You are on page 1of 349

Tension stiffening in reinforced concrete - instantaneous and

time-dependent behaviour

Author:
Wu, Mark Han Qing
Publication Date:
2010
DOI:
https://doi.org/10.26190/unsworks/23066
License:
https://creativecommons.org/licenses/by-nc-nd/3.0/au/
Link to license to see what you are allowed to do with this resource.

Downloaded from http://hdl.handle.net/1959.4/45041 in https://


unsworks.unsw.edu.au on 2024-01-29
TENSION STIFFENING IN REINFORCED
CONCRETE – INSTANTANEOUS AND
TIME-DEPENDENT BEHAVIOUR

Mark Han Qing Wu

A thesis submitted in partial fulfilment of

the requirements for the degree of Doctor of Philosophy

February 2010

UNSW
THE UNIVERSITY OF NEW SOUTH WALES, SYDNEY, AUSTRALIA

School of Civil and Environmental Engineering


DECLARATION OF ORIGNIALITY

I hereby declare that this submission is my own work and that, to the best of my knowledge

and belief, it contains no materials previously published or written by another person, nor

material which to a substantial extent has been accepted for the award of any other degree or

diploma at UNSW or any other institute of higher learning, except where due

acknowledgment is made in the thesis. Any contribution made to the research by others, with

whom I have worked at UNSW or elsewhere, is explicitly acknowledged in the thesis.

I also declare that the intellectual content of this thesis is the product of my own work, except

to the extent that assistance from others in the project’s design and conception and linguistic

expression is acknowledged.

______
Mark Han Qing Wu
Dedicated to my Lord and Saviour:

Jesus Christ

Also to my beloved parents:

Qiang Wu and Lin Huang


ACKNOWLEDGEMENT

The research reported in this thesis was undertaken in the School of Civil and Environmental

Engineering at University of New South Wales, Sydney. Sincere thanks is given to Professor

Ian Gilbert, who has supervised me through the entire research period with his gentle heart

and great suggestions. He has also cultivated my ability to think carefully about the academic

program. I also express my deep gratitude to Professor Stephen Foster, for his constructive

help in the computational modeling aspects of this study, without which this thesis would not

have been completed satisfactorily.

I would also like to thank the staff at the Heavy Structure Laboratory at the Randwick sub-

campus of UNSW, including the staff Ron Moncay, John Gilbert and Frank Scharfe, all of

whom have offered great and faithful assistance, so that the experimental program could be

carried out successfully. I am also grateful to the administrative staff at the School of Civil

and Environmental Engineering, for their consistent help over this 3.5 years academic period.

Finally I would like to thank my parents, who have always showed their steadfast love and

support throughout this PhD program.


LIST OF PUBLICATIONS

1. Wu, H. Q., and Gilbert, R. I. (2009a) "Modeling short-term tension stiffening in reinforced

concrete prisms using a continuum-based finite element model." Engineering structures,

31(10), 2380-2391.

2. Wu, H. Q., and Gilbert, R. I. (2009b) "Effect of shrinkage on time-dependent deflection of

reinforced concrete slabs." Proceedings of the 24th Biennial Conference of the Concrete

Institute of Australia, Sydney, Paper No.6b-1.

3. Wu, H. Q., (2008) “Time-dependent effect of tension stiffening on cracks of RC

structures.” Quality of Civil Engineering and Construction, 10(A), 620-625.

4. Gilbert, R. I., and Wu, H. Q. (2009). "Time-dependent stiffness of cracked reinforced

concrete elements under sustained actions." Australian Journal of Structural Engineering,

9(2), 151-158.
ABSTRACT

The research presented in this thesis is concerned with the time-dependent behaviour of

reinforced concrete and the mechanisms of tension stiffening. The research program

comprises both experimental investigation and numerical modeling. The aims are to provide a

comprehensive understanding about tension stiffening under short-term and long-term loading.

Of particular interests are the effects of concrete creep and shrinkage on the change of tension

stiffening with time.

The experimental program involved testing axially loaded reinforced concrete prisms and

simply-supported reinforced concrete beams and slabs under short-term and long-term

loadings. The experimental results have demonstrated that tension stiffening is greatly

influenced by the drying shrinkage of concrete, the extent of cracking and the deterioration

with time of the bond between the reinforcement and the concrete.

A two-dimensional continuum-based finite element model has been developed. The model

incorporates instantaneous and time-dependent constitutive laws for concrete, steel and

importantly the bond interface between them. A shrinkage-related bond model has been

proposed to accurately model the loss of tension stiffening under long-term loading.

The finite element model has been used to simulate the response of the test specimens and the

numerical results are in close agreement with the experimental results. A parametric study has

also been undertaken using the finite element model; the effects of different parameters

(concrete cover, reinforcing ratio, etc.) on time-dependent tension stiffening have been

examined. The significance of each parameter has been discussed.


I

TABLE OF CONTENTS

CHAPTER 1 INTRODUCTION 1
1.1 GENERAL ......................................................................................................................... 1
1.2 OBJECTIVES OF THE CURRENT STUDY.................................................................... 5
1.3 THESIS STRUCTURE...................................................................................................... 6

CHAPTER 2 CURRENT STATE OF KNOWLEDGE 8


2.1 OVERVIEW....................................................................................................................... 8
2.2 TENSION STIFFENING UNDER INSTANTANEOUS LOAD ...................................... 9

2.2.1 Micro and Macro behaviour in uniaxial tension members ................................................. 9

2.2.2 Micro and Macro behaviour in flexural members. ........................................................... 14

2.3 PREVIOUS EXPERIMENTAL INVESTIGATION ON TENSION STIFFENING ....... 17

2.3.1 Short-term experiments .................................................................................................... 17

2.3.2 Long-term experiments .................................................................................................... 19

2.3.3 Comment on the previous experimental works on tension stiffening .............................. 21

2.4 CURRENT METHODS FOR PREDICTING TENSION STIFFENING........................ 22

2.4.1 AS3600-2009 and Eurocode 2 Methods........................................................................... 22

2.4.2 Modelling Tension Stiffening in the Finite Element Context........................................... 26

2.5 BOND-SLIP ON THE CONCRETE-STEEL BAR INTERFACE................................... 32

2.5.1 Bond Zone Scales............................................................................................................. 32

2.5.2 Bond Mechanisms ............................................................................................................ 34

2.5.3 Characteristics of Bond Stress-Slip Relationship ............................................................. 37

2.5.4 Bond Models .................................................................................................................... 41

2.5.5 Time-dependent bond response ........................................................................................ 45

2.6 TIME EFFECTS OF PLAIN CONCRETE ..................................................................... 46

2.6.1 Concrete as a time-dependent material............................................................................. 47

2.6.2 Mathematical Modeling of Creep..................................................................................... 51


II

CHAPTER 3 TIME-DEPENDENT ANALYSIS OF TENSION STIFFENING 61


3.1 OVERVIEW..................................................................................................................... 61
3.2 TIME-DEPENDENT ANALYSIS OF A UNIAXIAL TENSION MEMBER ................. 61
3.3 TIME-DEPENDENT ANALYSIS OF A FLEXURAL MEMBER.................................. 64

3.3.1 Time dependent response of an uncracked singly-reinforced section .............................. 65

3.3.2 Time –dependent response of a cracked singly-reinforced section .................................. 67

3.3.3 Time-dependent tension stiffening in the cracked reinforced concrete beam .................. 69

3.4 SUMMARY ..................................................................................................................... 72

CHAPTER 4 EXPERIMENTAL PROGRAM AND RESULTS OF UNIAXIAL


TENSION TESTS 74
4.1 OVERVIEW..................................................................................................................... 74
4.2 SPECIMEN CONFIGURATION..................................................................................... 75
4.3 TEST INSTRUMENTATION.......................................................................................... 77
4.4 CURING AND TESING PROCEDURE ......................................................................... 79
4.5 MATERIAL PROPERTIES ............................................................................................. 83

4.5.1 Steel bar properties........................................................................................................... 83

4.5.2 Instantaneous properties of concrete ................................................................................ 83

4.5.3 Creep and shrinkage ......................................................................................................... 84

4.6 TESTING RESULTS OF THE UNIAXIAL TENSION PRISMS ................................... 87

4.6.1 Short-term test results....................................................................................................... 87

4.6.2 Long-term test results..................................................................................................... 108

4.7 SUMMARY ................................................................................................................... 120

CHAPTER 5 EXPERIMENTAL PROGRAM AND RESULTS OF FLEXURAL


TESTS 122
5.1 OVERVIEW................................................................................................................... 122
5.2 SPECIMEN CONFIGURATION................................................................................... 122
5.3 TEST INSTRUMENTATION........................................................................................ 125
5.4 CURING AND TESING PROCEDURE ....................................................................... 129
5.5 MATERIAL PROPERTIES ........................................................................................... 132

5.5.1 Concrete instantaneous properties.................................................................................. 132


III

5.5.2 Creep and Shrinkage ...................................................................................................... 133

5.6 TEST RESULTS FOR THE FLXURAL SPECIMENS................................................. 134

5.6.1 Short-term results ........................................................................................................... 134

5.6.2 Long-term results (SLTN4-12A and SLTN4-12B) ......................................................... 160

5.7 SUMMARY ................................................................................................................... 165

CHAPTER 6 FINITE ELEMENT FORMULATION AND MATERIAL


CONSTITUTIVE LAW 167
6.1 OVERVIEW................................................................................................................... 167
6.2 FOUR-NODE ISOPARAMETRIC CONCRETE ELEMENT ...................................... 168

6.2.1 Finite element formulation ............................................................................................. 168

6.2.2 Concrete element stress-strain relationship .................................................................... 172

6.3.1 Finite element formulation ............................................................................................. 183

6.3.2 Stress-strain relationship of the reinforcement ............................................................... 184

6.4 ZERO-WIDTH INTERFACE BOND ELEMENT ........................................................ 185

6.4.1 Finite element formulation ............................................................................................. 185

6.4.2 Constitutive relationship for bond stress-slip ................................................................. 188

6.5 SOLUTION PROCEDURES......................................................................................... 196


6.6 SUMMARY ................................................................................................................... 198

CHAPTER 7 FINITE ELEMENT ANALYSIS OF EXPERIMENTAL WORKS 200


7.1 OVERVIEW................................................................................................................... 200
7.2 FINITE ELEMENT ANALYSIS OF UNIXIAL TENSION PRISMS........................... 200

7.2.1 Finite element configuration .......................................................................................... 200

7.2.2 Finite element results ..................................................................................................... 205

7.3 FINITE ELEMENT ANALYSIS OF FLEXURAL SPECIMENS................................. 216

7.3.1 Finite element configuration .......................................................................................... 216

7.3.2 Finite element results ..................................................................................................... 219

7.4 NEJADI AND GILBERT (2004) ................................................................................... 226

7.4.1 Finite element configuration .......................................................................................... 228


IV

7.4.2 Finite element results ..................................................................................................... 229

7.5 GRIBNIAK (2009) ........................................................................................................ 234

7.5.1 Finite element configuration .......................................................................................... 235

7.5.2 Finite element results ..................................................................................................... 238

7.6 SUMMARY ................................................................................................................... 242

CHAPTER 8 PARAMETRIC STUDY 244


8.1 OVERVIEW................................................................................................................... 244
8.2 NUMERICAL EXPERIMENTS.................................................................................... 245

8.2.1 Benchmark material properties....................................................................................... 245

8.2.2 Beam specimens............................................................................................................. 246

8.2.3 Slab specimens ............................................................................................................... 248

8.2.4 Experimental variables ................................................................................................... 249

8.3 DISCUSSION OF TEST RESULTS.............................................................................. 251

8.3.1 Bottom concrete cover, cb – Series A ............................................................................. 251

8.3.2 Concrete tensile strength, fct – Series B.......................................................................... 256

8.3.3 Concrete elastic modulus, Ec – Series C......................................................................... 260

8.3.4 Tensile reinforcement area, Ast – Series D...................................................................... 264

8.3.5 Tensile bar diameter, db – Series E ................................................................................ 268

8.3.6 Magnitude of creep, ijc – Series F .................................................................................. 270

8.3.7 Magnitude of shrinkage, sh – Series G .......................................................................... 274

8.3.8 Compressive reinforcement area, Asc – Series H............................................................ 278

8.3.9 Quantity of transverse reinforcement – Series I ............................................................ 282

8.4 SUMMARY ................................................................................................................... 283

CHAPTER 9 CONCLUSIONS AND RECOMMENDATIONS 286


9.1 SUMMARY ................................................................................................................... 286
9.2 CONCLUSIONS............................................................................................................ 287
9.3 RECOMMENDATIONS FOR FUTURE RESEARCH................................................. 291

Before cracking ............................................................................................................................ 293


V

After cracking............................................................................................................................... 294

Working Examples ....................................................................................................................... 297

Uncracked beam section............................................................................................................... 302

Fully-cracked beam section.......................................................................................................... 305

WORKING EXAMPLES ............................................................................................................ 306

Appendix A AN ANALYTICAL MODEL TO PREDICT TIME-DEPENDENT


TENSION STIFFENING IN UNIXIAL TENSION PRISM 293

Appendix B TIME-DEPENDENT ANALYSIS OF A CRACKED OR UNCRACKED


BEAM SECTION 302

REFERENCES 312
VI

LIST OF FIGURES

Figure 2-1: Tensile stress in a uniaxial concrete tension member. .......................................... 10


Figure 2-2: Local bond stress and slip distribution between two primary cracks at
crack stablised stage. ............................................................................................... 12
Figure 2-3: Macro response of tension stiffening in a uniaxial tension member. .................... 13
Figure 2-4: Cracked reinforced concrete beam. ....................................................................... 15
Figure 2-5: Testing rig of Beeby and Scott's tests.................................................................... 21
Figure 2-6: Alternative stress-strain diagrams for concrete and steel in tension to
represent tension stiffening in the finite element context. ...................................... 27
Figure 2-7: Tension chord model (Marti et al. 1998)............................................................... 31
Figure 2-8: Bond zone scale classification............................................................................... 32
Figure 2-9: Typical pull-out test configuration and its idealised continuum bond zone.
................................................................................................................................. 34
Figure 2-10: Cover-controlled cracks by Goto's experiment. .................................................. 36
Figure 2-11: The governing mechanisms for bond stress-slip relationship. ............................ 37
Figure 2-12: The confining actions on bond capacity.............................................................. 39
Figure 2-13: Bond stress-slip relationship of CEB-FIP bond model (CEB-FIP 1997)............ 42
Figure 2-14: Three failure stages of bond stress-slip relationship by Tepfer (1979). .............. 43
Figure 2-15: Creep isochrones. ................................................................................................ 48
Figure 2-16: Compliance function with different loading age t'. ............................................. 50
Figure 2-17: Development of drying shrinkage strain. ............................................................ 51
Figure 2-18: Assumption made in RCM. ................................................................................. 53
Figure 2-19: A gradually applied stress history at t  t 0 .......................................................... 55
Figure 2-20: Model for the role of solidification in creep: (a) Kevin Chain description
for viscoelastic component; (b) microscopic creep compliance functions of
the solidified matter; (c) schematic representation of the solidification
creep model. ............................................................................................................ 56
Figure 3-1: Time-dependent response of an axially loaded tension member .......................... 62
VII

Figure 3-2: Time-dependent deformation and stress on an uncracked section of the


uniaxial tension member. ........................................................................................ 64
Figure 3-3: Time-dependent deformation on an uncracked beam section. .............................. 66
Figure 3-4: Time-dependent deformation on a cracked beam section. .................................... 68
Figure 3-5: Time-dependent moment verse deflection response of a short-term beam........... 69
Figure 3-6: Time-dependent moment verse deflection response of a long-term beam............ 73
Figure 4-1: Specimen details of uniaxial tension members. .................................................... 76
Figure 4-2: Strain gauge detailing of uniaxial tension members. ............................................ 77
Figure 4-3: Set up for short-term uniaxial tension tests. .......................................................... 78
Figure 4-4: Set up for long-term uniaxial tension tests............................................................ 79
Figure 4-5: Creep and shrinkage companion tests. .................................................................. 82
Figure 4-6: Creep coefficients for the uniaxial tension tests.................................................... 85
Figure 4-7: shrinkage strain development for uniaxial tension tests........................................ 86
Figure 4-8: Average load P versus strain (Hs.avg) for STN12.................................................... 88
Figure 4-9: Average load P versus strain (Hs.avg) for STN16.................................................... 89
Figure 4-10: Variation of forces in steel and concrete at different loading stages
(STN12)................................................................................................................... 91
Figure 4-11: Variation of forces in steel and concrete at different loading stages
(STN16)................................................................................................................... 92
Figure 4-12: Crack numbers and locations (STN12). .............................................................. 93
Figure 4-13: Crack numbers and locations (STN16). .............................................................. 93
Figure 4-14: Average load P versus strain (Hs.avg) for STS12 .................................................. 96
Figure 4-15: Average load P versus strain (Hs.avg) for STS16 .................................................. 97
Figure 4-16: Variation of forces in steel and concrete at different loading stages
(STS12). .................................................................................................................. 99
Figure 4-17: Variation of forces in steel and concrete at different loading stages
(STS16). ................................................................................................................ 100
Figure 4-18: Crack numbers and locations (STS12).............................................................. 102
Figure 4-19: Crack numbers and locations (STS16).............................................................. 102
Figure 4-20: The effects of shrinkage on the average concrete tensile stress. ....................... 104
Figure 4-21: The effect of reinforcing ratio on the average concrete tensile stress. .............. 105
Figure 4-22: Tension stiffening strain versus axial strain (STN12 and STN16). .................. 106
VIII

Figure 4-23: Concrete force distribution of STN12 around crack NO. 3. ............................. 107
Figure 4-24: Steel stress versus maximum crack width (uniaxial tension short-term
tests). ..................................................................................................................... 108
Figure 4-25: Average strain versus load response of LTN12A. ............................................ 110
Figure 4-26: Average strain versus load response of LTN12C.............................................. 111
Figure 4-27: Initial and final concrete and steel stresses LTN12A (P = 40 kN).................... 111
Figure 4-28: Initial and final concrete and steel stresses LTN12C (P = 40 kN). ................... 112
Figure 4-29: Crack numbers and locations (LTN12A). ......................................................... 113
Figure 4-30: Crack numbers and locations (LTN12C). ......................................................... 113
Figure 4-31: Average strain versus load response of LTN12B.............................................. 115
Figure 4-32: Average strain versus load response of LTN12D. ............................................ 116
Figure 4-33: Initial and final concrete and steel stresses LTN12B (P = 20 kN). ................... 116
Figure 4-34: Initial and final concrete and steel stresses LTN12D (P = 20 kN).................... 117
Figure 4-35: Crack numbers and crack locations (LTN12B)................................................. 117
Figure 4-36: Crack numbers and crack locations (LTN12D). ............................................... 118
Figure 4-37: Tension stiffening strain versus time for long-term tested prisms. ................... 119
Figure 4-38: Maximum crack width versus concrete age of LTN12A and LTN12B. ........... 120
Figure 5-1: Dimension and detail of BSTN2-16 and BSTS2-16. .......................................... 123
Figure 5-2: Dimension and detail of BSTN3-16 and BSTS3-16. .......................................... 124
Figure 5-3: Dimension and detail of all slabs. ....................................................................... 124
Figure 5-4: Set up for short-term flexural tests. ..................................................................... 126
Figure 5-5: Set up for long-term flexural tests....................................................................... 127
Figure 5-6: High-resolution laser transducers underneath the short-term testing slabs......... 128
Figure 5-7: Dial gauges underneath the long-term testing slabs............................................ 129
Figure 5-8: Casting and curing of the slabs............................................................................ 130
Figure 5-9: Creep coefficients for the flexural tests............................................................... 133
Figure 5-10: shrinkage strain development for the flexural tests........................................... 134
Figure 5-11: Moment versus mid-span deflection (BSTN2-16) ............................................ 136
Figure 5-12: Moment versus mid-span deflection (BSTN3-16) ............................................ 136
Figure 5-13: Moment versus mid-span deflection (SSTN4-12)............................................. 137
Figure 5-14: Variation of forces in steel at different stages of loading (BSTN2-16). ........... 140
Figure 5-15: Variation of forces in steel at different stages of loading (BSTN3-16). ........... 141
IX

Figure 5-16: Variation of forces in steel at different stages of loading (SSTN 4-12)............ 142
Figure 5-17: Crack numbers and crack locations (BSTN2-16).............................................. 143
Figure 5-18: Crack numbers and crack locations (BSTN3-16).............................................. 144
Figure 5-19: Crack numbers and crack locations (SSTN-4-12)............................................. 144
Figure 5-20: Moment versus mid-span deflection for BSTS2-16.......................................... 147
Figure 5-21: Moment versus mid-span deflection for BSTS3-16.......................................... 148
Figure 5-22: Moment versus mid-span deflection for SSTS4-12. ......................................... 148
Figure 5-23: Variation of forces in steel at different stages of loading (BSTS2-16). ............ 150
Figure 5-24: Variation of forces in steel at different stages of loading (BSTS3-16). ............ 151
Figure 5-25: Variation of forces in steel at different stages of loading (SSTS4-12). ............ 152
Figure 5-26: Crack numbers and locations (BSTS2-16)........................................................ 154
Figure 5-27: Crack numbers and locations (BSTS3-16)........................................................ 154
Figure 5-28: Crack numbers and locations (SSTS4-12) ........................................................ 154
Figure 5-29: The effects of initial shrinkage on the tension stiffening deflections................ 157
Figure 5-30: The effects of reinforcing ratio on the tension stiffening deflections. .............. 157
Figure 5-31: The effects of initial shrinkage on the maximum crack width of flexural
members. ............................................................................................................... 159
Figure 5-32: The effects of reinforcing ratio on the maximum crack width of flexural
members. ............................................................................................................... 159
Figure 5-33: The change of mid-span deflection with time (SLTN4-12A) ........................... 161
Figure 5-34˖ the change of mid-span deflection with time (SLTN4-12B).......................... 161
Figure 5-35: The change of tensile bar forces with time (SLTN4-12A and SLTN4-
1B)......................................................................................................................... 163
Figure 5-36: Crack numbers and locations (SLTN4-12A)..................................................... 164
Figure 5-37: Crack numbers and locations (SLTN4-12B)..................................................... 164
Figure 6-1: Four-node plane isoparametric element. ............................................................. 169
Figure 6-2: Biaxial strength envelop of Foster and Marti (2003). ......................................... 173
Figure 6-3: stress-strain relationship of concrete ................................................................... 174
Figure 6-4: Discretization of continuous retardation spectrum of solidifying
constituent. ............................................................................................................ 179
Figure 6-5: Two node truss element....................................................................................... 183
Figure 6-6: Stress-strain relationship for the truss element.................................................... 185
X

Figure 6-7: Zero-width bond interface element. .................................................................... 186


Figure 6-8: CEB-FIP bond stress-slip relationship. ............................................................... 189
Figure 6-9: Parameter 2 versus steel stress s....................................................................... 192
Figure 6-10: Non-local average scheme of bond element...................................................... 193
Figure 6-11: Bond breakdown with the amount of shrinkage by the proposed model. ......... 196
Figure 6-12: One load step in Standard Newton Raphson Method........................................ 198
Figure 7-1: Finite element model configuration for uniaxial tension prisms......................... 202
Figure 7-2: Compliance function curves for uniaxial tension tests........................................ 203
Figure 7-3: Comparison between experimental and finite element results for the
short-term uniaxial tension tests (Load versus average strain). ............................ 206
Figure 7-4: Comparison between experimental and finite element results for the
short-term uniaxial tension tests (tension stiffening factor versus average
strain)..................................................................................................................... 207
Figure 7-5: Finite element results for STN12 and STS12 (local response). .......................... 209
Figure 7-6: Finite element results for STN16 and STS16 (local response). .......................... 210
Figure 7-7: Comparison between experimental and finite element results for the long-
term uniaxial tension tests (tension stiffening strain versus time). ....................... 213
Figure 7-8: Finite element results for LTN12A and LTN12C (local response)..................... 214
Figure 7-9: Finite element results for LTN12B and LTN12D (local response)..................... 215
Figure 7-10: Finite element mesh configuration for flexural tests......................................... 217
Figure 7-11: Compliance function curves for flexural tests................................................... 219
Figure 7-12: Comparison between experimental and finite element results for the
shot-term flexural tests (moment versus mid-span deflection). ............................ 221
Figure 7-13: Finite element results for BSTN2-16 and BSTS2-16 (local response). ............ 222
Figure 7-14: Finite element results for LTN12B and LTN12D (local response)................... 223
Figure 7-15: Finite element results for SSTN4-12 and SSTS4-12 (local response). ............. 224
Figure 7-16: Comparison between experimental and finite element results for the
long-term flexural tests (mid-span deflection versus time)................................... 225
Figure 7-17: Finite element results for SLTN4-12A and SLTN4-12B (local response).
............................................................................................................................... 226
Figure 7-18: Details of cross section for the slabs (Nejadi and Gilbert 2004)....................... 227
Figure 7-19: mid-span deflection versus time for s1-a and s2-a. ........................................... 230
XI

Figure 7-20: Tensile strain contours for s1-a and s2-a........................................................... 233
Figure 7-21: Bond stress distributions for s1-a and s2-a........................................................ 233
Figure 7-22: Longitudinal (a) and cross-sectional (b) reinforcement of the test beams;
notation of cross-section (c) (Gribniak 2009). ...................................................... 234
Figure 7-23: Set-up of creep tests and variation of creep coefficient with time .................... 237
Figure 7-24: Moment versus mid-span deflection response of s-1 and s-1R......................... 239
Figure 7-25: Stress and strain distributions at the mid-span sections of Gribniak’s
first series beams just before loading (FEM). ....................................................... 241
Figure 7-26: Tension stiffening deflection versus mid-span moment of Gribniak’s
first series beams (experimental results). .............................................................. 242
Figure 8-1: Construction of beam specimens in the numerical experiment........................... 247
Figure 8-2: Finite element model of beam specimens in the numerical experiment. ............ 248
Figure 8-3: Construction of slab specimens in the numerical experiment. ............................ 249
Figure 8-4: Finite element model of slab specimens in the numerical experiment................ 249
Figure 8-5: Instantaneous moment versus mid-span deflection responses of series A
short-term specimens............................................................................................. 252
Figure 8-6: Tension stiffening deflection versus mid-span moment responses of
series A short-term specimens............................................................................... 253
Figure 8-7: Mid-span deflection versus time responses of series A long-term
specimens. ............................................................................................................. 254
Figure 8-8: Instantaneous moment verse mid-span deflection responses of series B
short-term specimens............................................................................................. 257
Figure 8-9: Tensile strain contour of series B short-term specimens at M = 30kN.m. .......... 257
Figure 8-10: Tension stiffening deflection verse mid-span moment responses of
series B short-term specimens............................................................................... 258
Figure 8-11: Mid-span deflection versus time responses of series B long-term
specimens. ............................................................................................................. 259
Figure 8-12: Instantaneous moment verse mid-span deflection responses of series C
short-term specimens............................................................................................. 261
Figure 8-13: Tension stiffening deflection verse mid-span moment responses of
series C short-term specimens............................................................................... 262
XII

Figure 8-14: Mid-span deflection versus time responses of series C long-term


specimens. ............................................................................................................. 263
Figure 8-15: Instantaneous moment verse mid-span deflection responses of series D
short-term specimens............................................................................................. 265
Figure 8-16: Tension stiffening deflection verse mid-span moment responses of
series D short-term specimens............................................................................... 266
Figure 8-17: Mid-span deflection versus time responses of series D long-term
specimens. ............................................................................................................. 267
Figure 8-18: Instantaneous moment verse mid-span deflection responses of series E
short-term specimens............................................................................................. 268
Figure 8-19: Tension stiffening deflection verse mid-span moment responses of
series E short-term specimens. .............................................................................. 269
Figure 8-20: Mid-span deflection versus time responses of series E long-term
specimens. ............................................................................................................. 270
Figure 8-21: Instantaneous moment verse mid-span deflection responses of series F
short-term specimens............................................................................................. 271
Figure 8-22: Tension stiffening deflection verse mid-span moment responses of
series F short-term specimens. .............................................................................. 272
Figure 8-23: Mid-span deflection versus time responses of series F long-term
specimens. ............................................................................................................. 273
Figure 8-24: Instantaneous moment verse mid-span deflection responses of series G
short-term specimens............................................................................................. 275
Figure 8-25: Tension stiffening deflection verse mid-span moment responses of
series G short-term specimens............................................................................... 276
Figure 8-26: Mid-span deflection versus time responses of series G long-term
specimens. ............................................................................................................. 277
Figure 8-27: Mid-span deflection versus time responses of series G long-term
specimens. ............................................................................................................. 278
Figure 8-28: Instantaneous moment verse mid-span deflection responses of series H
short-term specimens............................................................................................. 279
Figure 8-29: Tension stiffening deflection verse mid-span moment responses of
series H short-term specimens............................................................................... 280
XIII

