaviour of FRP on concrete. The
review showed that a great deal of experimental work has been done in this area and
different approaches to calculate the bond strength have been proposed. The main failure
modes were identified in the review and the existing prediction models were assessed
based on a large database of test results. It was shown that two failure modes of shear-lap
specimens with properly bonded composites are interfacial debond and shear-tension
failure. It was found that most theoretical models to predict the shear-lap bond strength
are based on an assumed bond-slip relationship. From the assessment study, it is clear that
despite the models being based on different relationships, they generally produce
reasonably accurate predictions. However, the review also highlighted that the validity of
this assumption needs further investigation especially when transverse cracks also exist.

The literature review also revealed that failure of retrofitted RC beams is more
complex with a larger number of failure modes. The two main debonding modes
identified are intermediate span debond and end debond. It was established that the
models proposed to predict debond loads are based on very different approaches, and
because of the complexity of the problem, many assumptions and simplifications are
required in the analyses. As part of the present study, the models were assessed with a
large test database. It was shown that that while some models produce inaccurate and
scattered results, some are not based on the actual mechanisms and therefore lacks
grounding.

361
CHAPTER 10

An experimental program was carried out to examine the bond behaviour of


CFRP bonded on concrete using a wet lay-up method and to examine the effect of a
number of variables, such as the CFRP bond length and the CFRP thickness. In the
program, 26 specimens were loaded in single shear. It was found that the effective bond
length is approximately 100 mm for the chosen CFRP and concrete dimensions. It was
established that when the load reaches a certain value, the effective bond area shifts
further from the loaded edge and there is little increase in the bond capacity. From the
experimental observation, it may be concluded that the bond-slip relationships based on
the average values between the gauges are nonlinear and can be approximated using
Popovics’ equation. The parametric study also revealed that the CFRP bond length and
width have a significant effect on the bond behaviour.

Two experimental programs were conducted to identify the debonding failure


mechanisms of retrofitted RC beams and to study the effect of several parameters. The
first experimental program involved 18 beams under four-point bending. Two failure
modes were observed and their failure mechanisms were established. End debond is the
failure which starts from the CFRP end and propagates along the tensile rebar level. It
appears to occur when the shear bond stress reaches a critical level of stress. Intermediate
span debond is the failure which initiates from the tip of a diagonal crack and propagates
along the CFRP bond surface. It appears to occur when the CFRP strain level reaches a
critical level of strain. It was also found that the performance of CFRP retrofitting is
influenced significantly by the CFRP bond length in the shear span, the CFRP stiffness
and the tensile steel reinforcement stiffness. Steel clamps prove to be effective in
preventing end debond and improving the beam ductility.

The second experimental program involved eight tests of beams in three-point


bending. The aim was to investigate the effectiveness of U-straps in preventing
debonding failures. It was found that U-straps can be used to limit end debond and
intermediate span debond by placing them at the CFRP end and along the shear span,
respectively. A method was developed for prestressing these straps. However, the
experimental results indicated that the prestressed system improves the beam capacity
only slightly better than the non-prestressed.

362
CONCLUSIONS

Finite element modelling was carried out for the shear-lap specimens using
smeared cracks. A fine mesh was generated near the bond line to simulate the bond
behaviour. The finite element model made use of the plane stress elements and a rotating
crack model. A reasonable correlation was achieved between the outcome of the finite
element analyses and the experimental results. The finite element analyses also showed
that the CFRP strain varies significantly along its thickness near the loaded edge, the end
and the transverse cracks. The analyses confirmed that even though the bond-slip curves
can vary along the bond due to the presence of transverse cracks, they appear to have
similar shapes and the bond behaviour can be approximated using a bond-slip relationship
following Popovics’ equation.

Smeared crack modelling of the retrofitted beams without anchorage was


conducted. As the localised failure within the concrete cover was quite complex, a
number of simplifications were adapted in this study. The finite element model made use
of both rotating and fixed crack models. Plane stress elements were used together with
beam and interface elements. A good correlation was achieved between the results of the
finite element analyses and the experimental results in terms of the crack patterns, the
peak loads and the strain distributions in the components. The finite element model also
allowed a more detailed study of the effect of several important parameters. It was
revealed from the parametric study that the failure mode changes from flexure failure to
intermediate span debond and then end debond as the FRP thickness increases. The
maximum capacity is achieved when the failure mode is a combination of intermediate
span and end debond. A similar shifting of the failure modes also occurs when the tension
reinforcement amount or the CFRP bond length increases. A similar smeared crack model
was used to model the retrofitted beams anchored with U-straps. As the stress state near
the straps was very complex, the failure was not fully simulated. Nevertheless, the effect
of the straps on the crack patterns and the beam capacity was predicted reasonably well.

A different modelling approach was also used to simulate the behaviour of the
shear-lap specimens and the retrofitted beams under testing. In these analyses, a
combination of discrete and smeared cracks was utilised. The analyses made use of an
interface crack model developed recently, in which concrete degradation in both tension
and shear was modelled. The simulations yielded similar results to those of the smeared
crack models. The simulations also confirmed that the two main debonding modes of

363
CHAPTER 10

retrofitted beams are the result of shearing of concrete along the tensile rebar level in the
shear span and of concrete along the bond surface near the tip of a flexure-shear crack.