Figure 8-30: Mid-span deflection versus time responses of series H long-term


specimens. ............................................................................................................. 281
Figure 8-31: Instantaneous moment verse mid-span deflection responses of series I
short-term specimens............................................................................................. 282
Figure 8-32: Mid-span deflection versus time responses of series I long-term
specimens. ............................................................................................................. 283
Figure A-1: Axial load versus average axial strain for STN12 and STS12. .......................... 299
Figure A-2: Axial load versus average axial strain forLTN12A............................................ 301
Figure B-1: A singly- reinforced beam section under bending. 302
Figure B-2: Transformed section and initial strain distribution. 303
XIV

LIST OF TABLES

Table 2-1: Parameters defined in CEB-FIP bond model (CEB-FIP 1997) .............................. 42
Table 4-1: Testing procedure of the uniaxial tension test specimens. ..................................... 80
Table 4-2: Reinforcing bar properties. ..................................................................................... 83
Table 4-3: Instantaneous concrete properties for uniaxial tension tests................................... 85
Table 4-4: Measured loads and strains for STN12................................................................... 89
Table 4-5: Measured loads and strains for STN16................................................................... 89
Table 4-6: Crack widths at different loading stage (STN12). .................................................. 94
Table 4-7: crack widths at different loading stage (STN16) .................................................... 95
Table 4-8: Measured loads and strains for STS12. .................................................................. 98
Table 4-9: Measured loads and strains for STS16. .................................................................. 98
Table 4-10: Crack widths at different loading stage (STS12)................................................ 103
Table 4-11: Crack widths at different loading stage (STS16)................................................ 103
Table 4-12: Crack width at different stages (LTN12A) ......................................................... 113
Table 4-13: Crack width at different stages (LTN12C) ......................................................... 113
Table 4-14: Crack width at different stages (LTN12B). ........................................................ 118
Table 4-15: Crack width at different stages (LTN12D) ......................................................... 118
Table 5-1: Testing procedure of all the specimens ................................................................ 131
Table 5-2: Instantaneous properties of concrete (flexural specimens)................................... 132
Table 5-3: Deflections of short-term non-shrunk flexural members...................................... 137
Table 5-4: Average tensile forces in the concrete and the steel over the constant
moment region (BSTN2-16, BSTN3-16 and SSTN4-12)..................................... 139
Table 5-5: Crack width at different stages (BSTN2-16) ........................................................ 145
Table 5-6: Crack width at different stages (BSTN3-16) ........................................................ 145
Table 5-7: Crack width at different stages (SSTN4-12) ........................................................ 145
Table 5-8: Deflections of short-term shrunk flexural members. ............................................ 149
Table 5-9: Average tensile forces in the concrete and the steel over the constant
moment region (BSTS2-16, BSTS3-16 and SSTS4-12)....................................... 153
Table 5-10: Crack width at different stages (BSTS2-16)....................................................... 155
XV

Table 5-11: Crack width at different stages (BSTS3-16)....................................................... 155


Table 5-12: Crack width at different stages (SSTS4-12) ....................................................... 155
Table 5-13: The measured and calculated mid-span deflection for SLTN4-12A and
SLTN4-12B........................................................................................................... 162
Table 5-14: Comparison between fully cracked section and measured steel forces .............. 163
Table 5-15: Crack width at different stages (SLTN4-12A) ................................................... 165
Table 5-16: Crack width at different stages (SLTN4-12B).................................................... 165
Table 7-1: Parameters for the instantaneous constitutive law of concrete (uniaxial
tension specimens). ............................................................................................... 202
Table 7-2: Major creep solidification data for uniaxial tension tests..................................... 204
Table 7-3: The retardation times  and the correlated E for uniaxial tension tests.............. 204
Table 7-4: Parameters for concrete instantaneous constitutive law (flexural tests)............... 218
Table 7-5: Major creep solidification data for flexural tests. ................................................. 218
Table 7-6: The retardation times  and the correlated E of the Kevin chain units for
flexural tests. ......................................................................................................... 218
Table 7-7: Details of slab specimens (bar diameter, db = 12 mm). ....................................... 228
Table 7-8: Creep coefficient and shrinkage strain (x10-6) (Nejadi and Gilbert 2004). .......... 228
Table 7-9: Kelvin chain properties for solidification creep model for Nejadi and
Gilbert’s slabs........................................................................................................ 229
Table 7-10: Calculated tension stiffening deflections for Nejadi and Gilbert’s slabs. .......... 231
Table 7-11: Details of beam specimens (Gribniak 2009). ..................................................... 235
Table 7-12: Calculated tensile strength and elastic modulus of the concrete at the tests. ..... 236
Table 7-13: Shrinkage deformations of 280 × 300 × 350mm prisms (Gribniak 2009)
and the corresponding modeling parameters......................................................... 237
Table 7-14: Kelvin chain properties for solidification creep model for Gribniak’s
beams..................................................................................................................... 238
Table 7-15: Measured and modeled Cracking moments for Gribniak’s beams..................... 240
Table 8-1: Creep coefficient and shrinkage strain (x10-6) for the benchmark specimens.
............................................................................................................................... 246
Table 8-2: Kelvin chain properties for solidification creep model. ....................................... 246
Table 8-3: Various parameters for the beam specimens in the numerical experiment. ......... 250
Table 8-4: Various parameters for the slab specimens in the numerical experiment. ........... 251
XVI

Table 8-5: Tension stiffening deflections of series A long-term specimens.......................... 255


Table 8-6: max for the series B test specimens (All in MPa). ................................................ 256
Table 8-7: Tension stiffening deflections of series B long-term specimens. ......................... 259
Table 8-8: Tension stiffening deflections of series C long-term specimens. ......................... 263
Table 8-9: Tension stiffening deflections of series D long-term specimens.......................... 267
Table 8-10: Tension stiffening deflections of series E long-term specimens. ....................... 270
Table 8-11: Tension stiffening deflections of series F long-term specimens. ....................... 273
Table 8-12: Tension stiffening deflections of series G long-term specimens........................ 278
Table 8-13: Tension stiffening deflections of series H long-term specimens........................ 281
Table A-1: Experimental and theoretical average concrete stresses after cracking 299
CHAPTER 1 INTRODUCTION

1.1 GENERAL
Deflection and crack control are both critical design requirements that must be considered in

the design of reinforced concrete structures at the serviceability limit states. They are often the

critical design considerations, especially for lightly reinforced flexural members, such as one-

way and two-way slabs. With the advent of higher strength materials, thinner slabs and longer

spans are possible, but the slenderness of modern slabs is limited by the deflection and

cracking behaviour under service loads. Structural design engineers need a sound

understanding of the interaction between the concrete and the reinforcement under service

loads and must have access to reliable methods for including the effects of concrete cracking

and the time-dependent deformation caused by creep and shrinkage in their calculations.

The prediction of deflection and crack widths of flexural members is complicated by the

difficulties in estimating the time-dependent stiffness after cracking. After the concrete first

cracks, the load-deformation response of the member becomes nonlinear. The deflection and

the crack widths increase significantly with increasing load and with increasing time after

loading. This decrement of structural stiffness and the corresponding increment of deflection

are partly due to the loss of tension stiffening with time.


2

When a plain concrete element is subjected to tension, cracking occurs when the tensile stress

in the concrete reaches the tensile strength of concrete. After cracking, the tensile force

decreases dramatically. If the rate of deformation is controlled at cracking, a small residual

tensile force can be carried across the crack by friction arising from aggregate interlock.

After cracking in a reinforced concrete element, the tensile force is carried by the

reinforcement crossing the crack. Either side of the crack, the tensile force in the

reinforcement is resisted by bond at the concrete-steel interface. Reinforcement thereby

greatly improves the behaviour of concrete in tension.

Between the cracks in a reinforced concrete element, the tensile stress in the reinforcement

varies with part of the tensile force being transferred to the tensile concrete by bond, resulting

in the phenomenon known as tension stiffening. The bond-slip relationship between the steel

and the concrete greatly affects the transfer of tension to the concrete and the contribution of

the tensile concrete to the member’s stiffness. As a consequence, therefore, the bond-slip

relationship plays an important role in the determination of service load behaviour, including

the spacing of cracks, the crack widths, and the deflection, both in the short-term and in the

long-term.

Tension stiffening has been defined variously as the contribution of the intact concrete

between the primary cracks to the stiffness of the member or the ability of the intact concrete

between the primary cracks to carry part of the resultant tensile force. It is particularly

significant in relatively lightly reinforced members, where the actual stiffness may be several

times larger than the stiffness calculated on the basis of fully-cracked cross-sections, where

the tensile concrete is ignored and only the embedded tensile reinforcement is taken into

consideration. For slabs containing small quantities of tensile reinforcement (typically


3

ȡ<0.003), tension stiffening may be responsible for more than 50% of the total stiffness of the

member ˄ Gilbert 2007 ˅ . It is generally agreed that tension stiffening decreases with

increasing external force. However, the decay of tension stiffening is not very well understood,

particularly the change in tension stiffening with time due to the combined effects of time-

dependent cracking, bond deterioration, tensile creep in the concrete and drying shrinkage. It

is generally agreed that tension stiffening reduces with time to about 50% of its initial value,

but this has not been satisfactorily confirmed and the magnitude and rate of change of tension

stiffening with time are uncertain.

Not many experimental studies of the time-dependent change of tension stiffening in

reinforced concrete elements have been reported in the literature. In the extensive

experimental program conducted by Beeby and Scott (2002), a number of axially loaded

reinforced concrete prisms were tested subjected to sustained service loads, and the concrete

and steel strains were measured with time. However, they did not measure the concrete drying

shrinkage both before and during the period of sustained loading and so a reliable

interpretation of the measurements is not possible. Consequently, Beeby and Scott’s

explanation for the decay of tension stiffening with time is not convincing. Drying shrinkage

plays a significant role in the time-dependent behaviour of reinforced concrete and it is a

crucial consideration in the study of the change in tension stiffening with time (Bischoff 2001).

Meanwhile, different numerical approaches have been developed to represent tension

stiffening in the analysis of reinforced concrete structures. Due to the lack of understanding of

the time-dependent behaviour of tension stiffening, most of these approaches try to model

tension stiffening based on experience. For example, Eurocode 2 (1992) proposes a method

based on using a distribution factor to obtain the effective stiffness and gives reduction factors
4

(0.5) to account for the decrease of bond with time on the decay of tension stiffening.

Therefore, it is suggested by the author that a more accurate numerical model must be

developed to predict time-dependent tension stiffening. The method must take appropriate

account of both creep and shrinkage, particularly the latter.

In the finite element context, tension stiffening is often modeled using a smeared crack

approach by adjusting the constitutive relationship for concrete in tension to include an

unloading branch after cracking (Barros et al. 2001, Ebead and Marzouk 2005, Gilbert and

Warner 1978, Guptam and Maestrini 1990, Lin and Scordelis 1975, Prakhya and Morley 1990,

Nayal and Rasheed 2006). Alternatively, tension stiffening has been included by adjusting the

constitutive relationship for the tensile reinforcement (Gilbert and Warner 1978). An

alternative approach is to model cracking as discrete discontinuities in the concrete and to

model the bond stress on the interface of steel and concrete as a function of slip between them.

Since tension stiffening is basically generated from the bond, this approach seems to be a

more realistic model of the state of stress and strain between the cracks in the tension zone.

This research program aims to offer a clear picture of the variation of tension stiffening with

time and the associated mechanisms, and hence give a better approach to model this

behaviour. In this thesis, several series of laboratory experiments investigating the time-

dependent change in tension stiffening are described and the results are presented. Reinforced

concrete prisms were tested in axial tension, including both short-term tests where the prisms

were gradually loaded to failure and long-term tests where the prisms were subjected to

prolonged periods of sustained service load. Short-term and long-term tests were also

undertaken on simply-supported reinforced concrete beams and slabs. The details of the

experimental work and results are presented to give a clear picture of the variation of tension
5

stiffening with time. The influences of creep, shrinkage, and cracking on the deterioration of

tension stiffening with time are carefully investigated and discussed. In addition, a two

dimensional non-linear finite element program, with a proposed bond model developed

concurrently with the experimental program, is presented. The proposed bond model is based

on the CEB-FIP bond model but incorporates other factors, such as steel and concrete stress,

concrete cracking and shrinkage. The program has been used to analyse the experimental

work and to conduct a parametric study, in which various parameters that may affect time-

dependent tension stiffening are carefully investigated.

1.2 OBJECTIVES OF THE CURRENT STUDY


The major part of the research program described in this thesis is an experimental

investigation of the short-term and long-term behaviour of cracked reinforced concrete,

including the mechanisms of tension stiffening and the change in tension stiffening with time

due to changes in either the sustained load or with increasing load. The thesis also outlines the

development of a numerical method that can accurately predict the time-dependent change in

tension stiffening. Accordingly, the specific objectives of the study are:

1 To undertake a range of laboratory experiments to investigate the time-dependent short-

term and long-term behaviour of tension stiffening in prisms subjected to uniaxial tension,

and in beams and slabs subjected to flexure, and to identify the mechanisms associated

with the loss of tension stiffening with time.

2 To develop a two dimensional finite element model, together with appropriate

constitutive laws, that can be used to predict the behaviour of cracked reinforced concrete

members and to satisfactorily model tension stiffening, both in the short-term and in the
6

long-term.

3 To demonstrate the applicability of the numerical model by comparing the finite element

results with the laboratory results.

4 To further investigate parameters affecting the time-dependent tension stiffening and

deflection by conducting numerical experiments using the finite element package.

1.3 THESIS STRUCTURE


In Chapter 2, a general review is given of the current state of knowledge related to tension

stiffening on both macro and micro scales. A literature review is also presented of previous

work on the development of numerical models of the bond mechanism, and the creep and

shrinkage characteristics of concrete.

Chapter 3 presents the author's understanding of tension stiffening behaviour and the factors

affecting the change of tension stiffening with time. The general responses of uniaxial tension

members and flexural members are discussed. The effects of creep and shrinkage are included

in the discussion.

The experiments undertaken at the Randwick Heavy Structures Laboratory of the University

of New South Wales are presented in Chapters 4 and 5. Details of the specimens, the test set

up, the construction and testing procedures and the measurements taken are described.

Shrinkage before or after cracking is an important parameter that was carefully measured

during these experiments. The results from the uniaxial tension tests and flexural tests are also

presented in these chapters.

In Chapter 6, material constitutive laws for concrete and steel bars are introduced in a two
7

dimensional finite element program. The formulation of each elements and the numerical

solution scheme are also presented. In addition, a bond-interface model is proposed to model

the change in bond between the concrete and the steel bar.

In Chapter 7, the finite element model introduced in Chapter 6 is used to analyse the tested

specimens in the experimental program and the numerical results are compared with the

experimental results. The model is further verified by analysing test specimens presented

elsewhere in the literature.

Chapter 8 consists of a series of numerical experiments conducted by using the finite element

program. A parametric investigation is undertaken to further study the factors affecting time-

dependent tension stiffening. The results are discussed and the influence of each parameter on

time-dependent tension stiffening is examined.

Chapter 9 summarises the entire thesis and presents the major findings of the study.

Suggestions for future research are also made.


CHAPTER 2 CURRENT STATE OF
KNOWLEDGE

2.1 OVERVIEW
This chapter gives a review of the current state of knowledge relating to tension stiffening.

The chapter starts with a description of tension stiffening in both uniaxial tension members

and flexural members under instantaneous monotonic loading, (the effects of creep and

shrinkage on the time-dependent behaviour will be discussed in Chapter 3). Section 2.3

provides a review of the previous experimental work investigating and characterizing tension

stiffening, the majority of which is considering only the behaviour under short-term loading.

Section 2.4 describes the methods that are used to predict tension stiffening in practice,

including the procedures specified in building codes and those in the finite element context.

Section 2.5 and 2.6 summarise and discuss previous work on the bond mechanism and the

effects of creep and shrinkage, as these factors are deemed to have a significant effect on the

time-dependent tension stiffening in reinforced concrete members. The basic characteristics of

bond, creep and shrinkage, and the numerical modeling methods are also reviewed in Section

2.5 and 2.6.


9

2.2 TENSION STIFFENING UNDER


INSTANTANEOUS LOAD
2.2.1 Micro and Macro behaviour in uniaxial tension members

Tension stiffening is the ability of tensile concrete to contribute to the stiffness of a reinforced

concrete member. There are a number of different ways to represent this phenomenon in

practice. Previous studies have considered tension stiffening in a reinforced concrete member

at either a macro scale or a micro scale. In this thesis, the micro scale is referred to as the level

concerning the stress (strain, or slip) distribution between two primary cracks. The macro

scale is based on a spatially larger level that explicitly contains several primary cracks and

implicitly considers member stiffness and the global average structural response.

Consider an axially loaded singly-reinforced concrete tension member (Figure 2-1a). The

member is subjected to an incremental uniaxial tensile force P at both ends of the protruding

bar, as shown. The tensile force is gradually transferred from steel bar to surrounding concrete

by bond at the concrete-steel interface. Before cracking, the concrete tensile stress increases

linearly as the applied load is increased. At an applied load Pcr, first cracking occurs at the

weakest cross-section and this is usually assumed to occur when the concrete tensile stress

reaches the lower characteristic value of the direct tensile strength fct. The concrete stress at

this cracked section suddenly drops to zero and at this section the entire force is carried by the

tensile steel bar (Figure 2-1b). Within a distance S from the crack, the concrete stress is

gradually built up due to the bond at the concrete-steel interface, until at distances greater than

S, the concrete stress is no longer affected by the crack.


10

a) A singly-reinforced concrete prism

Pcr Pcr

s s
Vc

b) At first crack

P>Pcr P>Pcr

Vc

c) At crack stabilised stage

Figure 2-1: Tensile stress in a uniaxial concrete tension member.

A relatively small increase in load will cause a second crack to develop at a cross-section at

some distance x t S from the first crack, thereby reducing the concrete stress in the vicinity of

that crack. Eventually, under increasing load, primary cracks form at somewhat regular

intervals along the member and the primary crack pattern is established. The concrete tensile

stress c at each crack is zero, rising to a maximum value (less than the tensile strength of the
11

concrete) mid-way between adjacent cracks, as shown in Figure 2-1c. For normal strength

concrete, this happens when P is in the range 1.2 to 1.6Pcr.

After the crack pattern stabilises and no more primary cracks form, the average tensile stress

in the concrete decreases progressively under increasing load. This phenomenon can be

attributed to internal cracks or cover-controlled cracks developing at the concrete-steel

interface. The cover-controlled cracks are the internal cracks that radiate from the bar

deformations and are contained within the concrete cover (Goto 1971). The bond stress is

reduced by the formation of cover-controlled cracks in-between two primary cracks and hence

the tensile stress in the concrete also gradually diminishes.

Figure 2-2 shows the micro responses, including concrete stress, bond stress and slip

distributions at the steel-concrete interface between two primary cracks. Bond stress is zero at

the cracked sections, because the concrete and the steel bar are not in contact. The slip at the

primary crack is at a maximum and decreases to zero mid-way between the two cracks, as

shown. Bond stress increases rapidly adjacent to each crack and then decreases to zero mid-

way between the cracks. After the crack pattern has stabilised, an increment of external load

causes a local decrease of the bond stress due to cover-controlled cracks and a widening of

existing primary cracks associated with increasing slip. Normally the deterioration of bond

stress due to cover-controlled cracks is fairly localised, however, in the figure, it is

represented as a smeared average decrease of bond stress between the two cracks.
12

Figure 2-2: Local bond stress and slip distribution between two primary cracks at crack
stablised stage.

Figure 2-3a shows the load versus average elongation response of a uniaxial tension member.

This is the macro-scale behaviour of tension stiffening under short-term load. The macro

response consists of three phases, namely linear elastic, primary crack formation and cover-

controlled cracking (stabilised crack) stages. Also plotted in the diagram is the bare bar

response, where the concrete is assumed to carry no tension and the entire tensile force is

carried by the steel bar at every section throughout the length of the member. The
13

instantaneous tension stiffening in the diagram can be denoted as the instantaneous tension

stiffening strain tsi, which is the difference between the bare bar strain and the actual member

strain at any particular load P (Figure 2-3a). The tension stiffening strain is a maximum tsi.max

at first cracking Pcr and then gets progressively smaller as the external load increases.

a) Load verses average strain response

b) Tension stiffening factor versus average strain


Figure 2-3: Macro response of tension stiffening in a uniaxial tension member.
14

Fields and Bischoff (2004) introduced the concept of a tension stiffening factor or bond factor t

which can be defined as the tension stiffening strain at load P (tsi) divided by the maximum

tension stiffening strain tsi.max at P = Pcr. Figure 2-3b shows a typical relationship between the

tension stiffening factor and the average member strain, where t reaches its maximum value

(1.0) at first cracking and then diminishes with increasing strain eventually approaching zero as

the strain becomes larger.

The overview of the micro and macro instantaneous responses of tension stiffening in a

uniaxial tension member shows that they are compatible with each other. The global response

is attributed to the local mechanisms of bond and cracking. The decrease of bond stress due to

both primary cracking and cover-controlled cracking triggers the decrease of concrete stress

locally and hence tension stiffening strain globally.

2.2.2 Micro and Macro behaviour in flexural members.

Figure 2-4a shows a typically cracked portion of a reinforced concrete beam subjected to an

applied moment Ms, greater than the cracking moment Mcr. The bottom fibres are in tension

while the top fibres are in compression. At the cracking moment, the tensile stress reaches the

flexural tensile strength of the concrete on the tension side of the beam, and primary cracks

occur. As moment increases, the cracks stabilise at a spacing about S. Under the moment Ms

even though the moment is constant along the beam, the second moment of area varies

depending on the crack locations and the curvature values along the beam. The second

moment of area of the cross-section is higher at the sections away from the cracks and lower

close to the cracks. The position of neutral axis is therefore not parallel to the geometric axis

of the member, as shown. However, it is usual to assume an average neutral axis location that

is at a constant depth below the compressive surface of the member (Figure 2-4a).
15

a) Singly reinforced beam under pure bending

b) Instantaneous moment versus deflection response

Figure 2-4: Cracked reinforced concrete beam.

The distribution of concrete tensile stress, bond stress and slip in the tensile zone in-between

the cracks are usually considered to be similar to those of the uniaxial tension member, as

shown in Figure 2-1 and 2-2. When the direct tensile strength of concrete in the weakest

section of the tensile zone is reached at moment Mcr, and the first crack occurs, the tensile

stress of concrete at the cracked section reduces to zero. The concrete tensile stress gradually

increases away from the cracks as the bond stress develops. Additional primary cracks form
16

with increasing applied moment until the primary crack pattern has stabilised. Thereafter

cover-controlled cracks appear at the concrete-steel interface and these locally break down the

bond stress and hence reduce the tensile stress in the concrete.

An area of concrete Act symmetrically located around the tensile steel can be defined in the

cracked regions of a beam to carry the concrete tensile stress c between the cracks and to

model the tension stiffening effect in flexural members. The dimensions of this average

tensile concrete area have been proposed by Clark and Speirs (1978), Gilbert (1983), Castel et

al. (2006) and others. Gilbert indicated that Act is dependent on the magnitude of the

maximum applied moment Ms, the area of the tensile reinforcement Ast, the amount and the

types of tensile steel bars, the tensile strength of concrete fct, and the loading type (short-term

or long-term loading). Clark and Speirs suggested that Act is also affected by tensile bar

diameters db and concrete cover cb. Other parameters, such as the location of the neutral axis

and the location of the reinforcements in the tensile area of the beam, were taken into account

by Castel et al. to calculate Act.

Globally, tension stiffening in reinforced concrete flexural members can also be interpreted in

terms of bending moment M versus deflection  (or average curvature ) response, as shown

in Figure 2-4b. The cracking development phases shown in the graph are analogous to that of

the uniaxial tension member. As soon as the bending moment exceeds the cracking moment

Mcr, a first crack forms and the response starts to become nonlinear. Tension stiffening in this

case can be characterized as the instantaneous tension stiffening deflection (tsi) between the

actual response and the fully-cracked response. The fully-cracked response is obtained by

assuming that every cross-section is fully-cracked and the flexural rigidity is EcIcr, where Ec is

the elastic modulus of concrete and Icr is the second moment of area of a fully-cracked cross-
17

section. In this case, tension stiffening factor t can be represented as the ratio between the

tension stiffening deflection tsi (or the average curvature tsi) at any applied moment and its

maximum value tsi.max (or tsi.max) at first cracking (Bischoff 2005). Experimental

investigations suggest that the instantaneous tension stiffening in beams or slabs decreases

with increasing moment. This is illustrated in Figure 2-4b where the actual response

approaches the fully-cracked response as the moment increases.

2.3 PREVIOUS EXPERIMENTAL INVESTIGATION


ON TENSION STIFFENING
2.3.1 Short-term experiments

Many short-term tests have been carried out to investigate the instantaneous behaviour of

reinforced concrete members after cracking. Many of these tests were undertaken on axially

loaded tension members where tension stiffening can be most directly measured. Most of

these tests have not considered the influence of initial concrete shrinkage prior to loading,

which may greatly affect the measured test results (Bischoff 2001). Chapter 3 provides a

detailed explanation of the effects of creep and shrinkage on the short-term tension stiffening.

In additional to the effect of initial shrinkage on the experimental results, there are other

factors that may significantly influence short-term tension stiffening behaviour. Early

experimental work by Clark and Cranstown (1979) showed that the decrease of tension

stiffening under short-term loading is not only a function of strain but also bar spacing in

reinforced concrete slabs. They suggested the decay of tension stiffening was greater when the

bars were widely spaced in slabs. This is more significant when the bar spacing exceeds a

critical value of 1.5 times the slab depth. Kishek (1983) undertook an experimental study on
18

the effect of orientation of reinforcement on tension stiffening in slabs under uniaxial bending.

The angle between the reinforcing bars and the principal bending moment direction varied

from 0 to 60 degrees. The results show that the tension stiffening increases as the angle of

reinforcement with the principal bending moment increases. This is attributed to the fact that

with greater angle, the tensile stresses in the reinforcement in the principal bending direction

are smaller. Furthermore, the experimental study by Sooriyaarachchi et al. (2005) showed that

there was no significant influence on tension stiffening when the bar diameter was varied

provided the reinforcement ratio remained unchanged.

Based on their experiments, Al-Fayadh (1997) and Abrishami and Mitchell (1996) concluded

that the use of high-strength concrete can increase tension stiffening but mainly at the crack

formation stage and the linear-elastic stage. However, they found that this influence decreases

gradually during the crack stabilisation stage. A similar conclusion was reported by Ouyang et

al. (1997) who found that the tension stiffening effect declines faster in high-strength concrete

members than in the normal-strength members as the average strain increases. They suggested

that this may be attributed to the more brittle cracking process in high-strength concrete.

Abrishami and Mitchell also indicated that crack spacing in high strength concrete specimens

is larger than in normal strength concrete specimens.

Gilbert (2007) discussed the tension stiffening deflections that were obtained from short-term

tests on a number of slabs with different reinforcing ratios. He concluded that for members

with a relatively small reinforcement ratio, tension stiffening provides a large proportion of

the postcracking stiffness of the slabs at moments up to 1.3Mcr.

In addition, other experimental works (Bischoff and Paixao 2004) suggested that the tension

stiffening contribution to the overall stiffness is greater in members reinforced with glass fibre
19

reinforced polymer (GFRP) than in members containing steel reinforcement. This is because

the GFRP reinforced concrete has higher cracking strain than the steel reinforced concrete,

thus the crack pattern in FRP reinforced concrete may not even be stabilised at service loads.

This thesis is only concerned with the tension stiffening in steel reinforced concrete elements,

and the FRP reinforced concrete is not considered further.

2.3.2 Long-term experiments

Relatively few experimental programs have been undertaken to investigate the long-term and

time-dependent behaviour of tension stiffening. As early as 1970, Stevens (1970) conducted a

series of long-term beam tests in which the specimens were subjected to sustained load for up

to 2 years, and the change of concrete and steel strains and the development of cracking were

recorded throughout the tests. Stevens concluded that the tensile force in the concrete is about

one third of that developed just before cracking, and it reduced with time at a reducing rate.

He suggested that the rate of reduction is akin to that of the relaxation of concrete stress in the

compression zone of a beam under sustained load.

The bond break-down with time due to internal cracks or cover-controlled cracks was first

discussed by Illston and Stevens (1972). In their study, a technique of drilling holes from each

crack on the side faces near the level of reinforcement was adopted. A vacuum pump was used

to draw the resin through the cracks to measure the internal crack width. The crack width at

the surface of the beam was 2.5 times that at the bar. However, this ratio reduced to 20% after

2 years of sustained loading. The beams under sustained loading also exhibited about 30%

loss of bond after 1.25 year. They suggested that under sustained load there is a length of bond

breakdown near the surface crack, which can cause the opening of existing surface cracks (or

primary cracks), initially restrained by the steel bars. It is interesting to note that the
20

increasing load can result in the instantaneous breakdown of bond and hence the loss of

tension stiffening, while under constant sustained loading, breakdown of bond is also

observed experimentally. The mechanism for this long-term decay of bond or tension

stiffening can be attributed to the time-dependent properties of the material, particularly the

deflection caused by concrete shrinkage and the restraint to shrinkage provided by the

embedded reinforcement.

The most relevant and recent experimental work was undertaken by Beeby and Scott (2002).

They specifically investigated tension stiffening in uniaxial tension members at different

sustained load levels. In the macro scale, they used Demec gauges to measure the average

elongation over several primary cracks; in the micro scale, strain gauges were used to measure

the local variation of strains along the tensile steel bars. They concluded that tension

stiffening decreases to about 50% of its initial value very rapidly (within 20 days) under

sustained loading. They attributed this phenomenon to several mechanisms including creep

and shrinkage of concrete, and development of internal cracks. Figure 2-5 depicts the test set

up for the experiment conducted by Beeby and Scott.


21

Figure 2-5: Testing rig of Beeby and Scott's tests.

2.3.3 Comment on the previous experimental works on tension stiffening

According to the literature, there are a variety of factors that may change tension stiffening in

reinforced concrete, including concrete strength, bar size, bar spacing, bar orientation, creep,

and shrinkage. However, most of the tests mentioned in the preceding sections did not

measure the creep and shrinkage characteristics of the concrete used in the experimental

programs. It has been conclusively demonstrated that initial shrinkage can greatly affect the

measured tension stiffening response in reinforced concrete tension members (Bischoff 2001).

Therefore, in the absence of measurements of shrinkage and creep, the reliability of many of

the previous tests is in doubt. The loss of tension stiffening is largely due to the breakdown of

bond and cracking of the concrete, both of which are affected by the development of

shrinkage in the concrete. Drying shrinkage can cause a significant increase in existing crack

widths and a decrease in the average crack spacing due to the formation of new crack (Nejadi
22

and Gilbert 2004a). Therefore, the effect of shrinkage on tension stiffening can not be ignored

and must be considered in the experimental program.

Despite the previous work, the magnitude and the rate of change of tension stiffening under

sustained load are still not well understood. Illston and Stevens found that the loss of bond is

only about 30% after 1.25 years of sustained load while Beeby and Scott detected a 50% loss

of tension stiffening within the first 20 days. The tremendous difference in these observations

may be ascribed to the neglect of shrinkage in the experimental evaluation and also the

different loading histories.

The behaviour of tension stiffening, especially its time-dependent responses and mechanisms

are not well understood, and there exists a range of opinions on the matter. The aim of the

current research is to give a clear picture of tension stiffening and its variation with time. As

there are a very limited number of tests incorporating the effects of shrinkage and creep on

tension stiffening reported in the literature, an experimental program with several uniaxial

tension and flexural members has been conducted in this study. The development of creep and

shrinkage has been monitored with time. The full details of the program are discussed in

Chapters 4 and 5.

2.4 CURRENT METHODS FOR PREDICTING


TENSION STIFFENING
2.4.1 AS3600-2009 and Eurocode 2 Methods

Various methods are specified in building standards to consider tension stiffening in the

calculation of deflection or crack width. The Australian Standard (AS3600-2009) adopts the

well-known Branson’s equation (Branson 1968), which takes account of tension stiffening by
23

calculating the so-called effective stiffness that is an empirical estimate of the member

stiffness between the fully-cracked stiffness and the uncracked stiffness, depending on the

level of moment at the critical sections. The Eurocode 2 (2004) provides an approach that

utilises a distribution coefficient to calculate post-cracking deformation according to the stress

state in the reinforcement under consideration. Both of these approaches are based on a

modified stiffness method, using an average or “smeared” moment-curvature relationship for

the member.

2.4.1.1 AS3600-2009 Method

In the simplified deflection calculation procedure in AS3600-2009, the effective second

moment of area Ief of a beam or a slab after first cracking is obtained from Equation 2.1. The

short-term deflection or curvature can then be derived based on the elasticity theory, using the

specified elastic modulus at the age of first loading.

3
§M ·
I ef I cr  I  I cr ¨¨ cr* ¸ d I ef . max
¸ 2.1
© Ms ¹

where Icr is the second moment of area of the cracked transformed section, which depends on

the modulus ratio and reinforcing ratio; I is the second moment of area of the uncracked

cross-section about the centroidal axis; Ms* is maximum bending moment at the section, based

on the short-term serviceability load or construction load; and M cr is the cracking moment.

M cr Z f cf‘  V cs  P / Ag  Pe t 0 2.2

where Z is the section modulus of the uncracked section, referring to the extreme fibre at

which cracking occurs; f’cf is the characteristic flexural tensile strength of concrete; cs is the

maximum shrinkage-induced tensile stress on the uncracked section at the extreme fibre at
24

which cracking occurs, and for a singly reinforced section may be taken as:

2.5 U w  0.8 U cw
V cs E s H cs* 2.3
1  50 U w

where cs* is the final design shrinkage strain; w is the web reinforcement ratio for the tensile

steel (= (Ast + Apt) / bwd); and cw is the reinforcement ratio for the compressive steel (= Asc /

bwd ); Asc and Ast are compressive and tensile reinforcement area respectively; and Apt is the

area of any prestressing steel in the tension zone. Equations 2.2 and 2.3 indicate that the

cracking moment Mcr is affected by the shrinkage strain of concrete and the reinforcement

ratio of the cross section.

For long-term deflection calculation of a reinforced concrete member, a deflection multiplier

kcs is specified as kcs >2  1.2 Asc / Ast @ t 0.8 , to account for the effects of creep and shrinkage,

where Asc / Ast is taken at — (a) mid-span, for a simply supported or continuous beam; or (b)

the support, for a cantilever beam. In the absence of more refined methods, the long-term

deflection of a beam or slab many be taken as kcs times the short-term deflection due to the

sustained load considered.

Equation 2.1 is also specified in the ACI code (ACI318-08) but in that standard no account is

taken of shrinkage (ie. cs = 0). Consequently ACI318-08 can overestimate stiffness and

underestimate deflection (Gilbert 1999).

2.4.1.2 Eurocode 2 Method

The approach adopted by Eurocode 2 is to calculate the curvatures at typical sections and then

determine the deflection by integration. The curvature of a section after first cracking is

calculated by:
25

N 9N cr  1  9 N uncr 2.4

where is the distribution coefficient, = 1 - 12 (Mcr/Ms)2; 1 is a coefficient that takes

account of the bond property of the reinforcement (1.0 for ribbed bars and 0.5 for plain bars);

2 is the coefficient that take accounts for the duration of loading or of repeated loading (1.0

for single, short-term loads and 0.5 for repeated or sustained loads); cr is the curvature at the

section ignoring concrete in tension; uncr is the curvature on the uncracked transformed

section; and Ms is the in-service moment.

2.4.1.3 Evaluation of the AS3600-2009 and Eurocode 2 approaches

It is generally agreed that both of these approaches give reasonable results for members with

reinforcing ratio in excess of about 0.5%. However, compared with experimental member

response of lightly reinforced members ( <0.005), tension stiffening is overestimated using

Branson’s equation. This is because when the reinforcement ratio is low, the maximum

moment Ms* in Branson’s equation is not significantly greater than the cracking moment Mcr,

and Ief remains close to the uncracked value I. Thus, the instantaneous deflection is

underestimated after first cracking (Bischoff 2005, Ghali 1993, Gilbert 2007). In contrast, the

Eurocode approach provides more accurate and reliable results particularly for lightly

reinforced concrete members (Gilbert 2007).

In order to compensate for this limitation of Branson’s equation Gao et al. (1998), Ghali et al.

(2001), Toutanji and Saafi (2002) proposed modifications to Branson's equation for deflection

prediction by introducing empirical factors that effectively decrease the ratio between gross

moment of inertia and cracked moment of inertia. Bischoff (2005) proposed an alternative

expression for Ief which may be derived from the Eurocode 2 approach (Equation 2.4).

Substituting the notations in equation 2.1 into equation 2.4 and rearranging gives the
26

following formula:

I cr
I ef
§ I ·§ M · 2 2.5
1  E 1 E 2 ¨1  cr ¸¨ cr ¸
¨ I ¸¹© M ¹
©

In terms of long-term deflection, AS3600-2009 method considers the influence of shrinkage

on the decay of tension stiffening by the inclusion of cs in Equation 2.2. Compared with

AS3600-2009, Eurocode 2 approach predicts the final value of tension stiffening as half of the

short-term value by setting the coefficient 2 = 0.5, when the specimen is under sustained load.

Under sustained loads, creep and shrinkage complicate the response of a cracked member, and

these are the most difficult issues to account in design. The simplified design methods in the

various codes account for long-term loss of tension stiffening, but do not specifically describe

how creep and shrinkage affects the behaviour.

2.4.2 Modelling Tension Stiffening in the Finite Element Context

In a finite element context, methods to take account of tension stiffening can be generally

categorized into two groups: (1) modifying the constitutive relationship of the tensile concrete

(Barros et al. 2001, Ebead and Marzouk 2005, Gilbert and Warner 1978, Guptam and

Maestrini 1990, Lin and Scordelis 1975, Prakhya and Morley 1990, Nayal and Rasheed 2006);

(2) modifying the constitutive relationship of the tensile reinforcement (Gilbert and Warner

1978). As shown in Figure 2-6, the former approach models tension stiffening with an

adjusted softening curve for concrete in tension, and the degree of softening is greater in the

finite element layer that is closer to the reinforcement layer. Perfect bond is usually assumed

between concrete and steel elements, even though in reality bond stress is always
27

accompanied by slip and the variation of micro-cracking at the concrete-steel interface.

a) Gradually unloading response of the tensile b) Modified stress-strain diagram for tensile
concrete after cracking (Lin and Scordelis) steel after cracking (Gilbert and Warner)

Figure 2-6: Alternative stress-strain diagrams for concrete and steel in tension to represent
tension stiffening in the finite element context.

The latter approach introduces an additional stress carried by the tensile steel bars that is

dependent on the strain levels, and assumes that concrete after cracking carries no stress

anywhere. It can be seen that both approaches do not correctly represent the true mechanisms

of tension stiffening. Therefore, these approaches are not adopted herein for the numerical

analysis in the thesis.

A more realistic method for modeling tension stiffening is to use a suitable bond-slip

relationship at the concrete-steel interface. The action of bond produces tensile stress in the

concrete around the bar. Balazs (1993) used the ascending branch of the bond-slip relationship

from CEB-FIP Code to describe the bond zone and obtain the steel stress distribution between

the cracks. However, the formula that he gives to calculate crack width or spacing is not

computationally friendly. Instead of assuming a bond stress versus slip relationship, Somayaji
28

and Shah (1981) assumed a function to represent the bond stress distribution between the

cracks in order to obtain crack width and to simulate tension stiffening. A constant bond stress

distribution within the cracked zone of a reinforced concrete member is often assumed. This

simplifies the mathematics (Floegl and Mang 1982, Marti et al. 1998).

The tension chord model developed by Marti et al. assumes that the bond stress is constant

and therefore that the steel stress distribution along the tension element is linear as shown in

Figure 2-7a. The concrete stresses are zero at the crack, and away from the cracks tensile

force is transferred from the reinforcement to the surrounding concrete by bond. The method

was extended to model the cracked membrane element (Kaufmann 1998, Kaufmann and

Marti 1998) for finite element implementation. For ordinary ribbed bars a stepped, perfectly

rigid-plastic bond shear stress-slip relationship was proposed in tension chord model, as

illustrated in Figure 2-7b. Marti et al suggested b0 = 2.0fct and b1 = fct before and after bar

yielding, while Gilbert (2006) recommended that b0 = 2.5fct in the short-term, and b0 =

1.25fct when the final maximum crack width is to be determined.

For the tension chord element between two consecutive cracks spaced at Srm shown in Figure

2-7a, the steel stress at the crack sr can be expressed as a function of the average strain m

W b 0 S rm
V sr EsH m  , ( V sr d f y ) 2.6a
db

W b 0 S rm W b1 S rm § W b 0 E s · E s 2
S rm
 f y  EsH m ¨  ¸ W b 0W b1 2
db d b ¨© W b1 E sh ¸¹ E sh db
V sr fy  2 , 2.6b
W b0 E s

W b1 E sh
29

( V s. min d f y  V sr )

§ f y · W b1 S rm
V sr f y  E sh ¨¨ H m  ¸ ( f y  V s. min )
© E s ¸¹ db 2.6c

Where db is the bar diameter, Esh = (fu – fy)/(u – fy/Es) = hardening modulus of steel, and s.min

is the minimum steel stress between the cracks. Note that in serviceability condition, steel

stress normally remains below the yielding stress, hence only Equation 2.6a is applicable.

The maximum crack spacing occurs when c.max = fct, and is given as

f ct d b
S r . max 2.7
2W b 0 U

where = Ast / Act is the reinforcement ratio of the tension chord element and db is the

reinforcing bar diameter. The crack spacing in a fully developed crack pattern is limited by

S r . max
d S rm d S r . max 2.8
2

Before the yielding of the steel, the average instantaneous steel strain is obtained from the

steel stress distribution

1
Hm T  0.375 f ct Act 2.10
E s Ast

The tension stiffening strain is obtained by subtracting the average strain from the bare bar

strain at the cracks, and according to the tension chord model it is constant after all the

primary cracks have stabilised, because the bond stress stays constant.
30

0.375 f ct Act
H tsi H sr  H m 2.11
E s Ast

The tension chord model simplifies the analysis by assuming the bond stress is constant,

hence giving rise to linear concrete and steel stress distributions. In reality the bond stress

distribution is nonlinear and significantly dependent on the level of steel stress and the extent

of slip. A more realistic model that can be used to correctly predict tension stiffening or

concrete tensile stress must incorporates a local bond stress-slip model that takes account of

these parameters. In this way, the interaction between the concrete and the steel bar can be

appropriately modeled and the tensile stress in the concrete will not remain constant after first

cracking. In the next section, a review and discussion is given of the bond stress-slip

mechanism, and the various existing models available in the literature.


31

a) Local stress distribution

b) Bond stress-steel stress relationship

Figure 2-7: Tension chord model (Marti et al. 1998).


32

2.5 BOND-SLIP ON THE CONCRETE-STEEL BAR


INTERFACE
2.5.1 Bond Zone Scales

There are generally three different scales in which the bond behaviour is characterized and

modeled in engineering practice, namely the rib scale, bar scale and member scale, (Lowe

1999) as illustrated in Figure 2-8. In the member scale, the whole member is considered as an

individual element in the analysis. The reinforcing bar is treated as being smeared over the

whole element and perfect bond between concrete and steel is assumed. The bond behaviour

can hardly be looked into in the member scale, as the combination of shear, flexure or torsion

on the member can affect the accurate investigation. In addition, this model can hardly be

applied in a finite element context as it can only simulate one particular structural element.

a) Member scale

b) Reinforcing bar scale

c) Rib scale

Figure 2-8: Bond zone scale classification.

The rib scale, at the other extreme, characterizes the bond response by discretizing the bond

zone around each rib. The bond effect in this scale is greatly affected by concrete behaviour at
33

the vicinity of the bar, such as fracture and crushing of concrete mortar and aggregates, load

transferred between concrete mortar and aggregate, etc. The bond behaviour is also greatly

influenced by the non-homogeneous response as the internal cracks on the top of the ribs

develop. The difficulty to investigate bond effects at this scale is that the material non-

homogeneous properties of concrete mortar and aggregates can rarely be precisely

demonstrated in the experiments. For the implementation of finite element, the rib-scale

model may require a sophisticated meshing solution and algorithm, and a need to introduce

material non-homogeneity in the model, which creates great complexity for the analysis.

Therefore, numerous pull-out tests examining the bond response have been conducted at the

scale of the reinforcing bar, and the mechanisms of bond at this scale have been widely

investigated. Most of the pull-out tests have been undertaken with embedment lengths

significantly larger than the rib spacing (usually over 6db). The bond response is assumed in

continuum within the bond zone, where the material properties of the concrete and the steel

bar are usually assumed to be homogenous. The size of the bond zone is normally smaller

than the crack spacing and therefore no primary cracks are allowed to form and disturb the

tests. External tensile force is imposed at the ends of the steel bar, and the parameters, such as

concrete strength, bar diameter, etc, on bond behaviour are considered. A typical example is

shown in Figure 2-9 with the usual idealization shown in Figure 2-9b.

The idealised continuum and the measured average bond stress-slip relationship greatly

simplify the problem and give good approximation of the global response of the whole

structure. Furthermore, the bond model identified in this scale can be directly applied to the

finite element model, together with the application of bond link elements. Accordingly, the

local bond stress-slip relationship that is proposed in this thesis is based on the steel bar scale.
34

a) Localized damage on the vicinity

b) Idealized continuum bond zone in the experiments

Figure 2-9: Typical pull-out test configuration and its idealised continuum bond zone.

2.5.2 Bond Mechanisms

The characteristics of bond zone response are associated with the localised damage that

occurs through three distinct mechanisms acting in parallel during the loading process (Lutz

and Gergely 1967), namely chemical adhesion, friction, and mechanical interlock between the

concrete and steel bar. The first two mechanisms are primarily responsible for the bond

response of plain bars, where roughness of the bar surface is the primary bond strength

parameter. In this case, when the bond stress at typical section of the member exceeds the

chemical adhesive and friction resistance, bond failure occurs.

For deformed steel bars, the mechanical interlock between concrete and steel has proven to be
35

the dominant mechanism. This phenomenon was demonstrated by Lutz et al. (1966). They

concluded that the load transferred between steel and concrete is significantly affected by

mechanical interaction. As the bar deformations bear onto the concrete, local fracture

(splitting) of concrete occurs. The presence of splitting crack can significantly leads to the

loss of tension stiffening, as demonstrated by Abrishami & Mitchell (1996). Lutz and Gergely

(1967) proposed that the slip can be induced by the wedging action at the front of the rib. A 30

to 40 degree effective rib face angle is formed by the crushing of concrete. Goto (1971)

investigated the wedging action by injecting ink into cracks around the reinforcing bar when

the specimens were loaded to the maximum force. Observation showed that besides primary

cracks, cover-controlled cracks (radial cracks) on the interface are initiated and inclined to the

bar axis at angle between 45 and 80 degrees (60 degrees on average). As the cover-controlled

cracks initiate, the bearing effect becomes the major mechanism causing damage to the bond

zone, where relatively little slip occurs. The bond stress induced by friction between concrete

and steel is a minimum in this phase. Splitting cracks along the steel bar axis form and

propagate from the steel interface to the surface of the concrete specimen. These splitting

cracks result from circumferential stresses within the concrete at the vicinity of the bond

interface. This is caused by the wedges bearing on the concrete at the front of each rib and the

radial stress perpendicular to the bar axis (Tepfers 1979). As soon as the circumferential stress

exceeds the tensile strength of concrete, longitudinal splitting crack forms. Therefore the bond

strength is closely related to the tensile strength of concrete.


36

Figure 2-10: Cover-controlled cracks by Goto's experiment.

Figure 2-11 shows the major mechanisms that govern the bond stress-slip relationship at

different stages (fib 2000). As the bond stress stays below 20% to 80% of the tensile strength

of concrete, bond stress is primarily provided by the chemical adhesion and friction. The bond

stress-slip response is relatively stiff at this stage. For plain bar, as soon as the adhesion

breaks down, bond failure occurs. In the case of deformed bars, after the chemical adhesion is

broken down, mechanical interlock with the ribs bearing onto the concrete becomes the

governing mechanism. The steel bar starts to crush the concrete and internal radial cracks start

to form from the tip of the ribs onto the surrounding concrete, and the bond stress-slip

response begins to become non-linear. As soon as the splitting cracks reach the surface of

concrete where little confinement is provided, bond failure occurs suddenly. If sufficient

confinement is provided around the bond zone (usually by transverse reinforcement), splitting

failure is delayed. After bond stress reaches its maximum value, it drops off with increasing

slip. Mechanical interaction decreases at this stage and the bond stress is mainly provided by

the residual friction on the rough surface between the concrete and steel. Eventually, due to
37

the loss of shear resistance from surface friction, the steel bar can be completely pulled out

from the concrete block.

Figure 2-11: The governing mechanisms for bond stress-slip relationship.

2.5.3 Characteristics of Bond Stress-Slip Relationship

Many experimental programs have been undertaken to evaluate the effects of a wide range of

parameters on bond, such as confinement (transverse steel, confinement pressure), steel

properties (bar sizes, anchorage length, bar spacing), and concrete properties (concrete

strength, concrete cover), etc. Few of these studies considered all parameters determining

bond effect, furthermore the testing methods and boundary conditions of the tests often varied

considerably.

It is generally agreed that bond strength increases as the concrete strength increases. It is
38

recognized that the bond strength is approximately proportional to the square root of concrete

compressive strength f cp (Eligehausen et al. 1983). In addition, Eligehausen et al.’s tests

also indicated that bond strength also depends on the bar diameter. The bond strength of a

19mm bar is 10% greater than that of 25mm bar, while the bond strength of 31mm bar is 10%

lower than that of 25mm bar. Morita and Fuji (1985) also reported that the measured slip in

their experiment was proportional to the bar diameter. The larger-size reinforcing bar is able

to induce greater slip.

The steel stress state and concrete stress state have also been shown to significantly affect the

bond capacity. Shima et al. (1987) conducted an experimental study on steel reinforced

concrete cylinders loaded in tension, with an anchorage length of 50db. Bar strain was measure

using foil gauges glued to the bar along the length without removing the ribs from the

reinforcing bar. Slip was obtained from the integral of the measured strain along the length of

the bar. They suggested that the local strain in the steel was an indicator of the surrounding

concrete state, and the value of steel and concrete strain or stress varies along the bar

according to the embedded length of the pull-out test. The local bond-slip relationship along

the reinforcing bar can also differ according to the boundary conditions of the tests (Kankam

1997). Most of the other experimental work assumes a constant average bond stress along a

relatively short embedded length. The influence of steel stress on local bond response has not

been widely investigated experimentally.

The bond characteristics in the post-yielding range of the steel bar are also dependent on steel

stress or strain state. Under typical in-service conditions, the steel strain is less than yield

strain, and therefore the bond response after the yielding of steel is not discussed further here.
39

Figure 2-12: The confining actions on bond capacity.

Bond capacity is greatly affected by confining pressure on the bond zone. Confining pressure

reduces radial stresses that lead to splitting in the bond zone (Tepfers 1979). If good

confinement is provided, the bond response can be very significantly enhanced and bar pull-

out can result without splitting. Eligehausen et al. (1983) concluded that the pull-out failure

response in a confined specimen is associated with higher bond strength and bond capacity

than the splitting failure response in a non-confined specimen.

The confinement actions can be attributed to the surrounding concrete, the transverse

reinforcement and the lateral pressure. Giuriani et al. (1991) evaluated previous pull-out tests

with transverse reinforcing bars and proposed the minimum amount of transverse

reinforcement that is sufficient to prevent slitting failure of the pull-out test. Early

investigations on the influence of lateral pressure on bond strength (Untrauer and Henry 1965)

demonstrated that the increase of bond strength due to external pressure is more significant at

higher stress states, and the increase is proportional to the product of square root of the

normal pressure and the square root of the concrete strength. A similar conclusion has also
40

been found by other researchers (Navaratnarajah and Speare 1987; Robin and Standish 1984).

Apart from the bond stress-slip relationship, another significant bond response is the radial

stress-displacement relationship. The measurement of radial stress has not often been made in

previous experimental studies. Two of the most comprehensive tests regarding this issue were

conducted by Gambarova et al. (1989) and Malvar (1992). In the former study, a preformed

splitting crack with prescribed opening on the specimen was maintained under the monotonic

loading process, thus the radial deformation was maintained constant and the confinement

stress was monitored. Malvar applied constant confining stress on the external surface of

concrete under increasing load until splitting cracks occurred, during which the dilation or

radial deformation was recorded. Gambarova et al. suggested that with increasing crack

opening, the radial stress reached its maximum value when the slip slightly exceeded the

value of slip corresponding to the shear bond strength, but the peak value of the confinement

stress is not strongly dependent on the opening of the splitting crack. Malvar concluded that

the dilation of the bond zone is a reaction to the damage of concrete surrounding the bond

zone, and thereby can be represented by the concrete material state. In addition the radial

expansion of the pull-out specimens decreases with increasing confinement pressure. The

influence of radial stress (or confinement pressure) is more obvious when the tests reach the

post-cracking range.

The previous experimental studies have also indicated that bond response can be affected by

the deformation pattern of the ribbed bar, including the rib spacing, height, face angle and

pattern. It is found that the bars with increased relative rib area (the rib area perpendicular to

the bar axis normalized by the bar surface area between ribs) exhibit increased bond stiffness

(Hamad 1995). In addition, splitting failure can be prevented with the use of bars with larger
41

rib area. Darwin and Graham (1993) concluded that the bond stress-slip relationship was a

function of the relative rib area of the bars, and independent of the specific combination of rib

height and rib spacing.

2.5.4 Bond Models

In this section, numerical models developed for the bond stress-slip relationship are reviewed

and discussed, such as CEB-FIP bond model, Bigaj’s bond model, Lowes’ model,

thermodynamic model and plasticity bond model. The sophistication of each model depends

on the way the bond zone mechanical interaction is simulated and the parameters that are

incorporated in the model. Each model is associated with different algorithm for finite

element implementation.

One of the most commonly used models is the non-linear relationship proposed by CEB-FIP

Mode Code (1997) and illustrated in Figure 2-13. The bond stress-slip relationship under

monotonic loading is decomposed into four phases. The bond stress in each phase is given by

identifying certain slip parameters s1, s2, s3 in the diagram (Figure 2-13). The first phase

(ascending part) is defined as a power function for the slip range s<s1, followed by a plateau

b = max for s1< s < s2. The curve then decreases linearly to the residual frictional bond

resistance f for a slip range s2< s < s3. For s > s3, the bond stress is assumed constant and

equal to f (Table 2-1). In order to distinguish the splitting bond failure and pullout bond

failure, CEB-FIP bond model simply specifies different values for these key stresses and slips

for the unconfined and confined states as shown in Table 2-1.


42

Bond Stress Slip Range

0.4
§ s1 · for 0 d s d s1
W b W max ¨¨ ¸
¸
© s2 ¹
W b W max for s1 d s d s 2

§ s  s2 ·
W b W max  W max  W f ¨¨ ¸ for s 2 d s d s 3
¸
© s3  s2 ¹
Wb Wf for s 3 d s

Figure 2-13: Bond stress-slip relationship of CEB-FIP bond model (CEB-FIP 1997).

Table 2-1: Parameters defined in CEB-FIP bond model (CEB-FIP 1997)

Unconfined Confined
Bond conditions Bond conditions
Good Others Good Others
S1 0.6mm 0.6mm 1.0mm
S2 0.6mm 0.6mm 3.0mm
S3 1.0mm 2.5mm Clear rib space
0.4 0.4
max 2 f cp f cp 2.5 f cp 1.25 f cp

f 0.15max 0.4max

Tepfer (1979) proposed a so-called thick-walled cylinder to characterize the bond zone. The

interaction between the concrete and the steel bar is represented by the shear stresses and

accompanied radial stresses. In his theory, three failure stages are involved in the bond stress-

slip relationship, namely uncracked elastic stage, partly cracked stage, and entirely cracked

stage (Figure 2-14), which characterises the propagation of radial stress from the interface.

Den Uijl and Bigaj (1996) extended Tepfers’ theory by incorporating a softening behaviour of

concrete caused by the internal radial cracks, where cracked concrete is defined on the basis

of the fictitious crack model (Hillerborg 1983) and the bilinear softening model by Roelfstra
43

and Wittmann (1986), rather than assuming it has no capacity after cracking as Tepfers

suggested. The number of radial cracks, which is related to the total radial and circumferential

displacements, is calibrated as a model parameter. The wedging effect is modeled by

assuming dry friction at the interface, so that the bond stress is proportional to radial stress

with the orientation of the interface. To distinguish the bond responses under splitting failure

and pull-out failure, Den Uijl and Bigaj suggested that the angle between radial displacement

and slip is a constant value. In the case of pull-out failure, they proposed that the cone angle

decreases during the development of a sliding plane on the concrete-steel interface.

a) Uncracked b) Partly cracked c) Fully cracked

Figure 2-14: Three failure stages of bond stress-slip relationship by Tepfer (1979).

Lowes et al (2004) introduced a series of multipliers to the bond-slip relationship, in order to

define the influence of confining pressure, concrete damage state, steel strain state, slip

history, bar size and bar spacing. The product of these multipliers contributes to the total peak

bond strength. For the convenience of finite element implementation, a non-local model is

used to include the adjacent concrete and steel material states into the bond elements. One of

the deficiencies of the model is that the multiplier to account for the damage of surrounding
44

concrete is defined based on their own concrete constitutive model. Thus it can hardly be used

with other concrete models, unless this multiplier is modified according to the corresponding

concrete model.

Ragueneau et al. (2006) describes the constitutive model of the bond interface elements based

on a thermodynamic framework. The model also couples continuum damage and plasticity

theory to consider the cracks parallel to the bar direction and inelastic strain due to sliding.

The mode- fracture failure and the frictional sliding are coupled, with a single damage

variable in the Helmholtz free energy description, to represent the energy transformed from

mechanical interaction to frictional sliding. An evolution law is required to identify the

variation of the damage yield function, within the elasto-damage domain. For the frictional

sliding part, the constitutive relationship is represented by the plasticity yield function. The

non-associated flow-rule is incorporated for the kinematic hardening of the material.

The plasticity-based bond model (Cox and Herrmann 1998) is established based on the

assumed yield surface, hardening and softening rule and non-associated flow rule of the bond

stress-slip relationship. Similar to the thermodynamic bond model, the fundamental factor that

determines the evolution of the bond response in this model is attributed to a damage factor.