A design method was developed based on the findings from the experimental
programs and numerical simulations. The method can be used to predict the capacity of a
retrofitted RC beam failing by one of three modes: flexural failure, intermediate span
debond and end debond. The method is relatively simple. It is based on the well-known
beam theory, which has been verified to be able to predict the FRP strain level at ultimate
quite accurately. The end debond model uses the average shear stress concept, which can
be checked against the average concrete shear strength at the tensile rebar level. The
intermediate span debond model is based on the maximum force that the composite could
take, which can be calculated using an equation modified from that for a shear-lap test. A
simple equation relating the shear strength to the concrete compressive strength was also
proposed. From the verification results of the design method against both the data from
the experimental programs carried out in this research and the large database of tests
available in the literature, it was found that the method is reliable and can be used for
design.

10.2 Recommendations for future work


Further research is required in several areas of the present study. They are listed
below.

• Additional experimental and numerical work is required in the area of CFRP-


concrete bond behaviour. In particular, more experiments are needed to study the effect of
the CFRP width and the concrete strength on the bond-slip curves. Three-dimensional
modelling of shear-lap specimens under testing should also be carried out to investigate
the combined effects of varying composite and concrete geometries.
• Further improvement of the debond models could be achieved. This should
involve large-scale testing of RC beams and also three-dimensional modelling of the
beams. Further investigation of concrete shear strength along the tension reinforcement
layer of a RC beam might provide a better estimation of the strength based on a larger
number of parameters besides the concrete compressive strength.

364
CONCLUSIONS

• Debonding of the U-straps from the side of a retrofitted beam needs to be


investigated further.
• Other possible future research areas are the interaction between shear and flexural
strengthening, and the debonding behaviour of prestressed CFRP composites
longitudinally bonded to RC beams.

365
REFERENCES

REFERENCES

ACI Committee 363 (1992), State of the art report on high-strength concrete. Detroit,
American Concrete Institute.

ACI Committee 440F (1999), Guidelines for the selection, design and installation of fiber
reinforced polymer (FRP) systems for external strengthening of concrete structures.
Detroit, American Concrete Institute.

ACI Committee 440F (2002), Guide for the design and construction of externally bonded
FRP systems for strengthening concrete structures.

Ahmed, O. and Van Gemert, D. (1999), Effect of longitudinal carbon fiber reinforced
plastic laminates on shear capacity of reinforced concrete beams. Proceedings of the
Fourth International Symposium on Fiber Reinforced Polymer Reinforcement for
Reinforced Concrete Structures, Maryland, USA, pp. 933-943.

Allen, D. E. (1992), “Canadian highway bridge evaluation: Reliability index”, Canadian


Journal of Civil Engineering, Vol. 19, No. 6, pp. 593-602.

Al-Mahaidi, R. (2002), Strength assessment & FRP strengthening of concrete bridges.


Short course in Use of FRP Composites in Retrofitting of Concrete Structures, Monash
University, Melbourne.

Al-Sulaimani, G. J., Sharif, A., Basunbul, I. A., Baluch, M. H. and Ghaleb, B. N. (1994),
“Shear repair for reinforced concrete by fiberglass plate bonding”, ACI Structural
Journal, Vol. 91, No. 4, pp. 458-464.

An, W., Saadatmanesh, H. and Ehsani, M. R. (1991), “RC beams strengthened with FRP
plates. II. Analysis and parametric study”, Journal of Structural Engineering ASCE, Vol.
117, No. 11, pp. 3434-3455.

Aprile, A., Limkatanyu, S. and Spacone, E. (2001a), Analysis of R/C Beams


Strengthened with FRP Plates. ASCE Structures Conference, Washington, DC, USA.

Aprile, A., Spacone, E. and Limkatanyu, S. (2001b), “Role of bond in RC beams


strengthened with steel and FRP plates”, Journal of Structural Engineering, Vol. 127, No.
12, pp. 1445-1452.

Arduini, M., Di Tommaso, A. and Nanni, A. (1997), “Brittle failure in FRP plate and
sheet bonded beams”, ACI Structural Journal, Vol. 94, No. 4, pp. 363-370.

Arya, C. and Farmer, N. (2001), Design guidelines for flexural strengthening of concrete
members using FRP composites. FRPRCS-5 - Fibre-reinforced plastics for reinforced
concrete structures, UK, pp. 167-176.

367
REFERENCES

ASTM International (2000), D3039/D3039M-00: Standard test method for tensile


properites of polymer matrix composite materials. ASTM Standards. American Society
for Testing and Materials. West Conshohocken, PA, United States.

Atadero, R. A., Lee, L. S. and Karbhari, V. M. (2004), Materials variability and reliability
of FRP rehabilitation of concrete. 4th International Conference on Advanced Composite
Materials in Bridges and Structures, Calgary, Canada.

Ayers, S. R. and Van Erp, G. M. (2002), An Australian structural design standard for
fibre composites. ACUN - 4: Composite Systems -Macrocomposites, Microcomposites,
Nanocomposites, pp. 160-170.

Bakht, B., Al-Bazi, G., Banthia, N., Cheung, M., Erki, M.-A., Faoro, M., Machida, A.,
Mufti, A. A., Neale, K. W. and Tadros, G. (2000), “Canadian Bridge Design Code
provisions for fiber-reinforced structures”, Journal of Composites for Construction, Vol.
4, No. 1, pp. 3-15.

Bakis, C. E., Bank, L. C., Brown, V. L., Cosenza, E., Davalos, J. F., Lesko, J. J.,
Machida, A., Rizkalla, S. H. and Triantafillou, T. C. (2002), “Fiber-reinforced polymer
composites for construction - state-of-the-art review”, Journal of Composites for
Construction, Vol. 6, No. 2, pp. 73-87.