The damage factor here is a function of the plastic slip (slip which occurred during the initial

splitting phase of the loading) and the yield function and the evolution law. The yield criteria

of the model consist of an exponential criterion (for large damage value) and a power criterion

(for small damage value) as a linear combination through a weighting function, which is

dependent of the magnitude of the damage factor. Therefore the yield surface changes from a

power criterion to an exponential criterion as the damage mechanism progresses. A tentative

non-associated flow rule (Cox 1994) is used to represent the bond stress versus radial dilation
45

response, which takes place near the peak bond stress. This radial dilation reduces as the

confinement pressure increases. A strong point of this model is that it does not require

establishing the geometric relationship between radial and shear stress. Instead, these two

bond stresses are automatically related through the variation of the plastic work in terms of

the non-associated flow rule.

2.5.5 Time-dependent bond response

There were quite few experimental works that focused on the long-term bond-slip response.

Under low-level sustained loading, creep can cause additional slip on the concrete-steel

interface. Creep therefore softens the stiffness of the ascending part of bond stress-slip

relationship. This primary mechanical aspect of bond creep at the steel-concrete interface has

been discussed by Franke (1976) and Svensvik (1981).

Rostasy and Kepp (1982) attempted to characterize the parameters of bond creep by their

long-term pull-out tests. Their testing series were classified with different variables, such as

concrete strength, sustained bond stress before the instantaneous bond strength, history of

bond stress, and relative area of ribs on the bar. They concluded that bond creep is

independent of concrete strength and bond stress states. In addition, the effect of aging on

bond creep can be ignored if the first loading age is greater than 35 days. This can greatly

simplify the numerical analysis. They presented the irrecoverable creep slip in terms of a

linear creep strain theory, where the bond creep displacement is proportional to the elastic

creep displacement by a bond creep coefficient b(t).

s sc t s siM b t 2.12

Where ssc(t) is the creep slip, ssi is the instantaneous slip induced by a certain level of short-
46

term bond stress, and ijb(t) is the bond creep coefficient, which performs non-linearly with

respect to time. If the creep coefficient in Equation 2.12 is non-aging, it can be depicted as a

function of elapsed time under load alone.

Koch and Balazs (1993) suggested that the time-dependent slip is higher, if the applied pull-

out force is higher, which indicates that the long-term slip is associated with internal or

external actions, as well as time.

In a sufficiently long specimen, the breakdown of bond with time is primarily the result of

restraint to concrete shrinkage by the reinforcing bar (or other boundary conditions) and the

effects of bond creep are secondary. If the restraint to concrete shrinkage is great enough,

cracks (either primary cracks or cover-controlled cracks) will start to propagate and cause a

local degrading of bond stress. This process is further complicated since it is also related to

the steel stress and concrete stress state, as addressed by Koch and Balazs. The experimental

work in the literature was mainly involved with small size specimens with embedment length

much less than the crack spacing. They do not reflect the influence of local damage of bond

due to shrinkage and cracking. A more comprehensive and realistic long-term bond stress-slip

relationship must consider the effect of shrinkage on the loss of bond at different steel stress

or concrete stress states.

2.6 TIME EFFECTS OF PLAIN CONCRETE


In this section, the property and characteristics of creep and shrinkage in plain concrete are

reviewed. Expressions for creep and shrinkage strains, together with the instantaneous

deformation are also presented.


47

2.6.1 Concrete as a time-dependent material

The concrete strain at any time at a point in a specimen at constant temperature can be

expressed as the sum of instantaneous İe(t), creep İcp(t) and shrinkage İsh(t) components.

Instantaneous and creep strains are induced by the stress in the concrete whereas shrinkage

strain is stress-independent. Hence the total strain in a specimen is the superposition of the

three components

H (t ) H e (t )  H cp (t )  H sh (t ) 2.13

where t is the time under consideration.

If the concrete element is under constant stress  and it remains less than about 0.4 of the

compressive strength (which usually is the case under service condition), creep strain is

proportional to the stress and the creep isochrones are approximately linear (Figure 2-15). At

higher stress level, creep strain increases at a faster rate than the applied stress and becomes

non-linear with respect to the stress. Therefore, at lower stress level, one can calculate the

creep strain approximately proportional to the stress level and Equation 2.13 can be expressed

as

H (t ) VJ t , t '  H sh (t ) 2.14

in which J(t,t’) is the compliance function representing the mechanical strain (instantaneous

plus creep) at time t caused by a unit constant stress that has been acting since time t’. J(t,t’)

can also be expressed as the sum of elastic compliance function 1/E(t’) and the creep

compliance function C(t,t’)


48

1  M (t , t ' )
J (t , t ' ) 1 / E t '  C t , t ' 2.15
E (t ' )

where E(t’) is the elastic modulus characterizing the instantaneous deformation at age t’; and

ij(t,t’) is called the creep coefficient which is the ratio between the creep strain and the

instantaneous strain at time t.

/fcp
1.0
Failure
limit

t=0

t=
0.5

Linear
creep range

Concrete strain

Figure 2-15: Creep isochrones.

Research on creep (Bažant and Wittmann 1983) showed that creep in the concrete is mainly

due to the breakage of bonds and contacts between colloidal sheets in the hydrated cement gel.

There are generally two types of creep, namely basic creep and drying creep. The former is

the creep strain occurring in a sealed moist environment, due to the breakage of bonds at

highly stressed sites within the colloidal microstructure of the calcium silicate hydrate gels in

the hardened cement paste. Pickett (1942) first observed drying creep, known as the Pickett

effect. The Pickett effect describes the fact that with drying, the deformation of a concrete

specimen under sustained load can exceed the sum of the drying shrinkage deformation of a

load-free specimen and of the deformation of a specimen that does not dry, i.e. is sealed. The
49

basic mechanism for drying creep is that drying can cause creep strain per unit stress to

become larger and thus produce additional deformation. Bažant and Xi (1994) suggested that

a major portion of the Pickett effect is attributed to the micro bond breakage within the

microstructure of calcium silicate hydrates by the drying effect, which can further increase

creep strain rate with time (Bažant et al. 1997).The other factors that may influence drying

creep were found and given by Bažant and Wu (1974), Bažant and Chern (1985), Wittmann

and Roelfstra (1980).

Creep is affected by many factors, including the maximum aggregate sizes, the water-cement

ratio, the concrete strength, the type of cement, the property of aggregates, temperature,

humidity, the degree of hydration, pore water content, specimen size and stress state, etc. The

detailed explanation of the effects of these parameters on creep can be found in the following

literature (Bazant et al. 1982; Bazant et al. 1976; Bažant and Chern 1985; Bažant and

Wittmann 1983; Domone 1974; Gamble and Parrott 1978; Gamble 1982; Gilbert 1988;

Lorman 1940; Marechal 1972; Neville et al. 1983).

Equation 2.15 explicitly indicates that an important property of concrete creep is aging, which

means that creep strain is greater when the age of the first loading is reduced. The creep rate

under sustained load is greater during the first few weeks for concrete loaded at an early age.

The aging of creep is caused by the chemical process of hydration in the concrete, which can

produce tri-calcium silicate hydrate gel and gradually fill the pores of hardened cement paste

hence stiffening the microstructure within the concrete and greatly reducing the creep strain.

Figure 2-16 shows the typical curves for the development of J(t,t’) with different loading ages

t’. J(t,t’) increases at a faster rate in the concrete loaded at earlier ages.
50

J(t,t')
t'=3days
30days

300days

3000days

30000days

1/E0
1 hour 30 years
log(t-t')

Figure 2-16: Compliance function with different loading age t'.

For concrete in tension, previous studies (Brooks and Neville 1977; Illston 1965) illustrate

that the compliance is generally observed to be as large as for uniaxial compression. However,

the mechanism and magnitude of tensile creep are still not well quantified. In this thesis, the

magnitude and rate of development of tensile creep is assumed to be identical to the

compressive creep at the lower stress levels.

Analogous to the creep strain, shrinkage strain can also be categorized into two different types:

drying shrinkage and autogenous shrinkage. Drying shrinkage commences as soon as the

concrete specimen begins drying. It results from increasing capillary tension of pore water

and solid surface tension of pore walls, caused by diffusion of pore water out of the specimen.

When the concrete specimen is immersed in water, swelling occurs, which is normally an

order of magnitude less than drying shrinkage and thus can not compensate for the drying

shrinkage. Autogenous shrinkage is the change of specimen volume caused by the chemical

hydration of concrete in a moisture-sealed condition. Autogenous shrinkage is relatively small

compared to drying shrinkage. It is only about 5% of the total drying shrinkage. The
51

autogenous shrinkage development stops when the concrete dries to a relative humidity 80%

or less in the pores and the chemical hydration stops.

sh

Upper bound

sh (drying shrinkage)

t0 (start of drying) t
Figure 2-17: Development of drying shrinkage strain.

Thus shrinkage strain is affected by the water-cement ratio, the relative humidity, the ratio of

exposed surface area to volume, the aggregate, etc. It develops very fast in the first several

months, especially at the beginning of drying, thereafter it increases at a decreasing rate. It is

generally agreed that shrinkage strain approaches a finite upper bound as time approaches

infinity (Figure 2-17).

2.6.2 Mathematical Modeling of Creep

2.6.2.1 Principle of superposition

The modeling of creep and its aging effect in a concrete structure is often a complex

procedure. Due to the time-varying stress history in almost every reinforced concrete

specimen, one can not directly derive creep strain at a typical time by equation 2.14. An

alternative way is to integrate the creep strain caused by the stress history d(t’). Each

infinitesimal stress at t’ produces the mechanical strain at time t according to its compliance

function d(t’) = J(t,t’)d(t’). The total strain at time t caused by a variable stress history can
52

then be obtained by summing these infinitesimal strains together with the stress-independent

strain (i.e. shrinkage strain)

t
Ht ³ J t , t ' dV t '
0
 H sh t 2.16

Equation 2.16 (Stieltjes integral) relies on the principal of superposition. Volterra (1913) first

introduced this method for aging material analysis, and McHenry (1943) first applied it to

concrete. The principal of superposition is equivalent to the hypothesis of linearity of the

constitutive equation that relates the stress and strain histories. Differentiating equation 2.16

gives the strain rate

V t t
wJ t , t '
H t ³ dV t ' 2.17
Et 0
wt

It has been demonstrated that the principle of superposition can give accurate results as long

as the following conditions are satisfied (Bazant 1988b; Gilbert 1988):

1. Concrete stress is lower than 0.4;

2. Unloading does not take place;

3. There is no significant change in moisture concrete distribution;

4. There is no large sudden stress increase after initial loading.

One major disadvantage of this approach is that a large amount of creep data is required.

2.6.2.2 Rate of Creep Method (RCM)

Due to the complication of creep aging effect, simplified methods, such as rate of creep

method (RCM) (Dischinger 1937; Whitney 1932) have been developed. The RCM assumes

parallel creep curves for concrete with different loading ages (Figure 2-18). That is the rate of
53

change of creep is assumed to be independent of the age at first loading. Therefore, only one

creep versus time curve is needed to obtain the creep strain for different stress histories. The

creep strain at time t caused by a constant stress  applied at time t’ can be calculated by

subtracting the creep strain induced from t0 (age at first loading) to t’ from the total creep

strain at t

V
H cp (t , t ' ) >M (t , t 0 )  M (t ' , t 0 )@ 2.18
Ec

Then the compliance function is given by

1 M (t , t 0 )  M (t ' , t 0 )
J t, t'  2.19
Et Ec

Substituting this creep strain into equation 2.17 gives:

1 V t
H t V t  M t , t 0 2.20
Et Ec

cp

F
(t1,t0)

B E

Actual
Response

(t1,t') D
Assumed
Response

A C
t0 t' t Time

Figure 2-18: Assumption made in RCM.


54

Integrating equation 2.20 with time or using the method of superposition, one can obtain the

total mechanical strain at time t. The greatest advantage of this method is that it requires no

previous stress history if the elastic modulus is given as a constant. However, larger error may

occur using this method, particularly for a decreasing stress history (Bazant and Najjar 1973).

2.6.2.3 Age-adjusted Effective Modulus Method (AEMM)

One of the simplest methods to predict creep in concrete for manual calculation is to adopt

some type of effective modulus. The convenience of this method is that it does not require a

step-by-step numerical integration. McMillan (1916) first introduced the Effective Modulus

Method (EMM), in which the creep for time t is obtained by transforming equation 2.15 to

obtain a so-called effective modulus:

1 E (t ' )
E e (t , t ' ) 2.21
J t, t ' 1  M (t , t ' )

However, this simplification obviously ignores the fact that equation 2.15 is only applicable

when the stress is held constant, which is normally not the case for the long-term response of

structures. EMM fails to take the aging effect into account and this method is only useful for a

single step stress history. If the concrete stress varies with time, this method yields inaccurate

solutions.

To improve the applicability of EMM, a so-called Aged-adjusted Effective Modulus Method

(AEMM) was introduced and formulated by Trost (1967) and Bazant (1972). This method

takes the aging effect on creep into account by simply reducing the creep coefficient ij(t,t0) if

the stress is gradually applied. An aging coefficient (t,t0) (with the range from 0.6 to 0.9) is

used and the magnitude of the reduced creep coefficient is given as (t,t0)ij(t,t0). Then the

creep strain at time t due to a gradually applied stress (t) at the time interval t – t0 can be
55

expressed as:

'V t
H cp (t ) F t , t 0 I t ,t 0 2.22
E t0

Therefore, for a continuous stress history, as shown in Figure 2.21, by assuming the change of

stress due the time interval t - t0 as (t) = (t0) – (t), the total strain at time t can be

calculated as the sum of strains produced by (t0), the strains produced by (t) and the

shrinkage strain.

V t0
H (t ) >1  M t , t 0 @  'V t >1  F t , t 0 M t , t 0 @  H sh t
E t0 Et
V t0 'V t 2.23
  H sh t
Ee t , t 0 Ee t , t 0

E t0
in which E e t , t ' is called the age-adjusted effective modulus.
1  F t, t0 M t, t0

(t 0)

(t0)
(t)

t0 t time

Figure 2-19: A gradually applied stress history at t  t 0 .

The aging coefficient F (t , t ' ) is dependent of the age at first loading, the duration of load, the

size and shape of the specimens and the rate of change of stress and so on (Gilbert 1988). For
56

a relaxation problem, it is approximately 0.8 when the time reaches infinity. For manual

calculation, AEMM is the most satisfactory alternative considering both efficiency and

accuracy.

2.6.2.4 Solidification creep theory

Volume fraction growth as a measure of aging

An improved basic creep formulation based on the principal of superposition is called the

theory of solidification (Bažant and Prasannan 1989a). The method describes all the

viscoelastic behaviour of concrete, including aging, by a series of Kevin chains. The aging

aspect of creep is due to growth of the volume fraction v(t) of the load-bearing portion of

solidified matter (i.e. hydrated cement), the properties of which are assumed to be age-

independent. The time-dependence of any material’s properties in this case is treated as a

consequence of a time-varying composition of the material, which is characterized by v(t).

g(v,t)

(a)

(c)
(b)
Figure 2-20: Model for the role of solidification in creep: (a) Kevin Chain description for
viscoelastic component; (b) microscopic creep compliance functions of the solidified matter;
(c) schematic representation of the solidification creep model.
57

Figure 2-20 shows the model for the role of solidification in creep of this theory. It is assumed

that the volume v of the hydrated cement grows by deposition of layers of solidified matter, as

shown in Figure 2-20c. ıg(v,t) is the stress at time t in the layer which solidifies when the total

volume of the solidified matter is v. At the moment it solidifies, the layer (dv) must be stress-

free, i.e. ıg[v(t),t] = 0. Therefore, the non-aging viscoelastic stress-strain relationship for the

layer that solidifies at time  is (Bazant 1977)

t
Hv t  Hv W ³ ) t  t ' V g >v W , dt '@ 2.24
W

in which ıg[v(),t’] = 0; v is the viscoelastic strain due to solidified matter and (t - t’) is the

microscopic creep compliance function of the solidified matter, representing the strain at age t

caused by a unit stress applied at . Function (t - t’) is assumed as an age-independent

function and only with one variable, namely the duration of time t - t’.

Further formulation of Equation 2.24 shows that ıg can be eliminated, and a macroscopic

stress-strain relation can be formed

J t 1 t 
Hv t ³ ) t  t ' dV t ' 2.25
vt vt 0

 t  t'
in which ) w) t  t ' / wt , (t) is regarded as the viscoelastic microstrain.

According to the analysis of test results, it is found that another component f, called flow,

must also be included in the concrete creep, as shown in Figure 2-20c. It is viscous rather than

viscoelastic. Similar to Equation 2.25, it can be described as

1 t 
H f (t ) ³ < t  t ' dV t ' 2.26
ht 0
58

< t  t' t  t ' /K 0 2.27

where 0 is called the effective viscosity of the solidified mater, as a constant; and h(t) is the

volume fraction of the solidified mater associated with the viscous strains.

Substituting Equation 2.27 into Equation 2.26 gives:

V t V t
H f t 2.28
K0h t Kt

in which (t) = 0 h(t) is the apparent (effective) macroscopic viscosity.

Constitutive relation for creep

According to the previous description, the total strain of concrete may be expressed as:

V
H cp  H cp  H sh , H cp Hv  H f 2.29
E0

in which /E0 is the ultimate elastic strain resulting from the deformation of the mineral

aggregate pieces in concrete and the microscopic elastic particles in the hardened cement

paste. Bažant and Prasannan suggested that the effective modulus E(t), which in ordinary

cases is an age-dependence value, be a constant asymptotic modulus. It represents a modulus

in a very short-time period, so that the age-dependent part of the ordinary elastic strain is

treated to be an apparent elastic deformation. Bazant and Bawja (Bazant and Baweja 1995a)

proposed that E0 = 1.6Ec28 where Ec28 is the elastic modulus of concrete at 28 days.

According to Equations 2.25, 2.28 and 2.29, the creep compliance function can be depicted as
59

1
t
§) (t  t ' ) 1 ·
J (t , t ' )  ³ ¨¨  ¸dt
K (t ) ¸¹
2.30
E 0 0 © v(t )

The empirical functions in Equations 2.25 and 2.28 were introduced by Bažant and Prasannan

(1986b) in the following forms

J t, t' q1  C t , t ' 2.31

C t, t ' >
q 2 Q t , t '  q3 ln 1  t  t '
n
@ q 4
§t·
ln¨ ¸
© t' ¹
2.32

m n 1
§ O0 · n W  t '
t
Q t, t ' ³t ' ¨© W ¸¹ O0  W  t ' n dW 2.33

in which q2, q3, q4, n, m, Ȝ0 are empirical constants. For all types of concrete, experimental

data indicated:

n 0.1 , m 0.5 , O0 1day 2.34

The binomial integral Q t ,t ' function in Equation 2.33 is approximated as

1 / r t '
ª § Q f t' ·
r t'
º
Q t, t' Q f «1  ¨¨ ¸¸ » 2.35
«¬ © Z t , t ' ¹ »¼

where:

8
0.12
r t' 1.7 t ' 2.36

Z t, t' t'
m
>
ln 1  t  t '
n
@ 2.37

log Q f t ' >


 0.112  0.4308 log t '0.0019 log t '
2
@ 2.38

The computational implementation of solidification creep theory will require transforming the
60

formulation into a rate-type law based on a rheological model with non-aging properties. The

means to obtain this formulation will be presented in Chapter 6 as this method will be used

for the finite element analysis.


CHAPTER 3 TIME-DEPENDENT
ANALYSIS OF TENSION STIFFENING

3.1 OVERVIEW
In this chapter, an analytical study is undertaken to describe the time-dependent behaviour of

reinforced concrete and the change of tension stiffening with time. The effects of creep and

shrinkage are taken into consideration, together with the deterioration of bond and the effects

of time-dependent cracking. The response of a uniaxial tension member will be analysed first

since tension stiffening in this case is more directly measurable. Thereafter the analysis of a

single-reinforced flexural member will be conducted.

3.2 TIME-DEPENDENT ANALYSIS OF A UNIAXIAL


TENSION MEMBER
Figure 3-1 shows the idealised instantaneous and time-dependent responses of a

concentrically reinforced concrete member subjected to axial tension (Figure 2-1), both before

and after cracking. As described earlier (section 2.2.1) the instantaneous response (curve OAB

in Figure 3-1) is linear up until first cracking at P = Pcr and non-linear after cracking. Before

cracking, the instantaneous tension stiffening strain tsi (which is the difference between the

strain in the specimen and strain in the bare bar) increases with load, but after cracking, tsi
62

decreases as P increases. The load at which cracking occurs depends on the tensile strength of

the concrete at the time of loading and also on the amount of drying shrinkage that has

occurred prior to loading. Early shrinkage is restrained by the embedded reinforcing bar and a

significant tensile stress may exist in the concrete before the external load P is applied to the

member.

Figure 3-1: Time-dependent response of an axially loaded tension member

Due to Creep (Without Shrinkage):

Figure 3-2 illustrates the effects of creep and shrinkage on an uncracked cross section of a

uniaxial tension member. If the load P is held constant with time, creep strain develops in the

concrete and the average axial strain changes. In fact, because the stress in the tensile concrete

is relatively small, creep strain is relatively small and the change of stress and strain in the

concrete is relatively small. In a member that does not shrink, tensile creep causes a softening
63

of the concrete in tension and a relatively small change in the load-deformation response with

time (as shown by the dashed curve labeled ‘after creep only’ in Figure 3-1). Tensile creep of

the concrete sheds some of the tensile force carried by the concrete into the bonded

reinforcement, thereby reducing tension stiffening with time and the ‘after creep’ response is

closer to the bare bar line than the instantaneous response.

Due to Shrinkage:

Shrinkage of the concrete while the member is under load causes a significant change in the

load-deformation response (which is shown in Figure 3-1 as curve O’A’D). For an uncracked

section, if unreinforced, shrinkage would cause a free shortening by an amount of shx

(Figure 3-2b). While reinforcement is provided, it restrains the free shrinkage of concrete. The

reinforcement would be subjected to a compressive force EsAss while an opposite tensile

force would be applied to the concrete (denoted as T in Figure 3-2b). Due to this restraint to

shrinkage, the shortening of the section reduces to sx. Therefore, before cracking, shrinkage

causes the member to shorten and, at low tensile loads, the load-deformation curve moves to

the left in Figure 3-1, as shown by the dashed curve O’A’. Restraint to shrinkage also causes a

gradual build-up of tensile stress in the concrete and this reduces the cracking load from Pcr to

Pcr.sh.

After first cracking (P > Pcr), cracking and the deterioration of bond caused by cracking

resulting in slip at the concrete-steel interface are the physical mechanisms that are primarily

responsible for the reduction in tension stiffening due to restraint to shrinkage. Without

cracking, incremental increases in stress caused by the restraint to shrinkage would cause an

increase in tension stiffening, but the formation of new cracks and the resulting bond slip

results in a subsequent overall (net) drop in tension stiffening. Therefore, at load P (> Pcr) in
64

Figure 3-1, the instantaneous tension stiffening strain tsi is represented by the horizontal

distance BE and this reduces to CE due to creep and further reduces to DE after both creep

and shrinkage. It is often assumed that tension stiffening reduces with time under sustained

loads to about 50% of its instantaneous value, but this is yet to be conclusively demonstrated.

a) Due to creep

b) Due to shrinkage
Figure 3-2: Time-dependent deformation and stress on an uncracked section of the uniaxial
tension member.

3.3 TIME-DEPENDENT ANALYSIS OF A


FLEXURAL MEMBER
The instantaneous response of a reinforced concrete beam subjected to gradually increasing

load was discussed in Section 2.2.2. The time-dependent response is described here.

The fundamental difference between a uniaxial tension member and a flexural member is the

strain gradient and the load carrying mechanism. In the uniaxial tension member, the stress
65

and strain distribution on the cracked section would not change with time, for concrete carries

no stress throughout. This is not the case for the cracked section of a flexural member, where

part of the concrete section is in compression. In a cracked reinforced concrete beam, the

time-dependent response of a particular cross-section depends on its proximity to the nearest

primary crack. In the following, the time-dependent deformations on an uncracked section

and on a cracked section are described separately. The analyses described in the following

sections are based on the assumptions that plane sections before loading remain plane after

loading. Therefore both the short-term and long-term strain distributions on a section are

assumed to remain linear.

3.3.1 Time dependent response of an uncracked singly-reinforced section

Due to Creep (Without Shrinkage):

Consider the uncracked singly reinforced rectangular section shown in Figure 3-3a subjected

to a relatively small sustained moment Ms , so that the maximum tensile stress in the concrete

is always less than the tensile strength of the concrete (i.e. the section is uncracked). In this

case, the time-dependent change in strain on the cross-section is caused by creep. In the top

fibres of the concrete cross-section, compressive creep causes a gradual increase in the

compressive strain with time. In the bottom fibres, tensile creep causes the increase in the

tensile strains. The increase in tensile strain is restrained by the tensile reinforcing bars. Strain

compatibility requires that the tensile steel strain increases as the tensile concrete strain at the

steel level increases causing an increase in the tensile steel stress and an increase in the force

in the steel. Equilibrium requires a redistribution of concrete stresses, involving a reduction in

the concrete tensile stress and a lowering of the neutral axis (see Figure 3-3a).
66

a) Due to creep

b) Due to shrinkage

Figure 3-3: Time-dependent deformation on an uncracked beam section.

Due to Shrinkage:

The effects of concrete shrinkage on an uncracked and unloaded singly reinforced concrete

cross-section are shown in Figure 3-3b. Shrinkage of the concrete compresses the

reinforcement, which in turn imparts an equal and opposite tensile force on the concrete at the

steel level. The magnitude of the restraining force depends on the quality of bond and the

amount of slip. For uncracked members, the bond may be reasonably assumed to be good

with very little slip. Due to the eccentricity of the tensile restraining force T, a curvature is

induced on the cross-section and a gradual warping of the beam occurs, as shown in Figure 3-

3b. The magnitude of T is also dependent on the area of tensile reinforcement and the

geometry of the cross-section. It also depends on whether or not the concrete has cracked, i.e.
67

on the magnitude of the applied moment. Therefore, although shrinkage is independent of the

applied load, shrinkage warping is not. Furthermore, if the shrinkage-induced restraining

tensile force is great enough, cracking may be induced with time in a previously uncracked

member.

3.3.2 Time –dependent response of a cracked singly-reinforced section

Due to Creep (Without Shrinkage):

On a fully-cracked section, the tensile concrete loses its ability to carry tensile stress and all

the tensile force is carried by the reinforcing bars. Thus, under sustained bending moment,

only the concrete in the compressive zone undergoes significant creep. As creep strain

develops in the compressive zone, the top fibre strain increases gradually, the neutral axis

moves downwards and the compression zone becomes deeper. Aging of the concrete causes

the creep strain distribution at the section to become non-linear and, as a result of strain

compatibility, the instantaneous (or stress related elastic) strain also become nonlinear during

the period of sustained loading, as does the compressive stress distribution. The stress at top

fibre will reduces marginally with time and this affects the final creep strain in the top fibre.

As a result, under constant sustained moment, the tensile steel strain has to increase by a small

amount to compensate for the reduced lever arm between the resultant compressive and

tensile forces on the cross-section and thereby to maintain equilibrium.


68

a) Due to creep

b) Due to shrinkage

Figure 3-4: Time-dependent deformation on a cracked beam section.

Due to Shrinkage:

After cracking, when subjected to the eccentric shrinkage induced restraining force imposed

by the reinforcement, the relatively small area of intact concrete above the crack must resist

the shrinkage induced restraining force imposed by the reinforcement T. Equilibrium and

compatibility considerations dictate that the magnitude of T on a cracked section is very

small (as indicated in Figure 3-4b). The shrinkage induced curvature on the cracked section is

clearly significantly greater than that on an uncracked section, as can be seen by comparing

the deformations in Figure. 3-3b and 3-4b.


69

3.3.3 Time-dependent tension stiffening in the cracked reinforced


concrete beam

A cracked reinforced concrete beam contains both cracked and uncracked sections and the

time-dependent response of an ‘average’ section in the beam is intermediate between the

response of the uncracked and cracked sections. The deformation of the beam increases with

time, with the rate of change being greatly dependent of the magnitude of shrinkage and creep

during the time period under consideration and the extent and severity of cracking.

Figure 3-5: Time-dependent moment verse deflection response of a short-term beam.

Under short-term loading:

If a beam begins to shrink prior to loading, as is commonly the case, shrinkage warping

occurs and the member deflects by an amount Gsh0,, when the applied moment is still zero (i.e.

M = 0), shown as point O’ in Figure 3-5. Therefore, the experimentally measured response of
70

a member that has already begun shrinking must start from a positive initial deflection, as is

shown in Figure 3-5. Because of the initial tensile stress induced by early shrinkage in the

tensile concrete, the moment required to cause first cracking Mcr.sh will be less than Mcr (as

indicated in Figure 3-5) and the moment-deflection response is given by the curve O’A’B’C’.