Batdorf, S. B. and Ghaffarian, R. (1984), “Size effect and strength variation of


unidirectional composites”, International Journal of Fracture, Vol. 26, No. 2, pp. 113-
123.

Bazant, Z. P. (2002), “Concrete fracture models: Testing and practice”, Engineering


Fracture Mechanics, Vol. 69, No. 2, pp. 165-205.

Bazant, Z. P. and Becq-Giraudon, E. (2002), “Statistical prediction of fracture parameters


of concrete and implications for choice of testing standard”, Cement and Concrete
Research, Vol. 32, No. 4, pp. 529-556.

Bazant, Z. P. and Oh, B. H. (1983), “Crack band theory for fracture of concrete”, Vol. 16,
No. 93, pp. 155-177.

Bizindavyi, L. and Neale, K. W. (1999), “Transfer lengths and bond strength for
composites bonded to concrete”, Journal of Composites for Construction, Vol. 3, No. 4,
pp. 153-160.

Blaschko, M. (1997), Strengthening with CFRP. Munchner Massivbau Seminar, TU


Muchen (In German).

Blaschko, M., Niedermeier, R. and Zilch, K. (1998), Bond Failure Modes of Flexural
Members Strengthened with FRP. Second International Conference on Composites in
Infrastructures, Tucson, Arizona, pp. 450-461.

Bonacci, J. F. and Maalej, M. (2001), “Behavioral Trends of RC Beams Strengthened


with Externally Bonded FRP”, Journal of Composites for Construction, Vol. 5, No. 2, pp.
102-113.

368
REFERENCES

Branson, D. E. (1977), Deformation of Concrete Structures. New York, McGraw-Hill,


Inc.

Brosens, K. and Van Gemert, D. (1997), Anchoring stresses between concrete and carbon
fibre reinforced laminates. Non-Metallic (FRP) Reinforcement for Concrete Structures,
Proc., 3rd Int. Symp, Sapporo.

Buyukozturk, O. and Hearing, B. (1998), “Failure Behaviour of Recracked Concrete


Beams Retrofitted with FRP”, Journal of Composites for Construction, Vol. 2, No. 3, pp.
138-144.

Camata, G., Spacone, E., Al-Mahaidi, R. and Saouma, V. (2004), “Analysis of test
specimens for cohesive near-bond failure of fiber-reinforced polymer-plated concrete”,
Journal of Composites for Construction, Vol. 8, No. 6, pp. 528-538.

Cervenka, J. (1994), Discrete crack modeling in concrete structures, PhD thesis,


Department of Civil, Environmental and Architectural Engineering, University of
Colorado.

Cha, J. Y., Balaguru, P. and Chung, L. (1999), Experimental and analytical investigation
of partially prestressed concrete beams strengthened with carbon reinforcement.
Proceedings of the Fourth International Symposium on Fiber Reinforced Polymer
Reinforcement for Reinforced Concrete Structures, Maryland, USA, pp. 625-633.

Chaallal, O., Nollet, M. J. and Perraton, D. (1998a), “Shear strengthening of RC beams


by externally bonded side CFRP strips”, Journal of Composites for Construction, Vol. 2,
No. 2, pp. 111-114.

Chaallal, O., Nollet, M. J. and Perraton, D. (1998b), “Strengthening of RC beams by


externally bonded fiber-reinforced-plastic plates: design guidelines for shear and flexure”,
Canadian Journal of Civil Engineering, Vol. Aug 1998, pp. 692-764.

Chajes, M. J., Finch, W. W., Jr., Januszka, T. F. and Thomson, T. A. J. (1996), “Bond and
force transfer of composite material plates bonded to concrete”, ACI Structural Journal,
Vol. 93, No. 2, pp. 208-217.

Chen, J. F. and Teng, J. G. (2001), “Anchorage Strength Models for FRP and Steel Plates
Bonded to Concrete”, Journal of Structural Engineering ASCE, Vol. 127, No. 7, pp. 784-
791.

Chen, J. F., Yang, Z. J., Pan, X. M. and Holt, G. D. (2001), Effect of test methods on
plate-to-concrete bond strength. FRPRCS-5 - Fibre-reinforced plastics for reinforced
concrete structures, UK, pp. 429-438.

Chen, W.-F. and Duan, L. (2000), Bridge Engineering Handbook. Boca Raton, CRC
Press.

Comite Euro-International du Beton (1985), CEB design manual on cracking and


deformations, Ecole Polytechnique Federale de Lausanne.

369
REFERENCES

Comite Euro-International du Beton (1991), CEB-FIB Model Code 1990. London, Great
Britain, Thomas Telford.

Concrete Society Technical Report 55 (2000), Design guidance for strengthening


concrete structures using fibre composite mateirials. Crowthorne, UK, The Concrete
Society.

CONTECVET (2002), A validated users manual for assessing the residual service life of
concrete structures - A manual for assessing corrosion-affected concrete structures, EC
Innovation Programme, IN30902I, GEOCISA.

Dai, J. G. and Ueda, T. (2003), Local bond stress slip relations for FRP sheets-concrete
interfaces. FRPRCS-6 - Fibre-reinforced polymer reinforcement for concrete structures,
Singapore, pp. 143-152.

David, E., Djelal, C., Ragneau, E. and Bodin, F. B. (1999), Use of FRP to strengthen and
repair RC beams: experimental study and numerical simulations. Proceedings of the
Eighth International Conference on Advanced Composites for Concrete Repair, London,
UK.

De Lorenzis, L., Miller, B. and Nanni, A. (2001), “Bond of FRP laminates to concrete”,
ACI Materials Journal, Vol. 98, No. 3, pp. 256-264.

de Witte, F. C. and Kikstra, W. P. (2003), DIANA User's Manual. Delft, The


Netherlands, TNO DIANA BV.

Dorr, K. (1980), Ein Beitrag zur Berechnung von Stahlbetonscheiben unter besonderer
Berücksichtigung des Verbundverhaltens, PhD thesis, University of Darmstadt (in
German).

Ellingwood, B., Galambos, T. V., MacGregor, J. G. and Cornel, C. A. (1980),


Development of a Probability Based Load Criterion for American National Standard A58.
Washington, D.C., National Bureau of Standards.

El-Mihilmy, M. T. and Tedesco, J. W. (2001), “Prediction of anchorage failure for


reinforced concrete beams strengthened with fiber-reinforced polymer plates”, ACI
Structural Journal, Vol. 98, No. 3, pp. 301-314.

Eurocode2 (2002), Basic of structural design, British Standard.

Fanning, P. J. and Kelly, O. (2000), “Smeared crack models of RC beams with externally
bonded CFRP plates”, Journal of Computational Mechanics, Vol. 26, No. 4, pp. 325-
332.

Fanning, P. J. and Kelly, O. (2001), “Ultimate Response of RC Beams Strengthened with


CFRP Plates”, Journal of Composites for Construction, Vol. 5, No. 2, pp. 122-127.

fib Bulletin 14 (2001), Design and use of externally bonded FRP reinforcement for RC
structures.

370
REFERENCES

Fukuyama, H. (1999), “Fibre-Reinforced Polymers in Japan”, Structural Engineering


International, Vol. No. 4, pp. 263-266.

Fu-quan, X., Jian-guang, G. and Yu, C. (2001), Bond strength between CFRP sheets and
concrete. FRP composites in civil engineering, Hong Kong, Elsevier, pp. 357-363.

Gao, B., Leung, W.-H., Cheung, C.-M., Kim, J.-K. and Leung, C. K. Y. (2001), Effects of
adhesive properties on strengthening of concrete beams with composite strips. FRP
composites in civil engineering, Hong Kong, Elsevier, pp. 423-432.

Garden, H. N., Hollaway, L. C. and Thorne, A. M. (1997), A preliminary evaluation of


carbon fibre reinforced polymer plates for strengthening reinforced concrete members.
Proc. Instn Civ. Engrs Structs & Bldgs, pp. 127-142.

Garden, H. N., Quantrill, R. J., Hollaway, L. C., Thorne, A. M. and Parke, G. A. R.


(1998), “An experimental study of the anchorage length of carbon fibre composite plates
used to strengthen reinforced concrete beams”, Construction & Building Materials, Vol.
12, pp. 203-219.

Gendron, G., Picard, A. and Guerin, M.-C. (1999), “A Theoretical Study on Shear
Strenthening of Reinforced Concrete Beams Using Composite Plates”, Composite
Structures, Vol. 45, pp. 303-309.

Harlow, D. G. and Phoenix, S. L. (1981), “Probability distributions for the strengthe of


composite materials II: A convergent sequence of tight bounds"”, International JOurnal
of Fracture, Vol. 17, No. 6, pp. 601-630.

Hearing, B. (2000), Delamination in reinforced concrete retrofitted with fiber reinforced


plastics, PhD thesis, Deparment of Civil and Environmental Engineering, Massachusetts
Institute of Technology.

Hillerborg, A., Modeer, M. and Petersson, P.-E. (1976), “Analysis of crack formation and
crack growth in concrete by means of fracture mechanics and finite element.s”, Cement &
Concrete Research, Vol. 6, pp. 773-782.

Holicky, M. and Markova, J. (2000), “Verification of load factors for concrete


components by reliability and optimization analysis: Background documents for
implementing Eurocodes”, Prog. Struct. Engng Mater, Vol. 2, pp. 502-507.

Hollaway, L. C. and Leeming, M. B. (1999), Strengthening of Reinforced Concrete


Structures - Using Externally-Bonded FRP Composites in Structural and Civil
Engineering. Cambridge, Woodhead Publishing Limited.

Holzenkampfer, O. (1994), Ingenieurmodelle des verbundes geklebter bewehrung fur


betonbauteile, Dessertation thesis, TU Braunschweig (in German).

Hordijk, D. A. (1991), Local approach to Fatigue of Concrete, PhD thesis, Delft


University of Technology.

371
REFERENCES

Hormann, M., Menrath, H. and Ramm, E. (2002), “Numerical investigation of fiber


reinforced polymers poststrengthened concrete slabs”, Journal of Engineering
Mechanics, Vol. 128, No. 5, pp. 552-561.

Hull, D. and Clyne, T. W. (1996), An introduction to composite materials, Cambridge


University Press.

Hussain, M., Sharif, A., Basunbul, I. A., Baluch, M. H. and Al-Sulaimani, G. J. (1995),
“Flexural behaviour of precracked reinforced concrete beams strengthened externally by
steel plates”, ACI Structural Journal, Vol. 92, No. 1, pp. 14-22.