The initial deflection of the fully-cracked member due to early shrinkage (Gsh0.cr) is

significantly larger than that of the uncracked member (Gsh0), so early shrinkage causes the

line representing the fully-cracked cross-sectional response to move significantly to the right

(shown as line O’’D’ in Figure 3-5). At any value of applied moment greater than the cracking

moment (such as Ms in Figure 3-4), the instantaneous tension stiffening deflection Gtsi is the

distance BE if no shrinkage has occurred prior to loading and is the distance B’E’ Gtsi(t) if

shrinkage has occurred prior to first loading. It is likely that the magnitude of the tension

stiffening deflection is dependent on the amount of shrinkage that occurs prior to first loading,

as of course is the magnitude of the cracking moment, but this is yet to be conclusively

demonstrated.

Under long-term loading:

For a simply-supported beam subjected to constant sustained loads, the bending moment Ms at

a particular cross-section is sustained over a time period t. If shrinkage is ignored, the

instantaneous mid-span moment versus deflection response of the member is shown as curve

OABC in Figure 3-6 (identical to curve OABC in Figure 3-5). The instantaneous fully-cracked

response (ignoring the tensile concrete) is also shown as line OD in Figure 3-6. If the member

does not shrink with time (i.e. Hsh remains at zero), creep causes an increase in deflection with

time and the member response shifts to curve OA’B’C’ in Figure 3-5a. The increase in

deflection with time is due to creep and may be expressed as 'G(t) = Gi M /D’, where Gi is the
71

instantaneous deflection, M is the creep coefficient and D’ is a factor that depends on the

amount of cracking and the reinforcement quantity and location. For reinforcement ratios

typical of beams, D’ is in the range 1.0 – 1.4 prior to cracking and in the range 4 - 6 when

cracking is extensive. With regard to the fully-cracked response (calculated assuming fully-

cracked cross-sections with no contribution from the tensile concrete), creep causes a

softening of the response shown as line OD’ in Figure 3-6a, with the slope of the line

increasing by the factor (1+ M /D’), where D’ is in the range 4 – 6.

When shrinkage after first loading is included, the deflection increases even further with time

due to shrinkage warping and the time-dependent response of the beam is shown as curve

O’A’B’C’ in Figure 3-6b. At zero loads (M = 0), the deflection increases due to shrinkage

warping of the uncracked member and the point O moves horizontally to O’. Due to the

restraint to shrinkage provided by the bonded reinforcement, tensile stress is induced with

time and this has the effect of lowering the cracking moment from Mcr to Mcr sh, as shown in

Figure 3-6b. For any flexural members subjected to a sustained moment in the range Mcr.sh <

Ms d Mcr, cracking will occur with time and the increase in deflection will be exacerbated by

the loss of stiffness caused by time-dependent cracking. In practice, many slabs are loaded in

this range. The fully-cracked response after creep and shrinkage is shown as line O’’D’ in

Figure 3-6b. The shrinkage induced deflection of the fully cracked member (ignoring the

tensile concrete) at zero loading (M = 0) is greater than that of the uncracked member and the

fully-cracked response is shifted horizontally from point O to point O’’ as shown. The slope

of fully-cracked response in Figure 3-6b is softened by creep and the slope of the line O’’D’

in Figure 3-6b is the same as the slope of line OD’ in Figure 3-6a.

With time, deflection increases from B to B’ while the fully-cracked response is also moved
72

from E to E’. The time-dependent tension stiffening deflection for the long-term tests

becomes from BE (tsi) to B’E’ (ts(t)). The difference between tsi and ts(t) in the figure

indicates the decay of tension stiffening with time.

An alternative approach to analyse the singly reinforced section in sustained bending for

cracked and uncracked section is given by Gilbert (1988). This method will be used to analyse

the experimental results, to obtain the fully-cracked response of a flexural member. The

details of this method and a case study are presented in the Appendix B.

3.4 SUMMARY
This chapter provides qualitative description of the time-dependent responses of uniaxial

tension members and flexural reinforced concrete members. It is noted that the loss of tension

stiffening with time is associated with concrete cracking, either primary cracks or the

formation of additional cover-controlled cracks that cause local bond damage between

concrete and steel. Shrinkage is deemed to be a significant factor since the restraint to

shrinkage from the reinforcing bar can produce additional tensile stress in the concrete which

may indirectly induce cracking. Tensile creep plays a minor role in the decay of tension

stiffening with time as it gradually releases the tensile stress in the concrete.

The discussion concludes that the way to predict time-dependent tension stiffening in

reinforced concrete structures consideration must be given to the effects of creep and

shrinkage, as well as the effects of the cracking of concrete. An alternative approach to predict

the time-dependent responses of the cracked uniaxial tension member and flexural member

are proposed by the author. The approach is similar to that outlined in Eurocode 2 (1992), but

modified to include the effects of time-dependent slip at the concrete-steel interface caused by
73

shrinkage and by time-dependent cracking. The proposed approach will be presented in

Appendix A of this thesis.

a) After creep only (no shrinkage)

b) After both creep and shrinkage

Figure 3-6: Time-dependent moment verse deflection response of a long-term beam.


CHAPTER 4 EXPERIMENTAL PROGRAM
AND RESULTS OF UNIAXIAL TENSION
TESTS

4.1 OVERVIEW
The experimental program involved the testing of reinforced concrete prisms in axial tension

and flexural members (beams and slabs) in bending. The tension prism tests were undertaken

before the flexural tests, to first measure the effects of tension stiffening in reinforced

concrete in direct tension, before considering the imposition of the strain gradient associated

with bending. This chapter describes the testing process and presents the results of the

uniaxial tension tests. Chapter 5 discusses the flexural specimens and testing program.

The major objectives of the experimental program were as follows:

ƒ To quantify tension stiffening in reinforced concrete members under increasing load and

to measure the effects on tension stiffening of concrete shrinkage prior to cracking;

ƒ To investigate the change in tension stiffening with time in a cracked member under

sustained service loads and, if significant, to measure the effect of creep and shrinkage on

the magnitude of tension stiffening; and

ƒ To investigate the effects of the time-dependent change in tension stiffening on the crack
75

width and crack spacing.

The primary test variables in the experiments were the load history, the reinforcement quantity,

and the creep and shrinkage characteristics of the concrete, which were measured throughout

the tests, from the commencement of drying until the end of the test. In addition, the

externally applied forces and reactions and the time-dependent displacements of the specimen

were recorded electronically throughout the test period, including the elongation for the

uniaxial tension members, the vertical deflection for the flexural members, the local steel

strains between the primary cracks, the crack width and the crack spacing.

4.2 SPECIMEN CONFIGURATION


A total of eight concrete prism specimens were tested. All specimens had the same overall

dimensions, with details given in Figure 4-1. Each concrete prism was of square cross-section

(100mm by 100mm) and 1100 mm long and contains a single ribbed reinforcing bar running

longitudinally through the centroid of the cross-section. The tensile force P was applied to the

ends of the reinforcing bar protruding from each end of the concrete prism. Part of the tensile

force is transferred into the concrete from the reinforcement bar through bond at the steel-

concrete interface. Over the middle 600 mm length of each specimen, 25 strain gauges were

attached to the reinforcing bar at 25mm centres in order to monitor the local steel strains.

Typical formwork boxes for the uniaxial tension specimens, prior to the placement of concrete,

are shown in Figure 4-2, where strain gauges were attached to the reinforcing bar without

removing the ribs so as to minimise the effect of the strain gauge on bond.

Four of the specimens (STN12, STN16, STS12 and STS16) were tested under short-term

monotonically increasing load, with loading continuing into the post-yield range up to an
76

average steel strain of 0.3%. These four tests are referred to as the short-term tests. The other

four specimens (LTN12A, LTN12B, LTN12C and LTN12D) were subjected to a constant

sustained service load for a period of about 50 days and are referred to as the long-term tests.

The first two letters in the designation of each specimen indicate the test duration; “ST” for

short-term and “LT” for long-term. The third letter indicates whether or not the specimen

commenced drying and began to shrink prior to the application of load; “S” if yes and “N” if

no. The next two digits indicate the reinforcing bar diameter, either 12 mm or 16 mm. For the

four long-term tests, the final letter A, B, C and D distinguishes the different load cases.

To investigate the influence of reinforcing ratio on tension stiffening, two bar sizes were used,

namely 12mm and 16mm diameter bars. Therefore, the reinforcement ratio for each specimen

was either 1.11% or 2.04%.

Figure 4-1: Specimen details of uniaxial tension members.


77

Figure 4-2: Strain gauge detailing of uniaxial tension members.

4.3 TEST INSTRUMENTATION


The loading system for the short-term uniaxial tension specimens (STN12, STN16, STS12

and STS16) is shown in Figure 4-3. The tests were undertaken in an Instron universal testing

machine. The specimens were positioned vertically within the testing machine as shown. The

member elongation was measured by attaching a pair of LVDTs (linear variable displacement

transducer) to a mounting frame that was clamped firmly to the specimen at both the front and

back surfaces. The gauge length of the LVDTs over which the elongation of the specimen was

measured was 900mm. The tests were undertaken in displacement control, with the rate of

application of axial displacement controlled and set to 0.05mm/min. The corresponding

tensile force was recorded throughout via the pressure transducer connected to the machine

grips. Loads, elongation and steel strains (strain gauges) were all recorded electronically using

an HBM amplifier and stored in the computer hardware by a data logger.


78

Figure 4-3: Set up for short-term uniaxial tension tests.

For the long-term tests, the rig shown in Figure 4-4 was designed and constructed. The

specimen was positioned vertically by fixing the reinforcing bar protruding from the

specimen through holes in the cross-heads both top and bottom. A constant sustained tensile

force was imposed on each specimen through an adjustable anchor-support at the top of the

rig and monitored using a load cell connected to the reinforcing bar at the base of the rig.

Five Demec targets were attached on opposite sides of each tension member at 250 mm

centres, so that the averaged elongation of the specimen at specific loads could be obtained.

The strain gauges attached to the reinforcing bar were connected to the HBM amplifier and

the readings were taken manually at regular time intervals throughout the period of sustained

loading. Crack widths were also measured throughout the test using a microscope with a
79

magnification factor of 40.

Figure 4-4: Set up for long-term uniaxial tension tests.

4.4 CURING AND TESING PROCEDURE


All the short-term specimens were cast from the same batch of ready-mixed concrete and

cured under wet burlap. The burlap was kept wet for 28 days to facilitate strength gain of the

concrete and to delay the commencement of drying shrinkage. The Demec gauge targets were

glued onto the concrete surface after removing the burlap. For the short-term tests, specimens

STN12 and STN16 were tested immediately after wet curing, so that relatively little shrinkage

had occurred at the time of testing. Specimens STS12 and STS16 were uncovered and

permitted to dry for a period of four weeks before testing.

For long-term test specimens, LTN12A and LTN12B were cast from one batch of concrete

(the same as the short-term specimens), while LTN12C and LTN12D were cast from a
80

different batch. All specimens were cured under wet burlap until the commencement of the

test. After curing, the specimens were put into the testing rig and the load was imposed by

manually adjusting the anchorage supports. LTN12A and LTN12C was loaded up to 40kN (all

cracks stablised), while LTN12B and LTN12D were subjected to 20kN (which was close to

the cracking load), and then the loads were maintained constant for 50 days. Table 4-1 shows

the testing regime for the uniaxial tension members.

The basic concrete properties were measured on companion specimens at the onset of each

test. The compressive strength and the elastic modulus of concrete were measured on standard

100 mm diameter concrete cylinders. The indirect concrete tensile strength was measured on

standard cylinders using the Brazil tests and the flexural tensile strength was measured on

100mm by 100mm by 600mm concrete prisms.

Table 4-1: Testing procedure of the uniaxial tension test specimens.

Specimens Day 1-28 Day 32 Day 57


STN12
Wet cured Tested -
STN16
under wet
STS12
burlap Start drying Tested
STS16
(a) Short-term uniaxial tension specimens

Specimens Day 1-28 Day 33 or 28 Day 32 - 83 Day 83


Loaded to
LTN12A
40kN
Loaded to
LTN12B Wet cured Load
20kN
under the Loaded to sustained for End test
LTN12C burlaps 50 days
40kN
Loaded to
LTN12D
20kN
(b) Long-term uniaxial tension specimens
81

The creep and shrinkage characteristics of the concrete were also measured on companion

specimens throughout the test. Unrestrained concrete prisms, each 600mm long with

100×100mm cross section, were used to measure the development of shrinkage strain with

time (Figure 4-5b). Brass plugs at 250 mm centres were embedded in opposite surfaces of the

prism during casting, and a Demec gauge measure was used to record the average shrinkage

strain between the two plugs in the prism. The deformation between the brass plugs was

measured as soon as the wet burlap was removed.


82

a) Creep specimens

b) Shrinkage specimens

Figure 4-5: Creep and shrinkage companion tests.

The standard creep rigs were set up as shown in Figure 4-5a. Two standard 150mm diameter

cylinders were loaded in the rig with a maximum 6MPa compressive stress applied by the

hydraulic jack at the bottom of the creep rig. The stress was then kept constant. Two other
83

standard cylinders were left unloaded and drying besides the rigs. The creep strain at any time

can therefore be calculated by subtracting the shrinkage strain of the unloaded cylinders and

the instantaneous strain of the loaded cylinders from the total strain measured on the loaded

cylinders at that time. Creep tests were initiated at the commencement of drying and at the

age of first loading for the long-term tests.

The stress-strain curves of the deformed reinforcing bars were obtained in the Instron

universal testing machine. The tensile force and the position of the upper rod were recorded.

Therefore stress can be calculated by dividing the tensile force by the bar area, and the

average strain can also be derived by dividing the measured elongation by the gauge length.

4.5 MATERIAL PROPERTIES


4.5.1 Steel bar properties

There are basically two different sizes of deformed 500 Grade reinforcing bars used in the

experimental program, namely N12 and N16 bars. Three N12 and three N16 bars that were

from the same provider were tested in the Instron machine. The measured yield stress and

elastic modulus were then averaged and the results are given in Table 4-2.

Table 4-2: Reinforcing bar properties.

Bar Db (mm) Ast (mm2) fsy (MPa) Es (MPa)


N16 16 201 500 204000
N12 12 113 500 200000

4.5.2 Instantaneous properties of concrete

The concrete used in this study was supplied by a pre-mixed concrete producer (BORAL

Limited) and the specification of the concrete mix is:


84

x Compression strength = 25MPa

x Maximum aggregate size = 20mm

x Slump = 80mm

x Water to cement ratio = 0.6

The compressive strength fcp, the indirect tensile strength fct, the flexural tensile strength fcf

and the elastic modulus of concrete Ec are presented in Table 4-3. For the short-term

specimens, these properties were measured at the same days of the tests; while for the long-

term specimens, they were measured at both the beginning and the end of the sustained

loading time. There were two identical companion specimens were tested to measure these

instantaneous properties at the time of each test. In general, the concrete strengths increase

with time by less than 16% during the sustained loading period.

4.5.3 Creep and shrinkage

Figure 4-6 shows the development of the creep coefficient c with time for the plain concrete

elements measured in the companion tests. The creep coefficient is the ratio of the creep

strain to instantaneous strain soon after first loading at different concrete ages. There were

two identical creep tests conducted for each batch of concrete as shown. The results are then

averaged and plotted in the diagrams.


85

Table 4-3: Instantaneous concrete properties for uniaxial tension tests.

Specimen f cp f ct f cf Ec
STN12
21.56 2.04 3.05 22400
STN16
STS12
24.73 2.15 3.08 21600
STS16
(a) Short-term uniaxial tension specimens

f cp f ct f cf Ec
Specimen
Start End Start End Start End Start End
LTN12A
21.56 24.88 2.04 2.44 3.05 3.54 22400 23600
LTN12B
LTN12C
27.73 29.02 2.85 3.02 4.00 4.40 20500 22300
LTN12D
(b) Long-term uniaxial tension specimens

* All units in MPa.

c 1.6 1.6
c

1.2 1.2

0.8 0.8
creep test 1
creep test 1 creep test 2
0.4 creep test 2 0.4 Averaged
Averaged

0.0 0.0
20 40 60 80 20 40 60 80
Concrete age (days) Concrete age (days)

(a) Batch 1 (STN12, STN16, STS12, STS16, (b) Batch 2 (LTN12C and LTN12D)
LTN12A and LTN12B)
Figure 4-6: Creep coefficients for the uniaxial tension tests.
86

For the short-term tests, the averaged creep coefficients at the onset of tests are 0.17 (STN12

and STN16) and 1.13 (STS12 and STS16) respectively. For the long-term tests, the averaged

creep coefficients by the end of the sustained loading periods are 1.38 (LTN12A and LTN12B)

and 1.11 (LTN12C and LTN12D) respectively.

Figure 4-7 shows the development of shrinkage strains after drying commenced. There were

also two companion shrinkage tests conducted for each batch of concrete. The measured

shrinkage strains are then averaged and plotted in the figures. For the short-term tests, the

averaged shrinkage strain at the onset of tests are -28 x 10-6 (STN12 and STN16) and -249 x

10-6 (STS12 and STS16) respectively. For the long-term tests, the averaged shrinkage strain

by the end of the sustained loading periods are -310 x 10-6 (LTN12A and LTN12B) and -356 x

10-6 (LTN12C and LTN12D) respectively.


Shrinkage strain (×10- 6)

400 400
Shrinkage strain ( ×10- 6

300 300

200 200
shrinkage test 1
100 100 shrinkage test 2
Averaged
0 0
0 20 40 60 80 0 20 40 60 80
Concrete age (days) Concrete age (day

(a) Batch 1 (b) Batch 2

Figure 4-7: shrinkage strain development for uniaxial tension tests.


87

4.6 TESTING RESULTS OF THE UNIAXIAL


TENSION PRISMS
4.6.1 Short-term test results

4.6.1.1 STN12 and STN16

Axial Load versus Average strain:

Specimen STN12 and STN16 were subjected to monotonically increasing deformation soon

after the end of moist curing. The elastic modulus of concrete at the time of testing at age 32

days was Ec = 22400 MPa (and with Es = 200000 MPa, n = 8.93) and the tensile strength of

concrete was f ct = 2.04 MPa. Prior to loading, the measured drying shrinkage strain in the

unreinforced companion member was small at Hsh0 = -28 x 10-6 and the corresponding creep

coefficient was c = 0.17. The calculated strains in the specimen (accounting for the restraint

provided by the reinforcement – see Appendix A Equation A.1) were -25 x 10-6 for STN12 and

-23 x 10-6 for STN16 respectively.

The measured axial load versus strain relationships for STN12 and STN16 are given in Figure

4-8 and 4-9 respectively, together with the bare bar response. Initially, prior to first cracking

in portion O’B of the curves in the figures, the specimens are at their stiffest and the load-

strain curves are steep. At first cracking (Point B when P = Pcr = 21.1kN and 23.0kN), there is

an abrupt change of stiffness and the stiffness continues to degrade under increasing

deformation as further cracks occur (portion BC). As the load increases, the tension stiffening

strain gradually reduces. In total, 5 primary cracks occurred in each specimen as loading

progressed, as indicated by the numbered peaks in the curves in Figure 4-8 and 4-9.

Tables 4-4 and 4-5 provide values of average axial strain Hs.avg and tension stiffening strain Htsi
88

at selected values of total applied load P, together with the average force carried by the

concrete and the steel, Pconc and Psteel, respectively, within the prisms. Psteel is derived by

multiplying the measured average axial strain with the elastic modulus and the cross-section

area of the steel bar. Subtracting Psteel from the applied load P gives Pconc.

It appears that with increasing load P, Htsi or Pconc drops off rapidly, whereas Psteel keeps

increasing. The decrease in Htsi or Pconc does not only occur at the crack formation stages, but

also after the cracks have stabilised. For example, Pconc of STN12 decreases from 4.4kN to

1.8kN as the load increases from 45kN to 55kN. Note that all the cracks had formed when P

reached 40kN. For STN16, Pconc decreases from 7.1kN to 5.0kN as the load increases from

65kN to 85kN after all the cracks had formed. The mechanism to induce this drop-off in

tension stiffening is likely to be a bond breakdown due to cover-controlled cracking.

70

60
Axial Load (kN)

50 Experimental C
Hts 5
40 Bare bar

30 4 Crack numbers
Pcr B1 2 3
20 = 21.1 kN

10 -6
Average Axial Strain (x10 )
0 O’
-500 0 500 1000 1500 2000 2500 3000
-25

Figure 4-8: Average load P versus strain (Hs.avg) for STN12.


89

120

100 Experimental
C
Axial load (kN)

Bare bar
80

60
 crack number
40 B

P cr =  tsi
20 -6
23.0kN Average Axial Strain (×10 )
O'
0
-500  0 500 1000 1500 2000 2500 3000

Figure 4-9: Average load P versus strain (Hs.avg) for STN16.

Table 4-4: Measured loads and strains for STN12.


Applied load, P (kN) 0 10 21.1 25 35 45 55
Average strain, Hs.avg (x 10-6) -25 9.2 55.5 561 1148 1795 2355
Tension stiffening strain Htsi (x 10 ) -6
25 433 878 545 401 196 79
Avge force in concrete Pconc (kN) 0.57 9.79 19.8 12.3 9.1 4.4 1.8
Avge force in steel bar Psteel (kN) -0.57 0.21 1.3 12.7 25.9 40.6 53.2

Table 4-5: Measured loads and strains for STN16.


Applied load, P (kN) 0 10 23.0 30 45 65 85
Average strain, Hs.avg (x 10-6) -23 8.2 57 386 866 1440 1990
Tension stiffening strain Hts (x 10 ) -6
23 241 523 360 253 177 124
Avge force in concrete Pconc (kN) +0.92 9.67 20.7 14.5 10.2 7.1 5.0
Avge force in steel bar Psteel (kN) -0.92 0.33 2.3 15.5 34.8 57.9 80.0

Variation of steel strains:

The variation of steel strains over the 600mm gauge length of the reinforcing bar were

measured as the test progressed using strain gauges and the variation in force in the

reinforcing steel was determined from the measured strains. The variation in concrete force

was obtained by subtracting the force in the steel at each strain gauge location from the total
90

force applied to the specimen.

Figure 4-10 and 4-11 show the variation in the tensile forces carried by steel and concrete at

different loading stages for the short-term specimen STN12 and STN16. As can be seen, the

formation of a primary crack causes a large local decrease of force in the concrete and an

increase of force in the steel bar. All the tensile force is carried by the steel at the cracked

section. At a certain distance S from the crack, the tensile force carried by the concrete is

unaffected by the crack as shown in Figure 4-10b and Figure 4-11b. Of course, at cracking,

the sudden loss of stiffness causes a drop-off in the applied load and the force carried by the

concrete reduces accordingly as can be seen by comparison of Figure 4-10a and b and Figure

4.11a and b.

For example, just before first cracking of STN12 at P = 21.1kN (Figure 4-10a), the tensile

force in the concrete is almost uniform and equal to about 19kN, while the force in the steel is

about 2.1kN. Soon after first cracking (Figure 4-10b), the applied load drops to 17.8kN. At the

cracked section, steel carries the entire load and the force in the concrete drops to zero. At

distances greater than S | 140mm from the first crack, the concrete force is again uniform and

equal to about 16kN. As the load increases to 22.4kN (Figure 4-10c), the maximum tensile

force in the concrete exceeds 20kN, which allows the concrete tensile stress to exceed the

tensile strength of the concrete at the second weakest cross-section, thus a second primary

crack occurs (Figure 4-10d).


91

25 25
S=140mm
20 20
Force (kN)

Force (kN)
15 15

10
Steel 10
Steel
Concrete Concrete
5 5

0 0
200 300 400 500 600 700 800 900 200 300 400 500 600 700 800 900
Distance along prism, mm Distance along prism, mm
(a) Just before first cracking, P = 21.1 kN. (b) Just after first cracking, P = 17.8 kN.
25 25
Steel
20 20 Concrete
Force (kN)

Force (kN)
15 15

Steel
10 10
Concrete

5 5

0 0
200 300 400 500 600 700 800 900 200 300 400 500 600 700 800 900
Distance along prism, mm Distance along prism, mm
(c) Just before 2nd crack, P = 22.4 kN. (d) Just after 2nd crack, P = 17.4 kN.
30 30
Steel Steel
25 25
Concrete Concrete
Force (kN)

20
Force (kN)

20

15 15

10 10

5 5

0 0
200 300 400 500 600 700 800 900 200 300 400 500 600 700 800 900
Distance along prism, mm Distance along prism, mm
(e) Just before 3rd crack, P = 21.8 kN. (f) Just after 3rd crack, P = 19.6 kN.
60
50

50
40

40
Force (kN)
Force (kN)

30
30
Steel Steel
20 Concrete
Concrete 20

10
10

0 0
200 300 400 500 600 700 800 900 200 300 400 500 600 700 800 900
Distance along prism, mm Distance along prism, mm
(g) At P = 40.0 kN. (h) At P = 50.0 kN.

Figure 4-10: Variation of forces in steel and concrete at different loading stages (STN12).
92

25
25
Steel S=150mm
20 Concrete
20

Force (kN)
Force (kN)

15 15
Steel
10 10
Concrete

5 5

0 0
200 300 400 500 600 700 800 900 200 300 400 500 600 700 800 900
Distance along prism, mm Distance along prism, mm
(a) Just before first cracking, P = 23.0 kN. (b) Just after first cracking, P = 20.8 kN.
30 25
Steel Steel
25 Concrete 20 Concrete
Force (kN)

20
Force (kN)
15
15
10
10

5 5

0 0
200 300 400 500 600 700 800 900 200 300 400 500 600 700 800 900
Distance along prism, mm Distance along prism, mm
(c) Just before 2nd crack, P = 24.3 kN. (d) Just after 2nd crack, P = 22.7 kN.
30
30 Steel
Steel
25 Concrete
25 Concrete
Force (kN)

20
Force (kN)

20

15 15

10 10

5 5

0 0
200 300 400 500 600 700 800 900 200 300 400 500 600 700 800 900
Distance along prism, mm Distance along prism, mm
(e) Just before 3rd crack, P = 24.9 kN. (f) Just after 3rd crack, P = 22.8 kN.
60 100
Steel 90
50 Concrete 80
Force (kN)

70
Force (kN)

40
60
30
50

20 40 Steel
30 Concrete
10
20

0 10
200 300 400 500 600 700 800 900 0
Distance along prism, mm 200 300 400 500 600 700 800 900
Distance along prism, mm
(g) At P = 40.0 kN.
(h) At P = 75.0 kN.
Figure 4-11: Variation of forces in steel and concrete at different loading stages (STN16).
93

Thereafter, the development of new primary cracks follows a similar pattern (Figure 4-10e

and f) until all the cracks have formed at P = 40kN (Figure 4-10g). It is found that there is no

dramatic change of local steel forces or concrete forces when 40kN < P < 50kN (Figure 4-

10h), just a gradual reduction in average concrete force with increasing load.

Similar to STN12, STN16 also undergoes the same pattern for the development of local forces

in the concrete and steel, as shown in Figure 4-11.

Crack spacing and crack width:

The crack locations and crack numbers for STN12 and STN16 are shown in Figure 4-12 and

4-13. The crack numbers indicate the order of the formation of each crack, corresponding to

the numbers shown in Figures 4-8 and 4-9.

5 2 3 1 4

150 95 320 175 190 170

Figure 4-12: Crack numbers and locations (STN12).

5 3 4 1 2

230 115 180 125 240 210

Figure 4-13: Crack numbers and locations (STN16).

The maximum crack spacings for STN12 and STN16 are 320mm and 240mm respectively,

while the average crack spacings are 195mm and 165mm. STN12 (with the smaller

reinforcing ratio) exhibits slightly greater crack spacings than STN16. Using Equation 2.7 of

the Tension Chord model, the maximum crack spacings for STN12 and STN16 are calculated
94

as 263mm and 192mm. It appears that the measured maximum crack spacing is greater than

the one predicted by the Tension Chord model. It is often assumed that the average spacing

can be taken as the average of maximum crack spacing Sr.max and minimum crack spacing

Sr.max/2 (Equation 2.8). Therefore the calculated average crack spacings according to the

Tension Chord model are 197.3mm and 144.2mm, which are close to the measured average

values.

The crack widths at various stages of loading are given in Tables 4-6 and 4-7, together with

the steel stress (Vs) at the crack at each loading stage. Generally crack width increases as the

steel stress increases. The data shows that after cracks have stabilised, the maximum crack

width is about 1.5 – 2.0 of the average crack width. For example, at, Vs  440MPa, the

average crack widths for STN12 and STN16 are 0.200mm and 0.245mm, while the maximum

crack widths are 0.375mm and 0.375mm respectively (1.875 and 1.53 times of the average

values). In addition, the maximum crack width seems not to be affected by the bar diameter.

At similar steel stresses after cracks have stabilised, the measured maximum crack widths are

similar for STN12 and STN16.

Table 4-6: Crack widths at different loading stage (STN12).