ISIS Canada (2001), Design Manual No. 4 - Strengthening reinforced concrete structures
with externally-bonded FRP.

Izumo, K., Asamizu, T., Saeki, N. and Shimura, K. (1997), “Shear strengthening of PRC
members by fiber sheets”, Transactions of the Japan Concrete Institute, Vol. 19, pp. 105-
112.

Izumo, K., Asamizu, T., Saeki, N. and Shimura, K. (1998), “Bond Behaviour of Aramid
and Carbon Fiber Sheeting”, Transactions of the Japan Concrete Institute, Vol. 20, pp.
269-279.

Jansze, W. (1997), Strengthening of RC members in bending by externally bonded steel


plates, PhD thesis, Delft University of Technology.

Jones, R., Swamy, R. N. and Charif, A. (1988), “Plate Separation and Anchorage of
Reinforced Concrete Beams Strengthened By Epoxy-Bonded Steel Plates”, Structural
Engineer, Vol. 66, No. 12, pp. 187-188.

Kanakubo, T., Furuta, T. and Fukuyama, H. (2003), Bond strength between fiber-
reinforced polymer laminates and concrete. FRPRCS-6 - Fibre-reinforced polymer
reinforcement for concrete structures, Singapore, pp. 133-142.

Karbhari, V. M. (2001), “Materials considerations in FRP rehabilitation of concrete


structures”, Journal of Materials in Civil Engineering, Vol. 13, No. 2, pp. 90-97.

Karbhari, V. M. and Engineer, M. (1996), “Investigation of bond between concrete and


composites: use of a peel test”, Journal of Reinforced Plastics & Composites, Vol. 15,
No. 2, pp. 208-227.

Karbhari, V. M., Engineer, M. and Eckel II, D. A. (1997), “On the durability of
composite rehabilitation schemes for concrete: use of a peel test”, Journal of Materials
Science, Vol. 32, pp. 147-156.

Karbhari, V. M. and Seible, F. (1999), “Fiber-reinforced polymer composites for civil


infrastructure in the USA”, Structural Engineering International, Vol. 4/99, pp. 274-277.

Kim, W. and White, R. N. (1991), “Initiation of shear cracking in reinforced concrete


beams with no web reinforcement”, ACI Structural Journal, Vol. 88, No. 3, pp. 301-308.

372
REFERENCES

Kishi, N., Mikami, H., Matsuoka, K. G. and KUrihashi, Y. (2001), Failure behavior of
flexural strengthened RC beams with AFRP sheet. FRPRCS-5 - Fibre-reinforced plastics
for reinforced concrete structures, UK, pp. 85-95.

Kotsovos, M. D. and Pavlovic, M. N. (1995), Structural concrete - Finite element analysis


for limit-state design. New York, Thomas Telfold.

Kupfer, H. B. and Grestle, K. H. (1973), “Behavior of concrete under biaxial stresses”,


ASCE Journal of the Engineering Mechanics Division, Vol. 99, No. EM4, pp. 853-866.

Lee, K. (2003), Shear strength of concrete T-beams repaired using carbon fibre reinforced
polymer, PhD thesis, Department of Civil Engineering, Monash University.

Lee, K. and Al-Mahaidi, R. (2003), Strength and failure mechanisms of RC T-beams


strengthened with CFRP plates. FRPRCS-6 - Fibre-reinforced polymer reinforcement for
concrete structures, Singapore, pp. 247-256.

Lee, K., Al-Mahaidi, R., Taplin, G. and Millar, D. (2002), Behaviour of Reinforced
Concrete T-beams Strengthened with Prefabricated L-shaped CFRP Plates. IABSE
symposium, Melbourne.

Lees, J. M., Winistörfer, A. and Meier, U. (2002), “External prestressed carbon fiber-
reinforced polymer straps for shear enhancement of concrete”, Journal of Composites for
Construction, Vol. 6, No. 4, pp. 249-256.

Maeda, T., Asano, Y., Sato, Y., Ueda, T. and Kakuta, Y. (1997), A Study on Bond
Mechanism of Carbon Fiber Sheet. Proc., 3rd Symp. - Non-Metallic (FRP)
Reinforcement for Concrete Struct., Japan, pp. 279-286.

Malek, A. M., Saadatmanesh, H. and Ehsani, M. R. (1998), “Prediction of failure load of


R/C beams strengthened with FRP plate due to stress concentration at the plate end”, ACI
Structural Journal, Vol. 95, No. 2, pp. 142-152.

Maruyama, K. and Ueda, T. (2001), JSCE design recommendations for upgrading of RC


members by FRP sheet. FRPRCS-5 - Fibre-reinforced plastics for reinforced concrete
structures, UK, pp. 441-446.

Matthys, S. (2000), Structural behaviour and design of concrete members strengthened


with externally bonded FRP reinforcement, thesis, Ghent University.

Menetrey, P. and Willam, K. J. (1995), “Triaxial failure criterion for concrete and its
generalization”, ACI Structural Journal, Vol. 92, No. 3, pp. 311-318.

MERLIN User's Manual (2001), Department of Civil Engineering, University of


Colorado, Boulder.

Mirza, S. A. and MacGregor, J. G. (1979), “Varitions in dimensions of reinforced


concrete members”, Vol. 105, No. 4, pp. 751-766.

373
REFERENCES

Mohamed Ali, M. S. (2000), Peeling of plates adhesively bonded to reinforced concrete


beams, PhD thesis, Department of Civil Engineering, Adelaide University.

Mukhopadhyaya, P. and Swamy, R. N. (2001), “Interface Shear Stress: A New Design


Criterion for Plate Debonding”, Journal of Composites for Construction, Vol. 5, No. 1,
pp. 35-43.

Naaman, A. E., Park, S. Y., Lopez, M. M. and Till, R. D. (2001), Parameters influencing
the flexural response of RC beams strengthened using CFRP sheets. FRPRCS-5 - Fibre-
reinforced plastics for reinforced concrete structures, UK, pp. 117-125.

Nakaba, K., Kanakubo, T., Furuta, T. and Yoshizawa, H. (2001), “Bond Behavior
between Fiber-Reinforced Polymer Laminates and Concrete”, ACI Structural Journal,
Vol. 98, No. 3, pp. 359-367.

Neubauer, U. and Rostasy, F. S. (1997), Design aspects of concrete structures


strengthened with externally bonded CFRP-plates. Proceeding of the 7th International
Conference on Structural Faults and Repairs, pp. 109-118.

Neubauer, U. and Rostasy, F. S. (1999), Bond failure of CFRP plates at inclined cracks -
experiments and fracture mechanics model. Fourth International Symposium - Fiber
Reinforced Polymer Reinforcement for Reinforced Concrete, American Concrete
Institute, pp. 369-381.

Nguyen, D. M., Chan, T. K. and Cheong, H. K. (2001), “Brittle Failure and Bond
Development Length of CFRP-Concrete Beams”, Journal of Composites for
Construction, Vol. 5, No. 1, pp. 12-17.

Niedermeier, R. (1996), Stellungnahme zur Richtlinie fu¨r das Verkleben von


Betonbauteilen durch Ankleben von Stahllaschen-Entwurf Ma ¨rz 1996, Schreiben 1390
vom 30.10.1996 des Lehrstuhls fu¨r Massivbau, Technische Universita¨t Mu¨nchen,
Munich, Germany (in German).

Niedermeier, R. (2000), Zugkraftdeckung bei klebearmierten bauteilen (Envelope line of


tensile forces while using externally bonded reinforcement), thesis, TU Munchen (in
German).

Niu, H. D. and Wu, Z. (2001a), Peeling-off criterion for FRP-strengthened R/C flexural
members. FRP composites in civil engineering, Hong Kong, Elsevier, pp. 571-578.

Niu, H. D. and Wu, Z. S. (2001b), “Interfacial debonding mechanism influenced by


flexural cracks in FRP-strengthened beams”, Journal of Structural Engineering, JSCE,
Vol. 47, No. A, pp. 1277-1288.

Oehlers, D. J. (1992), “Reinforced concrete beams with plates glued to their soffits”,
Journal of Structural Engineering ASCE, Vol. 118, No. 8, pp. 2023-2038.

Oehlers, D. J. and Moran, J. P. (1990), “Premature failure of externally plated reinforced


concrete beams”, Journal of Structural Engineering ASCE, Vol. 116, No. 4, pp. 978-995.

374
REFERENCES

Okeil, A. M., El-Tawil, S. and Shahawy, M. (2001), “Short-term tensile strength of


carbon fiber-reinforced polymer laminates for flexural strengthening of concrete girders”,
ACI Structural Journal, Vol. 98, No. 4, pp. 470-478.

Plevris, N., Triantafillou, T. C. and Veneziano, D. (1995), “Reliability of RC members


strengthened with CFRP laminates”, Journal of Structural Engineering ASCE, Vol. 121,
No. 7, pp. 1037-1044.

Popovics, S. (1973), “A numerical approach to complete stress-strain curve of concrete”,


Cement and Concrete Research, Vol. 3, pp. 583-599.

Quantrill, R. J., Hollaway, L. C. and Thorne, A. M. (1996), “Prediction of the maximum


plate end stresses of FRP strengthened beams: Part II”, Magazine of Concrete Research,
Vol. 48, No. 177, pp. 343-351.

Rahimi, H. and Hutchinson, A. (2001), “Concrete Beams Strengthened with Externally


Bonded FRP Plates”, Journal of Composites for Construction, Vol. 5, No. 1, pp. 44-56.

Raoof, M. and Hassanen, M. A. H. (2000), Peeling failure of reinforced concrete beams


with fiber-reinforced plastic or steel plates glued to their soffits. Proceedings of the
Institition of Civil Engineers: Structures and Buildings, pp. 291-305.

Raoof, M. and Zhang, S. (1997), An insight into the structural behaviour of reinforced
concrete beams with externally bonded plates. Proceedings of the Institition of Civil
Engineers: Structures and Buildings, pp. 477-492.

RILEM Technical Committee 90-FMA (1989), Fracture mechanics of concrete structures


- From theory to applications. London, Chapman and Hall Ltd.

Rizkalla, S. H. (2002), Use of FRP composites in retrofitting of concrete structures. Short


course in Use of FRP Composites in Retrofitting of Concrete Structures, Monash
University, Melbourne.

Roberts, T. M. (1989), “Approximate analysis of shear and normal stress concentrations


in the adhesive layer of plated RC beams”, Structural Engineer, Vol. 67, No. 12, pp. 229-
233.

Ross, C. A., Jerome, D. M., Tedesco, J. W. and Hughes, M. L. (1999), “Strengthening of


reinforced concrete beams with externally bonded composite laminates”, ACI Structural
Journal, Vol. 96, No. 2, pp. 212-220.