Load Vs Width of Width of Width of Width of Width of Average Maximum


1st crack 2nd crack 3rd crack 4th crack 5th crack crack width crack width
Stage (MPa)
(mm) (mm) (mm) (mm) (mm) (mm) (mm)
After 1st 158 0.075 - - - - 0.075 0.075
After 2nd 154 0.075 0.05 - - - 0.063 0.075
After 3rd 173 0.10 0.125 0.05 - - 0.092 0.125
After 4th 212 0.10 0.25 0.10 0.05 - 0.125 0.25
After 5th 338 0.10 0.30 0.20 0.125 0.025 0.150 0.30
40kN 354 0.15 0.30 0.25 0.125 0.1 0.185 0.30
50kN 442 0.125 0.375 0.25 0.15 0.1 0.200 0.375
95

Table 4-7: crack widths at different loading stage (STN16)

Maximum
Width of Width of Width of Width of Width of Average
Vs
Load crack
1st crack 2nd crack 3rd crack 4th crack 5th crack crack width
Stage (MPa) width
(mm) (mm) (mm) (mm) (mm) (mm)
(mm)
After 1st 104 0.013 - - - - 0.013 0.013
After 2nd 113 0.025 0.05 - - - 0.0375 0.05
After 3rd 114 0.05 0.075 0.05 - - 0.0583 0.075
After 4th 196 0.10 0.25 0.125 0.05 - 0.1313 0.25
After 5th 299 0.125 0.325 0.10 0.225 0.025 0.160 0.325
75 kN 373 0.15 0.325 0.225 0.15 0.125 0.215 0.325
90 kN 447 0.20 0.375 0.25 0.25 0.15 0.245 0.375
105 kN 522 0.25 0.50 0.25 0.275 0.225 0.3 0.50

4.6.1.2 STS12 and STS16

Specimen STS12 and STS16 were identical to specimen STN12 and STN16 respectively,

except that they were tested 3½ weeks later at age 57 days when the drying shrinkage had

increased to Hsh = -249 x 10-6 in each specimen At the time of testing at age 57 days, Ec =

21600 MPa (i.e. n = 9.26) and fct = 2.15 MPa. The creep coefficient associated with the initial

period of shrinkage was Mc = 1.13. The calculated average strain in the specimen prior to the

commencement of loading (accounting for the restraint provided by the reinforcement) is

calculated using Equation A.1:

H sh -6
H sh 0 209 u 10  6 (STS12) and -185×10 (STS16)
1 n U *

where n * E s / Ee 17.6 and E e E c /(1  FM ) 11350 MPa.

Axial Load versus Average strain:

The average load versus strain measured using LVDTs throughout the tests is plotted in
96

Figures 4-14 and 4-15, together with the bare bar responses for the specimens. Obviously the

bare bar response for the shrunk members is identical to that of the non-shrunk member, as

the bare bar response is unaffected by concrete deformation.

It is noted that the average axial strain in the specimens before loading (P = 0 kN) is negative

due to early shrinkage. The early shrinkage in the experiment also resulted in a significant

reduction in the load at first cracking Pcr (13.0 kN in STS12 compared to 21.1 kN in STN12

and 11.6kN in STS16 compared to 23.0kN in STN16). A total of 5 and 6 primary cracks

occurred in the specimen STS12 and STS16 respectively, as shown in the figures. The

responses of STS12 and STS16 after first cracking were similar to the non-shrunk specimens,

with the tension stiffening strain gradually diminishing and the responses of the prisms

approaching the bare bar responses as the applied load were increased.

70

60
Axial Load (kN)

C
50 Experimental Hts
40 Bare bar
Pcr.sh = 13.0 kN
30 5
2 3 4
20 B 1
10 -6
O’ Average Axial Strain (x10 )
0
-500 0 500 1000 1500 2000 2500 3000
-209

Figure 4-14: Average load P versus strain (Hs.avg) for STS12


97

120

100
C
Axial Load (kN)

80

60
Crack numbers
40
B 6 STS16
20 345 Pcr.sh = 11.6 kN Bare Bar
12

0
O’
-500 0 500 1000 1500 2000 2500 3000 3500
-185
-6
Average Axial Strain (x10 )

Figure 4-15: Average load P versus strain (Hs.avg) for STS16

Tables 4-8 and 4-9 provide values of average axial strain Hs.avg and tension stiffening strain Htsi

at selected values of total applied load P, together with the average force carried by the

concrete and the steel, Pconc and Psteel, respectively, within the shrunk prisms. Similar to the

non-shrunk members, the values of Htsi and Pconc decrease as the applied loads increase at both

crack formation stage and at the stabilised crack stage. However, compared with the non-

shrunk members, the values of Hs.avg and Psteel of the shrunk members have negative

magnitudes before cracking due to initial shrinkage, while tension stiffening is greater in

terms of the values of Htsi and Pconc before cracking. After first cracking, this difference seems

to be less significant, as the values of Htsi and Pconc drop off quickly as the applied loads

increase.
98

Table 4-8: Measured loads and strains for STS12.

Applied load, P (kN) 0 10 13.0 20 30 40 55

Average strain, Hs.avg (x 10-6) -209 -171 -154 347 934 1452 2243

Tension stiffening strain Hts (x 10-6) 209 613 729 538 393 318 191

Avge force in concrete, Pconc (kN) +4.7 13.9 16.5 12.2 8.9 7.2 4.3

Avge force in steel bar, Psteel (kN) -4.7 -3.9 -3.5 7.8 21.1 32.8 50.7

Table 4-9: Measured loads and strains for STS16.

Applied load, P (kN) 0 10 11.6 20 30 45 65 85

Average strain, Hs.avg (x 10-6) -185 -149 -140 179 469 879 1461 2049

Tension stiffening strain Hts (x 10-6) 185 398 429 319 277 240 156 65

Avge force in concrete, Pconc (kN) 7.4 16.0 17.2 12.8 11.1 9.7 6.3 2.6

Avge force in steel bar, Psteel (kN) -7.4 -6.0 -5.6 7.2 18.9 35.3 58.7 82.4

Variation of steel strains:

The variations of steel and concrete forces in the middle 600 mm of the prism at various

stages of loading are shown in Figure 4-16 and 4-17.


99

20
20

S=120mm
15
15
Force (kN)

Force (kN)
10 Steel
10
Concrete Steel
5 Concrete
5

0
200 300 400 500 600 700 800 900 0
200 300 400 500 600 700 800 900
-5
Distance along prism, mm -5
Distance along prism, mm
(a) Just before first cracking, P = 13.0 kN. (b) Just after first cracking, P = 11.1 kN.
20
20

15
15
Force (kN)

Force (kN)
Steel Steel
10 10
Concrete Concrete
5 5

0 0
200 300 400 500 600 700 800 900 200 300 400 500 600 700 800 900

-5 -5
Distance along prism, mm Distance along prism, mm
(c) Just before 2 crack, P = 15.0 kN. nd
(d) Just after 2nd crack, P = 13.3 kN.

25
25
Steel
Steel
20 Concrete 20
Concrete
Force (kN)

Force (kN)

15 15

10 10

5 5

0 0
200 300 400 500 600 700 800 900 200 300 400 500 600 700 800 900

-5 -5
Distance along prism, mm Distance along prism, mm
th
(e) Just before 4 crack, P = 18.1 kN. (f) Just after 4th crack, P = 15.9 kN.
60 60

Steel 50
50
Concrete
40
Force (kN)

40
Force (kN)

30
30 Steel
20 Concrete
20

10
10

0
0 200 300 400 500 600 700 800 900
200 300 400 500 600 700 800 900
-10
-10
Distance along prism, mm Distance along prism, mm
(g) At P = 40.0 kN. (h) At P = 50.0 kN.

Figure 4-16: Variation of forces in steel and concrete at different loading stages (STS12).
100

20 20
S=105mm
15 15
Force (kN)

Force (kN)
10 Steel 10 Steel
Concrete Concrete
5 5

0 0
200 300 400 500 600 700 800 900 200 300 400 500 600 700 800 900
-5
-5

-10
Distance along prism, mm Distance along prism, mm
-10

(a) Just before first cracking, P = 11.6 kN. (b) Just after first cracking, P = 11.0 kN.
25 25
Steel
20 20
Concrete
15 15
Steel
Force (kN)

Force (kN)
10 Concrete 10

5 5

0 0
200 300 400 500 600 700 800 900 200 300 400 500 600 700 800 900
-5 -5
Distance along prism, mm Distance along prism, mm
-10 -10

(c) Just before 2nd crack, P = 13.8 kN. (d) Just after 2nd crack, P = 12.7 kN.

25 Steel 25

Concrete Steel
20 20
Concrete
15 15
Force (kN)

Force (kN)

10 10

5 5

0 0
200 300 400 500 600 700 800 900 200 300 400 500 600 700 800 900
-5 -5
Distance along prism, mm Distance along prism, mm
-10 -10

(e) Just after 3rd crack, P = 13.0 kN. (f) Just after 4th crack, P = 15.6 kN.
80 100

80
60
Force (kN)

Force (kN)

60
40
40 Steel
Steel Concrete
20 Concrete
20

0 0
200 300 400 500 600 700 800 900 200 300 400 500 600 700 800 900
Distance along prism, mm
Distance along prism, mm
(g) At P = 50.0 kN.
(h) At P = 90.0 kN.
Figure 4-17: Variation of forces in steel and concrete at different loading stages (STS16).
101

Consider specimen STS12. Just before 1st cracking at P = 13.0kN (Figure 4-16a), the force in

the concrete is about 16.0kN, and the force in the steel is about -3.0kN. Soon after first

cracking at P = 11.0kN, the distance S away from the crack (within which concrete force is

built up through bond) is about 120mm, which is less than that of STN12 (140mm). It is

interesting to note that at the region where bond stress is not affected by the crack, the force in

the concrete exceeds the total applied load (due to the restraint to early shrinkage) and the

steel force still remains negative. As the load P increases, the local steel forces between the

cracks increase rapidly and then become positive at the crack stabilised stage (P > 40kN). A

comparison between STN12 and STS12 in terms of local concrete and steel forces at P =

40kN indicates that there is little difference between the two specimens at crack stabilised

stage. For example, the maximum forces in the concrete in-between the cracks for STN12 and

STS12 at P = 40kN are both around 19kN (Figure 4-10g and Figure 4-16g).

Crack spacing and crack width:

The crack locations and crack numbers for STS12 and STS16 are shown in Figure 4-18 and 4-

19. The maximum crack spacings for STS12 and STS16 are 245mm and 145mm, while the

average crack spacings are 177.5mm and 125mm respectively. Compared to the non-shrunk

members, the shrunk members exhibited smaller crack spacings. It has been already discussed

that the crack spacing Srm is dependent on the distance S, as indicated in the diagrams of local

steel forces and concrete forces (Figures 4-10, 4-11, 4-16, and 4-17). The values of S are

smaller in the shrunk member than in the non-shrunk member, as well as the values of Srm.

One can say that the influence of primary cracking on the local deterioration of bond around

the crack is more significant in the non-shrunk members than in the ones with significant

shrinkage. In the shrunk member, at first cracking, the tensile stress can be quickly transferred
102

to the surrounding concrete with a shorter distance S from the first crack, which indicates that

the second crack will occur in a closer distance to the first crack.

5 1 2 4 3

180 160 105 245 200 210

Figure 4-18: Crack numbers and locations (STS12).

1 3 6 4 2 5

265 140 120 85 135 145 210

Figure 4-19: Crack numbers and locations (STS16).

The crack widths at various stages of loading are given in Table 4-10 and 4-11, together with

the steel stress (Vs) at the crack at each loading stage. At the crack stabilised stage, the

measured maximum crack width is about 1.3 – 1.5 of the average crack width. For example,

at, Vs  440MPa, the average crack widths for STS12 and STS16 are 0.385mm and 0.288mm,

while the maximum crack widths are 0.525mm and 0.425mm respectively (1.36 and 1.48

times of the average values). With initial shrinkage, the member with less reinforcement ratio

(STS12) exhibited greater maximum crack width. At a steel stress of Vs  440MPa the

maximum crack width for STS12 is 0.525mm and for STS16 at Vs  440MPa, the maximum

crack width is 0.425mm.


103

Table 4-10: Crack widths at different loading stage (STS12).

Maximum
VsWidth of Width of Width of Width of Width of Average
Load crack
1st crack 2nd crack 3rd crack 4th crack 5th crack crack width
Stage (MPa) width
(mm) (mm) (mm) (mm) (mm) (mm)
(mm)
After 1st 98 0.025 - - - - 0.025 0.025
After 2nd 118 0.05 0.075 - - - 0.063 0.075
After 3rd 132 0.125 0.20 0.125 - - 0.15 0.20
After 4th 141 0.15 0.25 0.15 0.15 - 0.175 0.25
After 5th 265 0.175 0.35 0.225 0.175 0.05 0.195 0.35
40kN 354 0.125 0.40 0.50 0.375 0.30 0.34 0.50
50kN 442 0.25 0.425 0.525 0.425 0.30 0.385 0.525

Table 4-11: Crack widths at different loading stage (STS16).

Width Width Width Width Width Width Maximum


Vs Average
Load of 1st of 2nd of 3rd of 4th of 5th of 6th crack
crack width
Stage (MPa) crack crack crack crack crack crack width
(mm)
(mm) (mm) (mm) (mm) (mm) (mm) (mm)

After 1st 55 0.05 - - - - - 0.05 0.05


After 2nd 63 0.05 0.025 - - - - 0.038 0.05
After 3rd 65 0.075 0.05 0.05 - - - 0.058 0.075
After 4th 78 0.075 0.05 0.05 0.05 - - 0.056 0.075
After 5th 82 0.125 0.10 0.125 0.125 0.15 - 0.125 0.15
After 6th 136 0.125 0.125 0.125 0.15 0.175 0.025 0.121 0.175
50 kN 249 0.15 0.125 0.15 0.20 0.325 0.05 0.167 0.325
75 kN 373 0.175 0.10 0.25 0.325 0.375 0.25 0.246 0.375
90 kN 448 0.175 0.125 0.325 0.375 0.425 0.30 0.288 0.425

4.6.1.3 Comparison and further discussion

The curves in the following pages show the average concrete tensile stress versus the average

axial strain for each short-term test prism. Figure 4-20 illustrates the influence of initial

shrinkage while Figure 4-21 shows the influence of reinforcing ratio. The average concrete

stress herein is calculated by dividing the average concrete tensile force Pconc with the cross

section area.
104

2.5

Average concrete tensile stress (MPa


2.5
Average concrete tensile stress (MPa

STN16 (Experimental)
STN12 (Experimental) 2
2
STS16 (Experimental)
STS12 (Experimental)
1.5
1.5

1
1

0.5
0.5

0
0
-500 0 500 1000 1500 2000 2500
-500 0 500 1000 1500 2000 2500
Average strain ×10-6
Average strain ×10-6

(a) STN12 and STS12 (U = As/Ac = 0.011). (b) STN16 and STS16 (U = As/Ac = 0.021).

Figure 4-20: The effects of shrinkage on the average concrete tensile stress.

The dots at the left hand end of each curve in Figures 4-20 signify the situation at the

beginning of loading, where an initial tensile stress existed in the concrete due to shrinkage

(Hsh) prior to the application of the external load P. The tensile strength of concrete is a highly

variable material property and, in these four specimens, the concrete tensile stress at first

cracking was 2.00 MPa and 2.11 MPa for STN12 and STN16, respectively, and 1.67 MPa and

1.76 MPa for STS12 and STS16, respectively. The concrete in both specimens with

significant initial shrinkage (STS12 and STS16) had a significantly lower tensile strength than

the concrete in the other two specimens (despite being almost 28 days older at the time of

testing). It is generally known that tensile strength of concrete increases with time. As STS12

and STS16 were tested in 31/2 weeks after STN12 and STN16, the tensile strength of STS12

and STS16 must also be greater. The fact that the calculated tensile stresses at first cracking

for STS12 and STS16 are less than STN12 and STN16 may be attributed to an

underestimation of restraint to initial shrinkage measured during the drying period prior to

loading. Therefore, the dots on the left hand end for the curves of the shrunk member may

underestimate the initial tension.


105

After initial cracking, the average concrete stress in all specimens gradually reduced as the

deformation increased. There is minor difference of average concrete tensile stress between

the shrunk and non-shrunk specimens at in-service load levels (Figure 4-20). Although it

might appear that the concrete tensile stress (in effect the tension stiffening) decreases at a

slightly slower rate if significant shrinkage occurs prior to first cracking, but further testing is

required to gain confidence in this conclusion.

2.5 2.5
Average concrete tensile stress (MPa

Average concrete tensile stress (MPa


STN12 (Experimental) 2 STS12 (Experimental)
2
STN16 (Experimental) STS16 (Experimental)
1.5 1.5

1 1

0.5 0.5

0 0
0 500 1000 1500 2000 2500 -500 0 500 1000 1500 2000 2500
-6
Average strain ×10 Average strain ×10-6

(a) STN12 and STN16. (b) STS12 and STS16.


Figure 4-21: The effect of reinforcing ratio on the average concrete tensile stress.

The average concrete tensile stress versus axial strain curves for STN12 and STN16 (Figure

4-21a) are remarkably similar, as are the curves for STS12 and STS16 (Figure 4-21b). While

the average tensile stress in the concrete after cracking appears to be not significantly affected

by the bar diameter (and reinforcement ratio U = As/Ac), the contribution of the cracked tensile

concrete to the stiffness of the member is a larger proportion of the total stiffness when the

reinforcement ratio is smaller.

This is illustrated in Figure 4-22 where the tension stiffening strain is plotted against the

average axial strain caused by load. After cracking, the tension stiffening strain in STN12 is

significantly greater than in STN16 with the ratio between them under service conditions (500

x 10-6 < Hs.avg < 2000 x 10-6) in the range 1.7 to 1.8. It is noted that the inverse of the ratio of the
106

reinforcement ratios for each specimen is 1.79. It appears that the tension stiffening strain is

almost inversely proportional to the reinforcement ratio, with tension stiffening becoming

more significant as the reinforcement ratio decreases.

900
strai (×10 )
-6

-6 800
×10 STN12 (Experimental)
700
stiffeningstrain

STN16 (Experimental)
600
tensionstiffening

500
400
300
Tension

200
100
0
0 500 1000 1500 2000 2500
-6
Average strain ×10

Figure 4-22: Tension stiffening strain versus axial strain (STN12 and STN16).

In order to investigate the local deterioration of bond at the crack stablised stage, the tensile

forces in the concrete around the 3rd crack at 40kN and 50kN in STN12 (Figure 4-10g and h)

are plotted in Figure 4-23. The bond stress is proportional to the gradient of the tensile stress

distribution. The gradient of the tensile stress in the concrete decreases as the applied load

increases, which means that the bond stress decreases. Since no other primary crack has

occurred, this decrease of bond stress must be attributed to local damage of bond due to

microcracking at the bar interface.


107

Force (kN)
30
At 40kN
25 At 50kN
20
3rd crack section
15
10
5
0
400 500 600 700
Distance along the prism (mm)

Figure 4-23: Concrete force distribution of STN12 around crack NO. 3.

The maximum crack widths measured in the experiments and given in Tables 4-6, 4-7, 4-10

and 4-11 are plotted against steel stress at each crack in Figure 4-24 and a line of best fit for

each specimen is shown. It appears that shrinkage strain in the concrete prior to cracking has a

significant effect on maximum crack widths (as shown in Figure 4-24a and 24b). For a

particular steel stress, the crack widths are greater when the specimen has been exposed to

significant shrinkage prior to loading. The effect of changing the reinforcement ratio on crack

widths is shown in Figures 4-24c and 24d. Figure 4-24c compares the steel stress versus

maximum crack width curves for the two specimens without significant shrinkage. From

these tests it appears that reinforcement ratio has little effect. However, for the two specimens

with significant initial shrinkage, the specimen with the smaller reinforcement ratio has

significantly wider cracks, as discussed previously.


108

500 500

Steel Stress at Crack (MPa)


Steel Stress at Crack (MPa)

400 400

300 300

200 200

STN12
100 STS12 100 STN16
STS16
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6

Maximum Crack Width (mm) Maximum Crack Width (mm)

(a) STN12 and STS12. (b) STN16 and STS 16


500 500

Steel Stress at Crack (MPa)


Steel Stress at Crack (MPa)

400 400

300 STS16
300

STN12
200 200
STS12
STN16 100
100

0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6

Maximum Crack Width (mm) Maximum Crack Width (mm)

(c) STN12 and STN16 (d) STS12 and STS16


Figure 4-24: Steel stress versus maximum crack width (uniaxial tension short-term tests).

4.6.2 Long-term test results

4.6.2.1 LTN12A and LTN12C

For the long-term test, the average strain was measured at pre-selected times throughout the

period of loading using Demec gauges attached on the opposite sides of the concrete surface.

On each side, 5 Demec points were fixed to the concrete surface at 250mm centres and 4

readings were taken at each time instant. The readings were averaged to obtain the average

axial strain of the specimen (i.e. the sum of the elastic, crack opening, creep and shrinkage

strains in the concrete).

Specimen LTN12A was moist cured until two days before the test started (at age 33 days).
109

The measured elastic modulus of concrete at the time of first loading was Ec = 22400 MPa (n

= 8.93) and the tensile strength of concrete was ft = 2.04 MPa. The creep coefficient for this

initial period of drying (up to age 33 days) was 0.3 (i.e. E e = 18060 MPa, n * = 11.1). The

initial concrete tensile stress caused by restraint to shrinkage prior to loading is Vsh0= 0.11

MPa. The initial drying shrinkage strain was Hsh0 = -52 x 10-6 and, accounting for restraint

provided by the reinforcing bar, the strain in the concrete was -46 x 10-6. Specimen LTN12C

was tested right after moist curing (at age 28 days), therefore the shrinkage prior to loading is

small enough to be ignored herein.

The shrinkage strains at the end of each sustained load period were -310 x 10-6 for LTN12A

and -356 x 10-6 for LTN12C respectively. The creep coefficients associated with the sustained

load period were 1.38 and 1.11 respectively.

Both specimens were initially loaded at 5kN increments up to a load P = 40kN, at which load

5 cracks had developed, and this load level was then held constant for the duration of the tests

(i.e. For LTN12A from age 33 days to age 83 days).

Axial Load versus Average strain:

Figures 4-25 and 4-26 plot the axial load versus measured average strain responses of

LTN12A and LTN12C. It can be seen that in both tests, the loads were maintained constant at

40kN, after all primary cracks had formed. During the long-term loading period, the graph

indicates the increment of axial average strain with time. Since the bare bar strain did not vary

with time under constant axial force, tension stiffening strain must decrease with time.

For specimen LTN12A, the average axial strain immediately after application of the 40 kN

load was Hs.avg(0) = 1409 x 10-6 and the tension stiffening strain was Hts(0) = 361 x 10-6. The
110

average stress in the steel is therefore Vs(0)= Es Hs.avg(0) = 282 MPa and the average force

carried by the steel bar is Ps(0) = As Vs(0) = 31.84 kN. This means the average tensile force in

the concrete is Pc(0) = P - Ps(0) = 8.16 kN and the average concrete stress is Vc(0) = Pc(0)/Ac

= 0.825 MPa. After 50 days under load, the average axial strain had increased to Hs.avg (50) =

1556 x 10-6 and the tension stiffening strain had reduced to Hts(50) = 214 x 10-6. The

corresponding steel and concrete forces and stresses are Vs(50)= Es Hs.avg(50) = 311 MPa;

Ps(50) = As Vs(50) = 35.17 kN; Pc(50) = P - Ps(50) = 4.83 kN and the average concrete stress

is Vc(50) = Pc(50)/Ac = 0.49 MPa. It appears that for specimen LTN12A, initially with a fully

developed crack pattern, the average tensile stress in the concrete and the tension stiffening

strain reduced to 59.4% of their initial value in the first 50 days under sustained load.

As for specimen LTN12C, similar behaviour is found in the experiment. Within the first 50

days, tension stiffening strain Hts dropped from 291 x 10-6 to 207 x 10-6. Meanwhile, the

average tensile stress in the concrete Vc decreased from 0.67MPa to 0.47MPa. Both values

dropped to about 70% of its initial value during the sustained loading period.

60
LTN12A load sustained for 50 days
50
Axial Load,P (kN)

Bare bar
40

30

20

10
Average strain (x 10-6)
0
-250 250 750 1250 1750
-46

Figure 4-25: Average strain versus load response of LTN12A.


111

60

50 LTN12C Load sustained for 50 days


Bare bar
Axial load,P (kN)

40

30

20

10
Average strain (x 10-6)
0
0 500 1000 1500 2000 2500

Figure 4-26: Average strain versus load response of LTN12C.

Variation of steel and concrete stresses:

The variation of steel strains over a 600mm gauge length of the reinforcing bar was measured

using strain gauges as the time under load increased. Figures 4-27 and 4-28 show the variation

in the tensile stresses carried by concrete and steel immediately after the application of the full

sustained load and after 50 days under the sustained load for specimens LTN12A and

LTN12C, respectively.

3.0 400
Initial concrete stress (time = 0) 350
2.5
Final concrete stress (t = 50 days)
Stress (MPa)

300
Stress (MPa)

2.0
250
1.5 200

1.0 150
Initial steel stress (time = 0)
100
0.5 Final steel stress (t = 50 days)
50
0.0 0
200 300 400 500 600 700 800 900 200 300 400 500 600 700 800 900
Distance along specimen (mm) Distance along specimen (mm)
(a) Concrete stress. (b) Steel stress.
Figure 4-27: Initial and final concrete and steel stresses LTN12A (P = 40 kN).
112

3.0
400
2.5 350
Initial concrete stress (time = 0)
Final concrete stress (t = 50 days) 300
2.0
Stress (MPa)

Stress (MPa)
250
1.5 200

1.0 150 Initial steel stress (time = 0)


100 Final steel stress (t = 50 days)
0.5
50
0.0 0
200 400 600 800 200 400 600 800
Distance along specimen (mm) Distance along specimen (mm)

(a) Concrete stress (b) Steel stress.


Figure 4-28: Initial and final concrete and steel stresses LTN12C (P = 40 kN).

The results show that within the 50 days under sustained loads, significant decrease of

concrete stress and increase of steel stress are observed. For example, the maximum

instantaneous stress in the concrete of LTN12A had dropped from 1.65MPa to 0.7MPa at the

end of the test. Since the external applied loads remained constant during the tests and no

other primary cracks were observed, the loss of tensile stress in the concrete must be

attributed to the long-term decrease of bond stress. As discussed in Chapter 2, time-dependent

microcracking or cover-controlled cracking caused by the restraint to shrinkage is deemed to

be the major factor.

Crack spacing and crack width:

The crack patterns of LTN12A and LTN12C both immediately after loading and after 50 days

under sustained load are shown in Figure 4-29 and 4-30. The maximum crack spacings for

LTN12A and LTN12C are 295mm and 290mm, while the average crack spacings are 194mm

and 196mm respectively.


113

3 1 4 2 5

175 295 185 185 110 150

Figure 4-29: Crack numbers and locations (LTN12A).

5 1 2 4 3

180 190 185 290 120 135

Figure 4-30: Crack numbers and locations (LTN12C).

The crack widths at different stages for the specimens are tabulated in Table 4-12 and 4-13. It

appears that the maximum crack width of LTN12A and LTN12C increased by 80% and 16%

respectively during the sustained load period.

Table 4-12: Crack width at different stages (LTN12A)

Width of Width of Width of Width of Width of Average Maximum


Time 1st crack 2nd crack 3rd crack 4th crack 5th crack crack width crack width
(mm) (mm) (mm) (mm) (mm) (mm) (mm)
Beginning 0.075 0.125 0.5 0.375- 0.125 0.24 0.375
After 50 days 0.125 0.125 0.6 0.675 0.15 0.335 0.675

Table 4-13: Crack width at different stages (LTN12C)

Width of Width of Width of Width of Width of Average Maximum


Time 1st crack 2nd crack 3rd crack 4th crack 5th crack crack width crack width
(mm) (mm) (mm) (mm) (mm) (mm) (mm)
Beginning 0.375 0.475 0.275 0.175 0.125 0.285 0.475
After 50 days 0.55 0.5 0.325 0.125 0.225 0.345 0.55
114

4.6.2.2 LTN12B and LTN12D

Specimen LTN12B was cast, cured and tested together with LTN12A, and LTN12D was cast

cured and tested with LTN12C. Therefore, the creep and shrinkage characteristics of the

concrete in LTN12B and LTN12D were the same as those of LTN12A and LTN12C

respectively, both before and after the loading commenced. Both specimens were loaded up to

20kN, and this load level was then held constant for 50 days. Initially, a single crack occurred

in specimen LTN12B under the instantaneous load, and no cracking occurred in LTN12D.

During the period of sustained loads, three additional cracks developed in LTN12B and one

crack occurred in LTN12D primarily due to the tensile forces provided by the reinforcing bar

restraining drying shrinkage.

Axial Load versus Average strain:

Figure 4-31 and 4-32 show the load versus measured average axial strain responses of

LTN12B and LTN12D. During the long-term loading period, the graph also indicates the

increment of axial average strain with time and the decrease of tension stiffening strain.

For specimen LTN12B, the average axial strain immediately after application of the 20 kN

load was Hs.avge(0) = 89 x 10-6 and the tension stiffening strain was Hts(0) = 796 x 10-6. The

average stress in the steel is therefore Vs(0)= Es Hs.avge(0) = 17.8 MPa and the average force

carried by the steel bar is Ts(0) = As Vs(0) = 2.01 kN. This means the average tensile force in

the cracked concrete is Tc(0) = P - Ts(0) = 17.99 kN and the average concrete stress is Vc(0) =

Tc(0)/Ac = 1.82 MPa. At this time, only one crack had developed. After 50 days under load,

the average axial strain had increased to Hs.avge(50) = 239 x 10-6 and the tension stiffening

strain had reduced to Hts(50) = 646 x 10-6. The corresponding average steel and concrete forces
115

and stresses are Vs(50)= Es Hs.avge(50) = 47.8 MPa; Ts(50) = As Vs(50) = 5.40 kN; Tc(50) = P -

Ts(50) = 14.59 kN and the average concrete stress is Vc(50) = Tc(50)/Ac = 1.48 MPa. Several

additional cracks occurred with time due to restraint to shrinkage. For this relatively lightly

loaded member, the average tensile stress in the concrete and the tension stiffening strain

reduced to 81.2% of its initial value in the first 50 days under sustained load.

For specimen LTN12D, tension stiffening strain Hts dropped from 647 x 10-6 to 487 x 10-6

during the testing period. Meanwhile, the average tensile stress in the concrete Vc decreased

from 1.66MPa to 1.29MPa. Tension stiffening in LTN12D reduced to approximately 75% of

its instantaneous value in the 50 days.

30
Load sustained for 50 days LTN12B
25
Bare bar
Axial load,P (kN)

20

15

10

5
Average strain (x 10-6)
0
-200 0 200 400 600 800

Figure 4-31: Average strain versus load response of LTN12B.


116

Load (kN)
30
Load sustained for 50 days
25 LTN12D
Bare bar
20

15

10

5
Average strain (x 10-6)
0
0 200 400 600 800

Figure 4-32: Average strain versus load response of LTN12D.

Variation of steel and concrete stresses:

Figures 4-33 and 4-34 show the variation in the tensile stresses carried by concrete and steel

immediately after the application of the full sustained load and after 50 days under the

sustained load for specimens LTN12B and LTN12D, respectively. The formation of new

primary cracks with time lead to significant decreases in tension stiffening in LTN12B and

LTN12D, as time progresses.

3.0 300
Initial concrete stress (time = 0) Initial steel stress (time = 0)
2.5 250
Final concrete stress (t = 50 days) Final steel stress (t = 50 days)
Stress (MPa)
Stress (MPa)

2.0 200

1.5 150

1.0 100
0.5 50
0.0 0
200 300 400 500 600 700 800 900 200 300 400 500 600 700 800 900
Distance along specimen (mm) Distance along specimen (mm)
(a) Concrete stress. (b) Steel stress.
Figure 4-33: Initial and final concrete and steel stresses LTN12B (P = 20 kN).
117

3.0 300
Initialconcrete stress (time = 0) Initial steel stress (time = 0)
2.5 250
Final concrete stress (t = 50 days) Final steel stress (t = 50 days)

Stress (MPa)
200
Stress (MPa)

2.0

1.5 150

1.0 100

0.5 50

0.0 0
200 300 400 500 600 700 800 900 200 300 400 500 600 700 800 900
Distance along specimen (mm) Distance along specimen (mm)

(a) Concrete stress. (b) Steel stress.


Figure 4-34: Initial and final concrete and steel stresses LTN12D (P = 20 kN).

Crack spacing and crack width:

The crack pattern and crack width of both long-term tests are shown in Figure 4-35 and 4-36,

both immediately after loading and after 50 days under sustained load. The measured crack

widths at different stages of loading are summarised in Table 4-14 and 4-15.

265 835

(a) Immediately after first loading (LTN12B).

1 2 3 4

265 265 165 175 230

(b) After 50 days under load (LTN12B).

Figure 4-35: Crack numbers and crack locations (LTN12B).


118

No crack after instantaneous load

(a) Immediately after first loading (LTN12D).


1

825 275

(b) After 50 days under load (LTN12D).

Figure 4-36: Crack numbers and crack locations (LTN12D).

Table 4-14: Crack width at different stages (LTN12B).

Width of Width of Width of Width of Average Maximum


Time 1st crack 2nd crack 3rd crack 4th crack crack width crack width
(mm) (mm) (mm) (mm) (mm) (mm)
First loading 0.24 - - - 0.24 0.24
After 50 days 0.335 0.1 0.075 0.1 0.15 0.335

Table 4-15: Crack width at different stages (LTN12D)

Width of Average Maximum


Time 1st crack crack width crack width
(mm) (mm) (mm)
Beginning - - -
After 50 days 0.15 0.15 0.15

4.6.2.3 Comparison and futher discussion

Figure 4-37 shows the tension stiffening strain versus concrete age for the long-term tests.

The amount of tension stiffening strains ts at the beginning and the end of the tests are also

labeled in the graphs. The rate of reduction of ts does not appear to be influenced by the

loading levels. For example, with the identical -356 × 10-6 shrinkage occurring, tension

stiffening strain in LTN12C (P = 40kN) decreased by 29% (ts decreased from 291 × 10-6 to

207 × 10-6), while in LTN12D (P = 20kN), this value equals to 25% (ts decreased from 647 ×
119

10-6 to 487 × 10-6). This is a result as the decrease in tension stiffening with time does not

appear to depend on the number of primary cracks, lest rather on the development of

microcracks at the steel-concrete interface. Both primary cracks or microcracks cracks are

induced by the restraint to shrinkage during the sustained loading period.

The maximum crack widths measured in the experiments of LTN12A and LTN12B are plotted

against concrete age in Figure 4-38. It appears that at the stabilised crack stage (LTN12A),

crack width could be largely increased by both slip between the cracks caused by cover-

controlled cracking and the shrinkage of the concrete between the existing cracks. On the

contrary, the crack width of LTN12B widened with time at slower rate.
Tension stiffening strain (×10 -6

900
Tension stiffening strain (×10 -6)

900
646 800 LTN12C
800 647
700 LTN12D
700 487
796 600
600
500
500 361 LTN12A
400 291
400 LTN12B
300
300
200
200
214 100
100 207

0 0
33 43 53 63 73 83 28 38 48 58 68 78 88

Concrete age (days)


Concrete age (days)

(a) LTN12C and LTN12D


(a) LTN12A and LTN12B

Figure 4-37: Tension stiffening strain versus time for long-term tested prisms.
120

0.8
Maximum crack width(mm) LTN12A
0.7
0.6 LTN12B
0.5
0.4
0.3
0.2
0.1 Concrete age (days)
0
0 20 40 60 80 100

Figure 4-38: Maximum crack width versus concrete age of LTN12A and LTN12B.

4.7 SUMMARY
The experimental program and results of the time-dependent uniaxial tension tests have been

presented in this chapter. The experimental program has considered the significance of

shrinkage and the measurement of both creep and shrinkage was undertaken throughout the

tests. It has been demonstrated by the tests that shrinkage, both before and after loading, has a

profound effect on the load-deformation response of reinforced concrete in tension.

For the short-term tests, shrinkage before first loading is restrained by the embedded

reinforcement and a tensile restraining force is imposed on the concrete before any external

load is applied. This restraining force reduces the external load required to cause cracking and

induces an initial negative strain in the member. It is also observed that shrinkage has a minor

effect on tension stiffening at the stablised crack stages. Although the shrunk tension

specimens exhibited a reduced rate of change of tension stiffening decreasing rate and slightly

larger crack width after cracking, compared with the non-shrunk members, the amount of
121

tension stiffening strain does not vary too much. The short-term test results also indicate that

the tension stiffening strain is almost inversely proportional to the reinforcement ratio.

Tension stiffening becomes more significant as the reinforcement ratio in the tension prism

decreases.

For the long-term tests, the results show that tension stiffening dropped off to around 70% of

its initial values in the tested prisms after 50 days under sustained loading. It was found that

the decreasing rate of tension stiffening is not significantly influenced by the loading level.

However the major mechanisms that cause the loss of tension stiffening at different loading

stages could be different. For the specimen loaded at the crack formation stage, additional

primary cracks can significantly bring down the stress in the concrete. For the specimen

loaded at the stabilised crack stage, the loss of tension stiffening is attributed to local bond

breakdown due to microcracking. In both situations, the restraint due to shrinkage is the factor

that causes formation of new primary cracks and formation of cover-controlled cracks that

further break down bond between the steel and concrete. Therefore, consideration of drying

shrinkage is the key to understanding the long-term deformational response of reinforced

concrete in tension under typical service loads.


CHAPTER 5 EXPERIMENTAL PROGRAM
AND RESULTS OF FLEXURAL TESTS

5.1 OVERVIEW
In this chapter, the testing program for the beams and slabs is described and the results are

presented. The construction details of the test specimens and their curing and testing

progresses are first presented. Similar to the previous chapter, the material properties

measured in the experiment, including the instantaneous properties, and the creep and

shrinkage characteristics of the concrete, are next presented. The test results from the

specimens, including the moment versus deflection responses, the variation of forces in the

steel and the measured crack spacings and crack widths are then presented. Finally the results

are discussed. The effects of shrinkage and creep, the effects of primary cracks and cover-

controlled cracks on the change of tension stiffening and the time-dependent change of

stiffness are considered.

5.2 SPECIMEN CONFIGURATION


A total of six beams and four slabs were tested. All specimens were simply-supported over a

span of 2.4m and tested in four point bending. Figure 5-1, 5-2 and 5-3 show the dimensions

and details of the beam and slab specimens. Each beam had a cross section that was 200mm

wide by 400mm deep. All beam specimens were reinforced with a single layer of tensile
123

reinforcement. BSTN3-16 and BSTS3-16 were reinforced with three N16 deformed tensile

bars in the bottom of the section, while BSTN2-16 and BSTS2-16 had two N16 bottom bars.

The clear cover of the tensile bars is 50mm. All the beams were subjected to short-term

monotonic third-span point loads.

All the four slabs (SSTN4-12, SSTS4-12, SLTN4-12A, and SLTN4-12B) were identically

designed with a cross-section that was 800mm wide by 140mm deep. They were singly

reinforced containing four N12 bars at 200mm spacing in the bottom of the specimens as

shown in Figure 5-3. The clear cover of the bottom reinforcing bars is 20mm. SSTN4-12 and

SSTS4-12 were subjected to monotonically increasing load. SLTN4-12A and SLTN4-12B

were subjected to a prolonged period of sustained service load.

Figure 5-1: Dimension and detail of BSTN2-16 and BSTS2-16.


124

Figure 5-2: Dimension and detail of BSTN3-16 and BSTS3-16.

Figure 5-3: Dimension and detail of all slabs.

The first letter in the designation of each specimen indicates the specimen type, “B” for beam

and “S” for slab. The followed two letters represent the testing duration; “ST” for short-term

and “LT” for long-term. The fourth letter indicates whether or not the specimen had
125

developed significant drying shrinkage prior to the application of load; “S” if yes and “N” if

no. The next two digits are referred to as the number of tensile reinforcing bars and the bar

diameter. For instance, “2-16” means two 16mm bars reinforced on the tension side of the

specimen. For the two long-term slabs test specimens, the final letter A and B represent the

different load cases.

Local strains were also measured using strain gauges attached to the longitudinal tensile

reinforcing bars that are closest to the side surface of the specimens, thereby capturing the

influence of cracking on the local strain distribution in the tensile reinforcement. The strain

gauges were located at 40mm centres in the constant moment region.

5.3 TEST INSTRUMENTATION


A 500kN Instron actuator and a loading frame were used for the short-term tests of beams and

slabs. Figure 5-4 illustrates the set up for the short-term tests. Each specimen was simply

supported by a roller at one end and a pin at the other end. The loads were applied at the third

span points and monitored by a load cell placed on the bottom of the actuator. All loading

plates and support plates were 100 mm wide and extended across the full width of the

specimen.

A 0.3mm/min prescribed displacement was imposed by the Instron machine from the top and

transferred through a spreader beam to the two loading points on the top of the specimen. The

vertical displacements of the specimen at mid-span and at the two third-span points were

monitored with high-resolution laser transducers placed underneath the specimen (Figure 5-6).

All the measurements, including local steel strains, vertical displacement and the applied

loads, were recorded electronically by the HBM amplifier and recorded through the data
126

logger.

a) Beam set up

b) Slab set up
Figure 5-4: Set up for short-term flexural tests.
127

(a) Before loading

(b) After loading


Figure 5-5: Set up for long-term flexural tests.

The set up for the two long-term slab tests are shown in Figure 5-5. The specimen was placed

on top of two firm steel supports (one pin and one roller). At the two third-span points of the

test specimen two cross-beams were located. Pre-cast concrete blocks were carefully
128

positioned on the top of the cross-beams to produce two constant sustained point loads at the

third span points. Dial gauges were placed underneath the slabs to monitor the change in

deflection at mid span and at the third points, as shown in Figure 5-7. The strain gauges were

also connected to the HBM amplifier and the readings of steel strains are recorded throughout

the long-term testing period. Crack widths were also measured with the microscope on the

side of the members.

Figure 5-6: High-resolution laser transducers underneath the short-term testing slabs.
129

Figure 5-7: Dial gauges underneath the long-term testing slabs.

5.4 CURING AND TESING PROCEDURE


The four beams specimens were cast together from the same batch of concrete and moist

cured in a water tank to keep them wet for 14 days, so that little if any drying shrinkage

occurred during the curing period. Then on the 14th day all the beams were immediately

moved out of the tank to allow drying to start. BSTN2-16 and BSTN3-16 were placed on the

existing set up (Figure 5-4a) and tested within 24 hours. In the contrast, BSTS2-12 and

BSTS3-12 however were also simply supported, but remained unloaded for a period of 46

days while drying shrinkage developed. After this initial drying, the specimens were tested in

the loading frame. During the drying period, the change of mid-span deflection of these two

beams was monitored using dial gauges, in order to correctly characterize the influence of

shrinkage warping on the initial deflection of the specimens before loading.


130

(a) Casting of concrete

(b) Curing in the water tank


Figure 5-8: Casting and curing of the slabs

The four slabs were cast together from one batch of concrete, and cured in the water tank in

order to minimize initial shrinkage. Then on the 14th day, SSTN4-12 and SSTS4-12 were

moved out of the water tank. SSTN4-12 was immediately tested (Figure 5-4b), but SSTS4-12

was allowed to dry and unloaded for 45 days. The mid-span deflection of SSTS4-12

throughout the drying period was monitored with a dial gauge. On day 15, SLTN4-12A and

SLTN4-12B were taken out of the tank and the long-term tests commenced immediately

(Figure 5-5). Each of the long-term test specimens as loaded with concrete blocks on the top,

and the loads were maintained unchanged for 70 days. Specimens SLTN4-12A was subjected
131

to a total superimposed sustained load of 40kN, while specimen SLTN4-12B of 30kN. Table

5-1 shows the curing and testing procedure for all the flexural members in the experimental

program.

Because the creep strain starts to develop when the concrete is first loaded, whether by

external applied load or by internal restraint to shrinkage, the specimens were kept moist until

the time of loading so that creep due to load and due to restrained shrinkage started at the

same time. In this way the creep coefficient measured on the companion specimen would be

applicable to both the load and the shrinkage induced stress. In practice, creep and shrinkage

strains are likely to develop before any service loads are applied, and this may significantly

affect the deflections and lower the cracking moment.

Concrete properties were measured on companion specimens, including concrete strengths,

creep and shrinkage. The methods to measure creep strain and shrinkage strain were identical

to those described in Chapter 4 for the uniaxial tension specimens (Figure 4-5). The test

results are presented in the following sections.

Table 5-1: Testing procedure of all the specimens

Specimens Day 1-14 Day 14 Day 60


BSTN2-16
Tested -
BSTN3-16 Wet cured in
BSTS2-16 the water tank
Start drying Tested
BSTS3-16
(a) Beams

Specimens Day 1-14 Day 14 Day 15 Day 59 Day 85


SSTN4-12 Tested - - -
SSTS4-12 Wet cured Start drying - Tested
in the water
SLTN4-12A tank Start test End the
-
SLTN4-12B and drying test
(b) Slabs
132

5.5 MATERIAL PROPERTIES


5.5.1 Concrete instantaneous properties

The instantaneous properties of concrete measured in the companion tests on concrete

cylinders and prisms are given in Table 5-2. The concrete strength of the beams increased by

approximately 50% during the drying period. For the slabs, the concrete strength increased by

about 15% throughout the sustained load test. The elastic modulus of the concrete also

increased by 28% for the beams and by 16% for the slabs during the drying period. The

corresponding increase in the tensile strength, both indirect tensile strength and flexural

tensile strength, was about 10% for both beams and slabs.

Table 5-2: Instantaneous properties of concrete (flexural specimens)

Specimen f cp f ct f cf Ec

BSTN2-16
26.0 2.2 3.5 25000
BSTN3-16
BSTS2-16
40.0 2.4 3.9 32000
BSTS3-16
SSTN4-12 28.6 1.8 3.0 25000
SSTS4-12 32.6 2.0 3.3 29000
(a) Short-term flexural specimens
f cp f ct f cf Ec
Specimen
Start End Start End Start End Start End
SLTN4-12A
28.6 33.7 1.8 2.2 3.0 3.3 25000 29500
SLTN4-12B

(b) Long-term flexural specimens

* All units in MPa.


133

5.5.2 Creep and Shrinkage

For the concrete in the beams and slabs, creep was measured on two identical specimens. The

concrete was first loaded at age 14 days as soon as the sample started drying. The calculated

creep coefficients of the plain concrete from the creep tests for both beams and slabs are

plotted in Figure 5-9. Also plotted in the diagrams are the averaged values from the two creep

tests. According to the diagrams, the averaged c at the ends of the drying days for the short-

term beams and slabs tests were 0.92 and 1.37 respectively. By the end of long-term tests of

the slabs at day 85, c had reached 1.47.

c 1.2
c 1.6
1.0
1.2
0.8
Creep test 1
0.6 0.8 Creep test 2
Creep test 1
0.4 Averaged
Creep test 2
0.4
0.2 Averaged

0.0 0.0
14 34 54 74 0 20 40 60 80
Concrete age (days) Concrete age (days)

(a) Beams (b) Slabs

Figure 5-9: Creep coefficients for the flexural tests.

Figure 5-10 shows the measured drying shrinkage strains and their averaged values for the

beams and slabs. The averaged shrinkage strain for specimen BSTS2-16 and BSTS3-16

during the drying period (day 60) was -220 × 10-6. In the time interval between the two short-

term slab tests (from day 15 to day 59), the averaged drying shrinkage strain was -320 × 10-6

and at the end of the long-term tests (day 85), the averaged drying shrinkage strain was -400 ×

10-6.
134

300 500
Shrinkage strain (×10-6

Shrinakge strain ( ×10-6


400
200
300

200
100 Shrinkage test 1
Shrinkage test 2 100
Averaged
0
0
0 20 40 60 80 100
0 20 40 60 80 Concrete age (days)
Concrete age (days)
(b) Slabs
(a) Beams
Figure 5-10: shrinkage strain development for the flexural tests.

5.6 TEST RESULTS FOR THE FLXURAL


SPECIMENS
5.6.1 Short-term results

5.6.1.1 Members without initial shrinkage (BSTN2-16, BSTN3-16 and SSTN4-12)

BSTN2-16, BSTN3-16 and SSTN4-12 were tested soon after the end of moist curing before

significant shrinkage could occur. They were cured in a water tank for 14 days and then tested.

Minimum drying shrinkage and shrinkage-induced initial deflection sh0 occurred before the

tests.

Moment versus Mid-span deflection:

Figure 5-11, 5-12 and 5-13 display the global moment versus mid-span deflection responses

from the tests, together with the calculated fully-cracked response. The fully-cracked response

is obtained by assuming that all the sections along the member are fully-cracked and the

tensile concrete carries no tension. The mid-span deflection of the fully-cracked beam is

obtained by integrating the curvature (M/EcIcr) over the span of each member (2.4m). Thus,
135

the instantaneous tension stiffening deflection tsi at a particular mid-span moment is the

difference between the fully-cracked deflection and actual member deflection, as seen in the

diagrams.

As can be seen, the member behaviour is approximately linear-elastic before first cracking

(curve OB) and behave in a nonlinear manner after cracking (curve BC). The maximum

tension stiffening deflection tsi.max occurs at the onset of first cracking (i.e. at M = Mcr). Then,

as the load is increased, tsi gradually diminishes and the moment-deflection curve approaches

the fully-cracked response as the moment increases.

Table 5-3 summarises the experimentally measured mid-span deflections exp, the calculated

fully-cracked deflections fc and the corresponding tension stiffening deflections tsi after first

cracking for each specimen. The magnitude of tension stiffening deflections decreases as

applied moment increases, and this indicates the loss of tension stiffening. As the applied

moments increased from Mcr to 2.0Mcr, tsi decreases by 60%, 56% and 75% for BSTN2-16,

BSTN3-16 and SSTN4-12 respectively. As the moment approaches the magnitudes where the

steel bars approaches yielding, tsi approaches zero.


136

Moment (kN.m)

60

C
50

40
M cr = 22.1kN.m  tsi
30
B
20 BSTN2-16
 tsi.max
Fully-cracked
10
O Mid-span deflection (mm)
0
0 1 2 3 4 5 6

Figure 5-11: Moment versus mid-span deflection (BSTN2-16)


Moment (kN.m)

70

C
60

50

40 M cr = 26.6kN.m
 tsi
B
30
BSTN3-16
20  tsi.max
Fully-cracked

10
O Mid-span deflection (mm)
0
0 1 2 3 4 5

Figure 5-12: Moment versus mid-span deflection (BSTN3-16)


137

25
Moment (kN.m)

C
20

15
M cr =9.1kN.m
tsi tsi
B SSTN4-12
10
Bare bar
Fully-cracked
tsi.max
tsi.max
5
Mid-span deflection (mm)
0 O
0 3 6 9 12 15 18

Figure 5-13: Moment versus mid-span deflection (SSTN4-12)

Table 5-3: Deflections of short-term non-shrunk flexural members.

Specimen Ms kN.m exp mm fc mm tsi mm fc / exp


At Mcr (22.1kN.m) 0.82 2.16 1.34 (max) 2.63
1.25 Mcr (27.6kN.m) 1.41 2.70 1.29 1.91
BSTN2-16 1.5 Mcr (33.2kN.m) 2.24 3.23 0.99 1.44
1.75 Mcr (38.7kN.m) 3.00 3.78 0.78 1.26
2.0 Mcr (44.2kN.m) 3.78 4.31 0.53 1.14
At Mcr (26.6kN.m) 0.65 1.88 1.23 (max) 2.90
1.25 Mcr (33.3kN.m) 1.37 2.35 0.99 1.72
1.5 Mcr (39.9kN.m) 1.92 2.82 0.90 1.47
BSTN3-16
1.75 Mcr (46.6kN.m) 2.55 3.29 0.74 1.29
2.0 Mcr (53.2kN.m) 3.23 3.76 0.53 1.16
60kN.m 3.92 4.22 0.3 1.08
At Mcr (9.1kN.m) 1.95 6.87 4.92 (max) 3.52
1.25 Mcr (11.4kN.m) 5.50 8.53 3.27 1.55
SSTN4-12 1.5 Mcr (13.7kN.m) 7.91 10.26 2.35 1.30
1.75 Mcr (15.9kN.m) 9.88 11.92 2.04 1.21
2.0 Mcr (18.2kN.m) 12.02 13.65 1.23 1.14
138

Variation of steel forces:

Figure 5-14, 5-15, and 5-16 show the variations of steel forces in the middle 800 mm of each

specimen (i.e. the constant moment region) at various stages of loading. The steel force was

obtained directly from the measured local steel strains. The variation of steel force depends on

the locations of the cracks, with the tensile force in the steel at a peak at each crack. Away

from the cracks, tensile stress in the concrete is built up through bond, and the tensile stress in

the steel decreases. Since the strain gauges on the reinforcement in the flexural tests were

located at 40mm centres, there were only 3 – 6 measurements of local steel strains between

adjacent cracks. Therefore, it is difficult to accurately access the actual distance S away from

the primary cracks, where bond is affected. Future research can try to improve this by using

more strain gauges at closer centres without removing the ribs or reducing the local bond

mechanism.

Unlike the uniaxial tension members, the tensile stress in the concrete at typical sections

along the flexural members can not be directly obtained from the measured steel strains,

because the area of concrete under tension, as well as the tensile stress gradient, is also

uncertain. Therefore, another approach to represent the quantity of local tension stiffening is

to calculate the average tensile force carried by the concrete between the cracks.

Table 5-4 tabulates the average tensile bar force Ts.avge in BSTN2-16, BSTN3-16 and SSTN4-

12 at certain loading stages after first cracking. Ts.avge was determined by averaging the

measured tensile force in the steel over the constant moment region, and the average tensile

force carried by the concrete Tc.avge was determined by subtracting the average force in the

steel from the maximum force in the steel at each primary crack (where concrete is assumed

to carry no tensile force). The results show that after cracking, Tc.avge decreases with
139

increasing moment, which is indicated by the reduction of the proportion Tc.avge/Ts.max. For

example, as the applied moment increases from 16.0kN.m to 20.0kN.m in SSTN4-12, the

value of Tc.avge/Ts.max decreases from 0.15 to 0.12. Notice that all the cracks had formed after

M = 16.0kN.m, thus the average tensile stress in the concrete drops off at the stabilised crack

stages, which is similar to the uniaxial tension test results.

Table 5-4: Average tensile forces in the concrete and the steel over the constant moment
region (BSTN2-16, BSTN3-16 and SSTN4-12).

M Ts.max Ts.ave Tc.avg Tc.avg /


Specimens
(kN.m) (kN) (kN) (kN) Ts.max
28.8 40.0 26.6 13.4 0.33
BSTN2-16 40.0 59.0 51.1 7.9 0.13
51.2 79.0 71.8 7.2 0.09
40.0 40.5 25.9 14.6 0.36
BSTN3-16 51.2 52.0 41.0 11.0 0.21
60.8 62.5 53.5 9.04 0.14
12.0 28.6 20.8 7.8 0.27
SSTN4-12 16.0 38.2 32.4 5.8 0.15
20.0 47.7 41.9 5.8 0.12
140

20 30
25
15
Force (kN)

Force (kN)
20
10 15
10
5
5
0 0
-0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4 -0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4
Distance from the mid span (m) Distance from the mid span (m)

(b) Just after 1st and 2nd cracking, M = 22.1 kN.


(a) Just before 1st cracking, M = 22.0 kN.m.
40 50

30 40
Force (kN)

Force (kN)
30
20
20
10
10
0 0
-0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4 -0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4
Distance from the mid span (m) Distance from the mid span (m)

(c) Just after 3rd crack, M =.26kN.m (d) After 6th crack, M = 28.8 kN.m.

70 90
60 80
Force (kN)

Force (kN)

50 70
40 60

30 50

20 40
-0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4 -0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4
Distance from the mid span (m) Distance from the mid span (m)

(e) After 10th crack, M = 40kN.m (f) At M = 51.2kN.m


Figure 5-14: Variation of forces in steel at different stages of loading (BSTN2-16).
141

20 30
25
15

Force (kN)
Force (kN)

20
10 15
10
5
5
0 0
-0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4 -0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4
Distance from the mid span (m) Distance from the mid span (m)

(a) Just before first cracking, M = 24.6 kN.m. (b) Just after first cracking, M = 26.6 kN.

30 50
25 40
Force (kN)
Force (kN)

20
30
15
20
10
5 10
0 0
-0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4 -0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4
Distance from the mid span (m) Distance from the mid span (m)

(c) After 3rd and 4th cracks, M = 28.2 kN.m. (d) At M = 40 kN.m.
60 70

50 60
Force (kN)

Force (kN)

40 50

30 40

20 30

10 20
-0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4 -0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4
Distance from the mid span (m) Distance from the mid span (m)

(e) At M = 51.5 kN.m. (f) At M = 60.8 kN.m.

Figure 5-15: Variation of forces in steel at different stages of loading (BSTN3-16).


142

20 30
25
15
Force (kN)

Force (kN)
20
10 15
10
5
5
0 0
-0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4 -0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4
Distance from the mid span (m) Distance from the mid span (m)

(a) Just before first cracking, M = 8.5 kN.m. (b) Just after first cracking, M = 9.1 kN.m.

30 40
25
30
Force (kN)

Force (kN)
20
15 20
10
10
5
0 0
-0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4 -0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4
Distance from the mid span (m) Distance from the mid span (m)

(c) Just after 2nd crack, M = 9.8 kN.m. (d) at M = 12 kN.m.

50 60

40 50
Force (kN)

Force (kN)

30 40

20 30

10 20
-0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4 -0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4
Distance from the mid span (m) Distance from the mid span (m)

(e) at M = 16 kN.m. (f) at M = 20 kN.m.

Figure 5-16: Variation of forces in steel at different stages of loading (SSTN 4-12).
143

Crack spacing and crack width:

The crack patterns of BSTN2-16, BSTN3-16 and SSTN4-12 at the end of the tests are shown

in Figure 5-17, 5-18 and 5-19. Within the constant moment regions, the numbers of cracks in

each specimen are 5, 6 and 7. The crack spacings shown in the figures are measured at the

levels of the tensile reinforcing bars. The maximum crack spacings of BSTN2-16, BSTN3-16

and SSTN4-12 in the constant moment region are 243mm, 153mm and 157mm respectively.