Rots, J. G. (1989a), Bond of reinforcement. Fracture mechanics of concrete structures -


From theory to applications. Elfgren, L. London, Chapman and Hall Ltd: 245-262.

Rots, J. G. (1989b), “Crack models for concrete: discrete or smeared? fixed, multi-
directional or rotating”, Heron, Vol. 34, No. 1.

Rots, J. G., Nauta, P., Kusters, G. M. A. and Blaauwendraad, J. (1985), “Smeared crack
approach and fracture localization in concrete”, Heron, Vol. 30, No. 1.

375
REFERENCES

Saadatmanesh, H. and Ehsani, M. R. (1990), “Fiber composite plates can strengthen


beams”, Concrete International: Design & Construction, Vol. 12, No. 3, pp. 65-71.

Saadatmanesh, H. and Malek, A. M. (1998), “Design guidelines for flexural strengthening


of RC beams with FRP plates”, Journal of Composites for Construction, Vol. 2, No. 4,
pp. 158-164.

Saafi, M. and Buyle-Bodin, F. (1997), Shear capacity and ductility of reinforced concrete
beams strengthened by plate bonding. Non-Metallic (FRP) Reinforcement for Concrete
Struct. Proc., 3rd Symp., Japan, pp. 351.

Saouma, V. (2002), Lecture notes in Fracture Mechanics. Boulder, Department of Civil


Environmental and Architectural Engineering, University of Colorado.

Sebastian, W. M. (2001), “Significance of midspan debonding failure in FRP-plated


concrete beams”, Journal of Structural Engineering ASCE, Vol. 127, No. 7, pp. 792-798.

Sharif, A., Al-Sulaimani, G. J., Basunbul, I. A., Baluch, M. H. and Ghaleb, B. N. (1994),
“Strengthening of initially loaded reinforced concrete beams using FRP plates”, ACI
Structural Journal, Vol. 91, No. 2, pp. 160-168.

Shehata, I. A. E. M., Cerqueira, E. C., Pinto, C. T. M. and Shehata, L. C. D. (2001),


Strengthening of R. C. beams in flexure and shear using CFRP laminate. FRPRCS-5 -
Fibre-reinforced plastics for reinforced concrete structures, UK, pp. 97-106.

Shokrieh, M. M. and Malevajerdy, S. A. M. (2001), Strengthening of reinforced concrete


beams using composite laminates. FRP composites in civil engineering, Hong Kong,
Elsevier, pp. 507-515.

Smith, G. and Teng, J. G. (2003), “Shear-bending interaction in debonding failures of


FRP-plated RC beams”, Advances in Structural Engineering, Vol. 6, No. 3, pp. 183-199.

Smith, S. T. and Teng, J. G. (2001), “Interfacial stresses in plated beams”, Engineering


Structures, Vol. 23, pp. 857-871.

Smith, S. T. and Teng, J. G. (2002), “FRP-strengthening RC beams. II: assessment of


debonding strength models”, Engineering Structures, Vol. 24, pp. 397-417.

Standards Australia (1991), AS1391-1991: Methods for tensile testing of metals. Sydney,
Australia.

Standards Australia (1999), AS1012.9-1999: Methods of testing concrete - Method 9:


Determination of the compressive strength of concrete specimens. Sydney, Australia.

Standards Australia (2001), AS3600-2001: Concrete structures. Sydney, Australia.

Swamy, R. N., Jones, R. and Bloxham, J. W. (1987), “Structural Behaviour of Reinforced


Concrete Beams Strengthened By Epoxy-Bonded Steel Plates”, Structural Engineer, Vol.
65A, No. 2, pp. 59-68.

376
REFERENCES

Täljsten, B. (1997), “Defining anchor lengths of steel and CFRP plates bonded to
concrete”, International Journal of Adhesion & Adhesives, Vol. 17, No. 4, pp. 319-327.

Tasuji, M. E. and Slate, F. O. (1978), “Stress-strain response and fracture of concrete in


biaxial loading”, ACI Journal, Vol. 75, pp. 306-312.

Teng, J. G., Smith, G., Yao, J. and Chen, J. F. (2001), Strength model for intermediate
flexural crack induced debonding in RC beams and slabs. FRP composites in civil
engineering, Hong Kong, Elsevier, pp. 579-587.

Teng, J. G., Zhang, J. W. and Smith, S. T. (2002), “Interfacial stresses in reinforced


concrete beams bonded with a soffit plate: a finite element study”, Construction &
Building Materials, Vol. 16, pp. 1-14.

Thorenfeldt, E., Tomaszewicz, A. and Jensen, J. J. (1987), Mechanical properties of high-


strength concrete and applications in design. Proc. Symp. Utilization of High-Strength
Concrete (Stavanger, Norway), Tapir.

Toutanji, H. and Ortiz, G. (2001), “The effect of surface preparation on the bond interface
between FRP sheets and concrete members”, Composite Structures, Vol. 53, pp. 457-462.

Triantafillou, T. C. (1997), Shear Strengthening of Concrete Members using Composites.


Proceedings of the Third International Symposium - Non-Metallic (FRP) Reinforcement
for Concrete Structures, pp. 523-530.

Triantafillou, T. C. (1998), “Shear strengthening of reinforced concrete beams using


epoxy-bonded FRP composites”, ACI Structural Journal, Vol. 95, No. 2, pp. 107-115.

Triantafillou, T. C. and Antonopoulos, C. P. (2000), “Design of concrete flexural


members strengthened in shear with FRP”, Journal of Composites for Construction, Vol.
4, No. 4, pp. 198-205.

Triantafillou, T. C. and Plevris, N. (1992), “Strengthening of RC beams with epoxy-


bonded fibre-composite materials”, Materials & Structures, Vol. 25, pp. 201-211.

Trunk, B. and Wittmann, F. H. (1998), Experimental investigation into the size


dependence of fracture mechanics parameters. Third international conference of fracture
mechanics of concrete structures, D-Freiburg: Aedificatio Publ., pp. 1937-1948.

Tumialan, G., Serra, P., Nanni, A. and Belarbi, A. (1999), Concrete cover delamination in
reinforced concrete beams strengthened with carbon fiber reinforced polymer sheets.
Proceedings of the Fourth International Symposium on Fiber Reinforced Polymer
Reinforcement for Reinforced Concrete Structures, Maryland, USA, pp. 725-735.

Ulaga, T. and Vogel, T. (2003), Bilinear stress-slip bond model: theoretical background
and significance. FRPRCS-6 - Fibre-reinforced polymer reinforcement for concrete
structures, Singapore, pp. 153-162.

Val, D. V. (2003), “Reliability of fiber-reinforced polymer-confined reinforced concrete


columns”, Journal of Structural Engineering, Vol. 129, No. 8, pp. 1122-1130.

377
REFERENCES

Van Gemert, D. (1980), “Force transfer in epoxy bonded steel/concrete joints”,


International Journal of Adhesion and Adhesives, Vol. No. 1, pp. 67-72.

van Mier, J. G. M. (1997), Fracture processes of concrete - Assessment of material


parameters for fracture models. Boca Raton, CRC Press.

Vecchio, F. J. and Bucci, F. (1999), “Analysis of repaired reinforced concrete structures”,


Journal of Structural Engineering ASCE, Vol. 125, No. 6, pp. 644-652.

Vecchio, F. J. and Collins, M. P. (1986), “The modified compression-field theory for


reinforced concrete elements subjected to shear”, ACI Journal, Vol. 83, No. 2, pp. 219-
231.

Warner, R. F. (1998), Concrete structures. Sydney, Addison Wesley Longman Australia


Pty Limited.

Watstein, D. and Parsons, D. E. (1943), “Width and spacing of tensile cracks in


reinforced concrete cylinders”, J. Res. Nat. Bur. Stand., Vol. 31, No. PR545, pp. 1-24.

Wittmann, F. H. (2002), “Crack formation and fracture energy of normal and high
strength concrete”, Sadhana, Vol. 27, No. 4, pp. 413-423.

Wittmann, F. H., Mihashi, H. and Nomura, N. (1990), “Size effect on fracture energy of
concrete”, Engineering Fracture Mechanics, Vol. 35, No. 1-3, pp. 107-115.

Wong, R. S. Y. and Vecchio, F. J. (2003), “Towards modeling of reinforced concrete


members with externally bonded fiber-reinforced polymer composites”, ACI Structural
Journal, Vol. 100, No. 1, pp. 47-55.

Wong, W. F., Chiew, S. P. and Sun, Q. (2001), Flexural strengthening of RC beams


strengthened with FRP plate. FRP composites in civil engineering, Hong Kong, Elsevier,
pp. 633-640.

Wu, Z., Yuan, H. and Niu, H. D. (2002), “Stress transfer and fracture propagation on
different kinds of adhesive joints”, Journal of Engineering Mechanics, Vol. 128, No. 5,
pp. 562-573.

Yang, Z. J., Chen, J. F. and Proverb, D. (2002), “Finite element modelling of concrete
cover separation failure in FRP plated RC beams”, Construction and Building Materials,
Vol. 17, pp. 3-13.

Yao, J., Teng, J. G. and Chen, J. F. (2005), “Experimental study on FRP-to-concrete


bonded joints”, Composites Part B: Engineering, Vol. 36, No. 2, pp. 99-113.

Yuan, H., Wu, Z. S. and Yoshizawa, H. (2001), “Theoretical solutions on interfacial


stress transfer of externally bonded steel/composite laminates”, J. Struct. Mech. and
Earthquake Engrg., Vol. No. No.675/I-55, pp. 27-39.

378
REFERENCES

Zarnic, R. and Bosiljkov, V. (2001), Behaviour of Beams Strengthened with FRP and
Steel Plates. The 2001 Structural Congress and Exposition, Washington, D.C.

Zarnic, R., Gostic, S., Bosiljkov, V. and Bosiljkov, B. (1999), Improvement of Bending
Load-Bearing Capacity by Externally Bonded Plates. Proc. Creating with Concrete,
London, Thomas Telford, pp. 433-442.

Zhang, J.-P. (1997), “Diagonal cracking and shear strength of reinforced concrete
beams”, Magazine of Concrete Research, Vol. 49, No. 178, pp. 55-65.

Zhang, S. and Raoof, M. (1995), Prediction of peeling failure of reinforced concrete


beams with externally bonded steel plates. Proceedings of the Institition of Civil
Engineers: Structures and Buildings, pp. 257-268.

Ziraba, Y. N., Baluch, M. H., Basunbul, I. A., Sharif, A. M., Azad, A. K. and Al-
Sulaimani, G. J. (1994), “Guidelines toward the design of reinforced concrete beams with
external plates”, ACI Structural Journal, Vol. 91, No. 6, pp. 639-646.

379

You might also like