The average crack spacings are 154mm, 121mm and 122mm respectively. Similar to the

observation in uniaxial tension tests, the crack spacing in the members with the smaller

reinforcement ratio is greater than for the members with larger , i.e. more cracks had

formed in the latter. This indicates that in the flexural tests, the length S away from the

primary cracks where significant bond stresses develop is also dependent on the

reinforcement ratio.

Figure 5-17: Crack numbers and crack locations (BSTN2-16).


144

Figure 5-18: Crack numbers and crack locations (BSTN3-16).

Figure 5-19: Crack numbers and crack locations (SSTN-4-12).

The cracked widths at the different loading levels are tabulated in Table 5-5, 5-6 and 5-7.

Only the crack widths within the constant moment regions are shown, together with the

calculated steel stress at the cracked sections (Vs). Generally the crack width increases as the

steel stress Vs increases. Compared with the uniaxial tension members, the difference between

the maximum crack width and the average crack width is not that significant in the flexural

members. The ratio of maximum crack width over average crack width after all the cracks had

formed is about 1.25. In addition, the study of the crack width on the beams shows that at the

same steel stress level, maximum crack width seems independent of the reinforcement ratio.
145

This phenomenon will be discussed further.

Table 5-5: Crack width at different stages (BSTN2-16).

Load Vs Width of Width of Width of Width of Width of


Average crack
Maximum
1st crack 2nd crack 3rd crack 6th crack 10th crack crack width
Stage (MPa) width (mm)
(mm) (mm) (mm) (mm) (mm) (mm)

25.6kN.m 184 0.01 0.025 0.05 - - 0.028 0.05


36.8kN.m 259 0.05 0.075 0.1 0.05 - 0.069 0.1
44.8kN.m 318 0.1 0.115 0.15 0.1 0.125 0.118 0.15

Table 5-6: Crack width at different stages (BSTN3-16).

Load Vs Width of Width of Width of Width of Width of Width of Average Maximum


1st crack 3rd crack 4th crack 8th crack 9th crack 10th crack crack width crack width
Stage (MPa) (mm) (mm) (mm) (mm) (mm) (mm) (mm) (mm)

26kN.m 125 0.033 0.02 0.01 - - - 0.021 0.033

36kN.m 180 0.05 0.025 0.015 0.05 0.05 - 0.038 0.05


46kN.m 232 0.075 0.05 0.025 0.075 0.1 0.05 0.0625 0.1
60kN.m 317 0.125 0.1 0.125 0.15 0.125 0.075 0.117 0.15

Table 5-7: Crack width at different stages (SSTN4-12).

Average Maximum
Vs Width of Width of Width of Width of Width of Width of Width of
Load crack crack
1st crack 2nd crack 3rd crack 4th crack 5th crack 6th crack 12th crack
Stage (MPa) width width
(mm) (mm) (mm) (mm) (mm) (mm) (mm)
(mm) (mm)

12kN.m 253 0.05 0.075 0.1 0.1 0.075 0.075 - 0.08 0.075
16kN.m 338 0.175 0.175 0.15 0.15 0.125 0.1 0.075 0.136 0.175
20kN.m 422 0.225 0.25 0.2 0.225 0.225 0.15 0.125 0.20 0.25

5.6.1.2 Members with significant initial shrinkage (BSTS2-16, BSTS3-16 and SSTS4-12)

Specimens BSTS2-16, BSTS3-16 and SSTS4-12 were allowed to dry for a period of time (45-

46 days) before they were tested. At the time of testing days, significant shrinkage strains (-

220 × 10-6 for beams and -320 × 10-6 for slabs) had developed prior to loading. The

corresponding creep coefficients were 0.92 and 1.37 respectively. The initial deflection Gsh.0

caused by shrinkage warping was also measured by the dial gauges located underneath the
146

mid span of the specimens and the results were Gsh.0 = 0.17mm, 0.14mm and 0.65mm for

BSTS2-16, BSTS3-16 and SSTS4-12, respectively. Therefore, at zero moment (M = 0), the

specimens had already undergone some initial deflection.

Moment versus Mid-span deflection:

The moment versus mid-span deflection measured throughout the tests are plotted in Figure

5-20, 5-21 and 5-22. Also plotted are the corresponding fully-cracked responses. It has been

mentioned in Section 3.3.3 that the fully-cracked response for short-term flexural members

with initial shrinkage must be shifted to the right by an amount of Gsho.cr, in order to consider

the effect of initial shrinkage on the deflection of a fully-cracked member. The procedure to

calculate Gsho.cr can be found in Appendix B. The values of Gsho.cr are also shown in the figures.

It is also clear in these diagrams that the structural behaviour moves from linear to nonlinear

at the onset of cracking. At the linear-elastic stage (OB), the specimens are relatively stiff and

little deflection has occurred compared with the nonlinear stage (BC). Early shrinkage before

loading can greatly affect the member response. Compared with the specimens without early

shrinkage, the cracking moment Mcr.sh is significantly reduced. For example, the cracking

moment of BSTS2-16 (17.0kN.m) is much smaller than BSTN2-16 (22.1kN.m). The tension

stiffening deflection tsi also gradually diminish after cracking and the experimental responses

approach the fully-cracked responses.

Table 5-8 provide values of measured mid-span deflections exp, calculated fully-cracked

deflections fc and instantaneous tension stiffening deflections tsi at selected values of total

applied moment M. Similar to the members without shrinkage, the values of tsi decrease with

increasing applied moments after first cracking. As the applied moment increases from Mcr to
147

2 Mcr, tsi drops by 55%, 36% and 45% for BSTS2-16, BSTS3-16 and SSTS4-12 respectively.
Moment (kN.m)

60

BSTS2-16
50
Fully-cracked C

40

30
M cr.sh =17.0kN.m  tsi
B
20
 tsi.max
10

O
 sh0.cr = 0.5mm Mid-span deflection (mm)
0
0 1 2 3 4 5 6
 sh.cr = 0.17mm

Figure 5-20: Moment versus mid-span deflection for BSTS2-16.


148

70
Moment (kN.m)

C
60

50

40
 tsi
M cr.sh =21.4kN.m
30
B BSTS3-16
20  tsi.max
Fully-cracked
10
O  sh0.cr = 0.5mm Mid-span deflection (mm)
0
0 1 2 3 4 5 6
0.14
 sh.cr = 0.17mm

Figure 5-21: Moment versus mid-span deflection for BSTS3-16.

25
Moment (kN.m)

SSTS4-12
20 Bare bar
Fully-cracked C

15
M cr.sh = 6.0kN.m  tsi
10
B

5  tsi.max
Mid-span deflection (mm)
O  sh0.cr = 2mm
0
0 3 6 9 12 15 18
 sh0 = 0.65mm

Figure 5-22: Moment versus mid-span deflection for SSTS4-12.


149

Table 5-8: Deflections of short-term shrunk flexural members.

Specimen Ms kN.m exp mm fc mm tsi mm fc / exp


At Mcr (17.0KN.m) 0.67 2.26 1.59 (max) 3.37
1.25 Mcr (21.3kN.m) 1.34 2.63 1.29 1.96
1.5 Mcr (25.5kN.m) 1.88 3.04 1.16 1.62
BSTS2-16
1.75 Mcr (29.8kN.m) 2.42 3.45 1.03 1.43
2.0 Mcr (34.0kN.m) 3.16 3.88 0.72 1.23
45kN.m 4.64 4.94 0.30 1.06
At Mcr (21.4kN.m) 1.02 2.23 1.21 (max) 2.19
1.25 Mcr (26.8kN.m) 1.34 2.52 1.18 1.88
1.5 Mcr (32.1kN.m) 1.82 2.92 1.10 1.60
BSTS3-16 1.75 Mcr (37.5kN.m) 2.44 3.31 0.87 1.36
2.0 Mcr (42.8kN.m) 2.94 3.71 0.77 1.26
50kN.m 3.68 4.23 0.55 1.15
60kN.m 4.76 4.96 0.2 1.04
At Mcr (6.0 kN.m) 1.64 6.55 4.91 (max) 4.00
1.25 Mcr (7.5kN.m) 2.78 7.64 4.86 2.75
1.5 Mcr (9.0kN.m) 4.46 8.84 4.38 1.98
SSTS4-12
1.75 Mcr (10.5kN.m) 6.24 9.90 3.66 1.59
2.0 Mcr (15.0kN.m) 10.53 13.25 2.72 1.26
18.2kN.m 13.84 15.7 1.86 1.13

Variation of steel strains:

The variations of steel forces in the middle 800 mm of the beams at various stages of loading

are shown in Figure 5-23, 5-24 and 5-25. Before loading, since the reinforcing bars were

subjected to compressive force due to the shrinkage in the concrete, the steel strains were

negative. Therefore, the shrinkage-induced compressive strains must be subtracted from the

measured steel strains to obtain the actual strains at a typical moment. The procedure to

calculate the time-dependent initial compressive steel strains of the uncracked beam or slab

sections is explained in Appendix B.


150

10 30

5 20

Force (kN)
Force (kN)

0 10
-0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4
-5 0
-0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4
-10 -10
Distance from the mid span (m) Distance from the mid span (m)

(a) Just before first cracking, M = 16.7kN.m. (b) Just after first cracking, M = 17.0 kN.m.

30 50
40
20 Force (kN)
Force (kN)

30
10 20

10
0
-0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4 0
-10 -0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4
Distance from the mid span (m) Distance from the mid span (m)

(c) Just after 2nd crack, M = 18.9 kN.m. (d) At M = 28.8 kN.m.

70 90
60 80
Force (kN)

Force (kN)

50 70
40 60
30 50
20 40
-0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4 -0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4
Distance from the mid span (m) Distance from the mid span (m)

(e) At M = 40kN.m. (f) At M = 51.2 kN.m.

Figure 5-23: Variation of forces in steel at different stages of loading (BSTS2-16).


151

20 30

20
Force (kN)

10

Force (kN)
10
0
-0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4 0
-0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4
-10 -10
Distance from the mid span (m) Distance from the mid span (m)

(a) Just before first cracking, M = 19.4kN.m. (b) Just after 3rd cracking, M = 21.4 kN.m.

30 50

40
20
Force (kN)
Force (kN)

30
10 20

0 10
-0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4 0
-10 -0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4
Distance from the mid span (m) Distance from the mid span (m)
(c) Just after 4thcrack, M = 28.4 kN.m. (d) At M = 40 kN.m.
70 70
60 60
Force (kN)

Force (kN)

50 50
40 40
30 30
20 20
10 10
-0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4 -0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4
Distance from the mid span (m) Distance from the mid span (m)

(e) At M = 51.2kN.m. (f) At M = 60.8 kN.m.


Figure 5-24: Variation of forces in steel at different stages of loading (BSTS3-16).
152

20 20

15 15

Force (kN)
Force (kN)

10 10
5 5
0 0
-0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4 -0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4
-5 -5
Distance from the mid span (m) Distance from the mid span (m)

(a) Just before first cracking, M = 5.9kN.m. (b) Just after first cracking, M = 6.0 kN.m.

40
25
30
Force (kN)
Force (kN)

15 20

5 10

0
-5-0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4
-0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4
Distance from the mid span (m)
Distance from the mid span (m)
(c) Just after 3rd crack, M = 8.0 kN.m. (d) At M = 12 kN.m.
50 60

40 50
Force (kN)
Force (kN)

30 40

20 30

10 20
-0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4 -0.4 -0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4
Distance from the mid span (m) Distance from the mid span (m)

(e) At M = 16kN.m. (f) At M = 20.0 kN.m.

Figure 5-25: Variation of forces in steel at different stages of loading (SSTS4-12).

Table 5-9 tabulates the calculated average tensile forces in the concrete and the steel over the

constant region for BSTS2-16, BSTS3-16 an SSTS4-12. The comparison with Table 5-4

shows that the average tensile force in the concrete Tc.avg is larger in the members that had

undergone significant shrinkage. For instance, the Tc.avg in BSTS2-16 at M = 40.0kN is


153

16.64kN, while it is only 7.9kN in BSTN2-16 at the same moment level. The increase of the

average tensile force in the concrete in the member with early shrinkage is attributed to tensile

force induced by the restraint to initial shrinkage by the reinforcing bars.

Table 5-9: Average tensile forces in the concrete and the steel over the constant moment
region (BSTS2-16, BSTS3-16 and SSTS4-12).

M Ts.max Ts.ave Tc.avg Tc.avg /


Specimens
(kN.m) (kN) (kN) (kN) Ts.max
28.8 41.0 20.40 20.60 0.50
BSTS2-16 40.0 59.8 43.16 16.64 0.28
51.2 77.2 61.57 15.63 0.20
40.0 40.7 25.62 15.08 0.37
BSTS3-16 51.2 55.5 41.55 13.95 0.25
60.8 64.0 51.07 12.92 0.20
12.0 29.3 20.03 9.27 0.32
SSTS4-12 16.0 39.1 31.05 8.04 0.20
20.0 49.1 41.07 8.03 0.16

Crack spacing and crack width:

The crack locations and crack numbers for BSTS2-16, BSTS3-16 and SSTS4-12 are shown in

Figure 5-26, 5-27 and 5-28, and the crack widths at various stages of loading are given in

Table 5-10, 5-11 and 5-12, together with the steel stress (Vs) at the crack at each loading stage.

The members with early shrinkage exhibited similar numbers of cracks to the members

without shrinkage, except for BSTS2-16 which had two more cracks than BSTN2-16. The

maximum crack spacings in the constant moment regions for BSTS2-16, BSTS3-16 and

SSTS4-12 are 122mm, 162mm and 168mm respectively, and the average crack spacings are

110mm, 133mm and 139mm respectively. The difference of crack spacings between the

members with early shrinkage and without shrinkage can be hardly told.
154

Figure 5-26: Crack numbers and locations (BSTS2-16)

Figure 5-27: Crack numbers and locations (BSTS3-16)

Figure 5-28: Crack numbers and locations (SSTS4-12)


155

Table 5-10: Crack width at different stages (BSTS2-16)

Vs Width of Width of Width of Width of Width of Width of


Average Maximum
Load crack crack
(MPa) 1st crack 2nd crack 3rd crack 4th crack 6th crack 9th crack
Stage width width
(mm) (mm) (mm) (mm) (mm) (mm)
(mm) (mm)

20.0kN.m 156 0.025 0.025 0.05 0.0375 - - 0.0344 0.05


30.0kN.m 233 0.075 0.05 0.0625 0.1 0.05 - 0.0675 0.1
40.0 kN.m 310 0.125 0.05 0.1125 0.125 0.1 - 0.1025 0.125
48.0kN.m 371 0.15 0.05 0.125 0.15 0.1 0.0375 0.1021 0.15

Table 5-11: Crack width at different stages (BSTS3-16)

Vs Width of Width of Width of Width of Width of


Average Maximum
Load crack crack
(MPa) 1st crack 2nd crack 3rd crack 4th crack 7th crack
Stage width width
(mm) (mm) (mm) (mm) (mm)
(mm) (mm)

30.0kN.m 159 0.0375 0.025 0.025 0.0125 - 0.025 0.0375

40.0kN.m 212 0.05 0.0375 0.075 0.075 0.05 0.0575 0.075


48.0kN.m 252 0.1 0.05 0.125 0.075 0.075 0.085 0.125
60.0kN.m 317 0.15 0.1 0.175 0.15 0.125 0.140 0.175

Table 5-12: Crack width at different stages (SSTS4-12)

Average Maximum
Vs Width of Width of Width of Width of Width of Width of
Load crack crack
1st crack 2nd crack 3rd crack 5th crack 7th crack 8th crack
Stage (MPa) width width
(mm) (mm) (mm) (mm) (mm) (mm)
(mm) (mm)

12kN.m 253 0.125 0.075 0.125 0.125 0.075 0.05 0.096 0.125
16kN.m 338 0.175 0.1 0.175 0.15 0.125 0.075 0.133 0.175
20kN.m 442 0.275 0.175 0.275 0.25 0.25 0.15 0.229 0.275

5.6.1.3 Comparison and further discussion

The tension stiffening deflections given in Table 5-3 and 5-9 are plotted against the calculated

maximum steel stresses at the cracks in Figure 5-29 and 5-30. Figure 5-29 demonstrates the

influence of initial shrinkage on the short-term tension stiffening deflections. The dots at the

left hand end of each curve in the figure signify the situation at the onset of first cracking,

where the tension stiffening deflections are at maximum. It appears that with the same steel
156

stresses at cracks, the tension stiffening deflections are not significantly affected by the

amount of early shrinkage. For the beam specimens, tension stiffening deflections appear to

be slightly greater in the beams without shrinkage (Figure 5-29 (a) and (b)). This is typically

obvious in the specimens BSTS2-16 and BSTN2-6. At the same stress level after first

cracking, the calculated tension stiffening deflection in BSTN2-16 is about 0.2mm greater

than that in BSTS2-16 at the same steel stress level. For BSTN3-16 and BSTS3-16, tension

stiffening deflection in the former specimen is only about 0.1mm greater than that in the latter.

In the slabs, the difference of tension stiffening deflection varies at different steel stress levels.

As steel stress is less than 220MPa, SSTN4-12 exhibits greater tension stiffening.

Nevertheless, the tension stiffening deflection is higher in SSTS4-12 has the steel stress

exceeds 220MPa.

The initial shrinkage increases the tensile stress in the concrete before cracking, and thus

increases tension stiffening. After cracking, some tensile stress in the concrete between the

cracks caused by shrinkage still exists. After first crack develops, the tensile stress in the

concrete away from the crack in the member with shrinkage will be greater than for a similar

member without shrinkage, if good bond conditions exist. However, if this tensile stress is

greater than the tensile strength of concrete at the weakest section, a second crack occurs and

tension stiffening is gradually broken down. On the contrary, if the tensile stress produced by

restraint to shrinkage is not great enough or bond condition is not good, additional cracks will

not occur and tension stiffening will not be reduced further.


157

2 2
Tension stiffening

Tension stiffening
deflection (mm)

deflection (mm)
1.6 1.6

1.2 1.2

0.8 0.8
BSTN2-16 BSTN3-16
0.4 0.4
BSTS2-16 BSTS3-16
0 0
0 100 200 300 400 0 100 200 300 400
Tensile steel stress at crack (MPa) Tensile steel stress at crack (MPa)

(a) BSTN2-16 and BSTS2-16 (b) BSTN3-16 and BSTS3-16

6
Tension stiffening
deflection (mm)

SSTN4-12
4
SSTS4-12

0
0 100 200 300 400 500
Tensile steel stress at crack (MPa)

(c) SSTN4-12 and SSTS4-12.


Figure 5-29: The effects of initial shrinkage on the tension stiffening deflections.

2
Tension stiffening

2
Tension stiffening

deflection (mm)
deflection (mm)

1.6
1.6

1.2
1.2

0.8
0.8
BSTS2-16
BSTN2-16 0.4
0.4 BSTS3-16
BSTN3-16
0
0
0 100 200 300 400
0 50 100 150 200 250 300 350
Tensile steel stress at crack (MPa)
Tensile steel stress at crack (MPa)

(a) BSTN2-16 and BSTN3-16 (b) BSTS2-16 and BSTS3-16


Figure 5-30: The effects of reinforcing ratio on the tension stiffening deflections.
158

Figure 5-30 illustrates the effects of reinforcing ratio on short-term tension stiffening

deflections after cracking. It can be seen that the tension stiffening deflection is greater in the

specimens with the smaller reinforcing ratio. The tension stiffening deflections of BSTN2-16

and BSTS2-16 (with reinforcement ratio U = As/Ac= 0.005) is approximately 20% higher than

that of BSTN3-16 and BSTS3-16 (U = 0.0075) at the same steel stress level. This is similar to

that relating to axial deformation in the uniaxial tension tests.

The maximum crack widths measured in the experiments and given in Tables 5-5 to 5-7, and

Tables 5-10 to 5-12 are plotted against steel stress in Figure 5-31 and a best fit curve for each

specimen is also plotted. Unlike the uniaxial tension members, the effect of initial shrinkage

on maximum crack width is not very significant. At a particular steel stress at the crack

section, the crack widths of the shrunk and non-shrunk specimens are of little difference, as

seen in Figure 5-31. The effect of reinforcement ratio on crack widths is also not significant

as shown in Figures 5-32. For the members without initial shrinkage, BSTN3-16 has slightly

greater maximum crack width than BSTN2-16 at all steel stress levels. For the members with

significant initial shrinkage, BSTS2-16 has greater crack width than BSTS3-16 when the steel

stress is less than 200MPa. However, at higher steel stress levels, BSTS2-16 exhibits smaller

crack width than BSTS3-16.


159

500
Steel stress at crack (MPa)

Steel stress at crack (MPa)


500
400 400
300 300
200 200
100 BSTN2-16 BSTN3-16
100
BSTS2-16 BSTS3-16
0 0
0 0.05 0.1 0.15 0.2 0 0.05 0.1 0.15 0.2
Maximum crack width (mm) Maximum crack width (mm)

(a) BSTN2-16 and BSTS2-16 (b) BSTN3-16 and BSTS3-16.


Steel stress at crack (MPa)

500

400

300

200

100 SSTN4-12
SSTS4-12
0
0 0.1 0.2 0.3
Maximum crack width (mm)
(c) SSTN4-12 and SSTS4-12
Figure 5-31: The effects of initial shrinkage on the maximum crack width of flexural
members.

500
Steel stress at crack (MPa)

Steel stress at crack (MPa)

500
400 400
300 300
200 200
100 BSTN2-16 100 BSTS2-16
BSTN3-16 BSTS3-16
0 0
0 0.05 0.1 0.15 0.2 0 0.05 0.1 0.15 0.2
Maximum crack width (mm) Maximum crack width (mm)

(a) BSTN2-16 and BSTN3-16 (b) BSTS2-16 and BSTS3-16.


Figure 5-32: The effects of reinforcing ratio on the maximum crack width of flexural
members.
160

5.6.2 Long-term results (SLTN4-12A and SLTN4-12B)

SLTN4-12A and SLTN4-12B were tested after being cured in the water tank for 15 days. The

shrinkage strains in the specimens before the tests were very small and were ignored in the

analysis. Pre-cast concrete blocks were placed on the top of the specimen to provide sustained

dead loads. The dead loads were transferred as point loads to the specimens via two spandrel

beams at the third span points of the slabs to create constant moment regions (see Figure 5-5

(b)). SLTN4-12A and SLTN4-12B were subjected to two equal point loads of 20.0kN and

15kN, respectively, at the third span points, and the sustained moments in the constant

moment region of each slab were 17.9kNm and 13.9kNm, respectively. The shrinkage strain

and creep coefficient at the end of the long-term loading were -400×10-6 and 1.47.

Increment of mid-span deflection with time

Figure 5-33 and 5-34 show the time-dependent change of mid-span deflections of SLTN4-12A

and SLTN4-12B under sustained service loads. Both slabs did not begin to shrink until the

day of first loading. If the tensile concrete is entirely ignored (fully-cracked), the

instantaneous curvature on each cross-section can be determined using modular ratio theory

and the time-dependent curvature on each cross-section due to creep and shrinkage can be

calculated using the age-adjusted effective modulus method (Appendix B). By integrating the

curvature diagram at any time instant, the mid-span deflection of the fully cracked member

(here called the fully-cracked deflection fc(t) ) can be estimated. The tension stiffening

deflection ts(t) at any time after first loading is obtained by subtracting the measured

deflection of the member at mid-span from the fully-cracked deflection at particular moments.

The results are tabulated in Table 5-13.


161

20
Deflection (mm

15

10

5 SLTN4-12A
first 15days
0
0 10 20 30 40 50 60 70

Time (days)

Figure 5-33: The change of mid-span deflection with time (SLTN4-12A)

15
Deflection (mm

10

5
SLTN4-12B
first 15days
0
0 10 20 30 40 50 60 70
Time (days)

Figure 5-34˖ the change of mid-span deflection with time (SLTN4-12B)


162

Table 5-13: The measured and calculated mid-span deflection for SLTN4-12A and SLTN4-
12B.

Instant behaviour t = 15 days t = 70 days


M
Specimen exp fc ts exp(t) fc(t) ts(t) exp(t) fc(t) ts(t)
kN.m
mm mm mm mm mm mm mm mm mm
SLTN4-12A 17.9 10.86 15.9 5.04 15.93 19.8 3.87 17.91 21.7 3.79

SLTN4-12B 13.9 7.89 12.3 4.33 12.69 15.8 3.11 14.6 17.5 2.9

At first loading, the estimated tension stiffening deflections for SLTN4-12A and SLTN4-12B

were 46% and 55% of the measured total deflections. After 15 days under sustained loading,

the tension stiffening deflection ts(t) decreased by 23% for SLTN4-12A and 28% for SLTN4-

12B. From time 15 days to the end of the test, the tension stiffening deflection decreased by

only a small amount, and at t = 70 days, ts(t) had decrease to 75% and 67% of its

instantaneous value for SLTN4-12A and SLTN4-12B, respectively.

Variation of steel forces:

The measured forces in a steel bar derived from the strain gauges for both slabs are plotted in

Figure 5-35 and the average steel and concrete tensile forces are tabulated in Table 5-14. The

steel stress away from the crack appears to increase relatively quickly after first loading (in

the first 15 days under load) and the concrete tensile force suffers a corresponding reduction,

but in the period t = 15 days to t = 70 days the changes are relatively small. In addition, the

decreasing rate of the average tensile force in the concrete seems independent of the loading

levels.
163

50 50
t = 0 days
Tensile bar force (kN)

t = 15 days
40

Tensile bar force (kN)


40
t = 70 days

30 30
t = 0 days
t = 15 days
20 20
t = 70 days

10 10
1 8 3 10 5 2 9
3 8 2 1 6 9
Crack number
Crack number
0 0
-0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4 -0.4 -0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4
Distance from mid-span (m) Distance from mid-span (m)

(a) SLTN4-12A (b) SLTN4-12B


Figure 5-35: The change of tensile bar forces with time (SLTN4-12A and SLTN4-1B).

Table 5-14: Comparison between fully cracked section and measured steel forces

Slab 3 (17.9kN.m) Slab 4 (13.9kN)


Time
Tcr Ts.ave Tc Tcr Ts.ave Tc
Tc/Ts.cr Tc/Ts.cr
(kN) (kN) (kN) (kN) (kN) (kN)
Instant 42.8 33.3 9.5 0.22 33.3 24.8 8.5 0.26

t = 15 days 43.8 38.1 5.7 0.13 34.1 29.3 4.8 0.14

t = 70 days 44.0 39.4 4.6 0.10 34.2 30.1 4.1 0.12

Crack spacing and crack width:

Figure 5-36 and 5-37 show the crack patterns for SLTN4-12A and SLTN4-12B at the

beginning and the end of the tests. For both SLTN4-12A and SLTN4-12B, 3 new primary

cracks were observed to develop during the first 15 days under load, and the formation of

these new cracks significantly contributed to the drop off in tension stiffening. For the last 55

days under load, no new primary cracks were observed in either slab, and the very small

reduction in tension stiffening in this period may be caused by a gradual deterioration of bond
164

at the concrete-steel interface.

(a) At first loading

(b) After 70 days

Figure 5-36: Crack numbers and locations (SLTN4-12A).

a) At first loading

b) After 70 days

Figure 5-37: Crack numbers and locations (SLTN4-12B).


165

Table 5-15 and 5-16 gives the measured crack widths during the tests. The maximum crack

width had increased by about 70% - 80% during the sustained loading period.

Table 5-15: Crack width at different stages (SLTN4-12A)

Width of Width of Width of Width of Width of Width of Width of Average Maximum


Load
1st crack 2nd crack 3rd crack 5th crack 8th crack 9th crack 10th crack crack width crack width
Stage
(mm) (mm) (mm) (mm) (mm) (mm) (mm) (mm) (mm)
First
0.175 0.175 0.125 0.105 0.1 0.075 - 0.126 0.175
loading
After
0.225 0.3 0.17 0.15 0.125 0.15 0.125 0.178 0.3
70 days

Table 5-16: Crack width at different stages (SLTN4-12B)

Width of Width of Width of Width of Width of Width of Average Maximum


Load
1st crack 2nd crack 3rd crack 6th crack 8th crack 9th crack crack width crack width
Stage
(mm) (mm) (mm) (mm) (mm) (mm) (mm) (mm)
First
0.125 0.125 0.1 0.1 - - 0.112 0.125
loading
After 70
0.175 0.175 0.225 0.15 0.125 0.1 0.158 0.225
days

5.7 SUMMARY
Four beams and four identical slabs were tested to investigate the time-dependent tension

stiffening in short-term and long-term bending. The experimental program was specified in

this Chapter. The tests have demonstrated the significance of drying shrinkage on the

deflection of beams and slabs and have identified the gradual loss of tension stiffening in

flexural members both under short-term and long-term loading.

Similar to the uniaxial tension tests in Chapter 4, it can be concluded that for both short-term

and long-term tests, the loss of tension stiffening shall be directly attributed to the formation

of new cracks and bond deterioration, which can be caused by applied bending moment or

restraint to shrinkage. Short-term test results show that the shrinkage prior to loading can
166

reduce the cracking moment and cause initial deflection. However, early shrinkage appears to

have relatively little effect on tension stiffening and crack width after first cracking.

The long-term tests indicates that tension stiffening can be greatly reduced due to the

formation of new cracks in the first 15 days and then bond deterioration further breaks down

tension stiffening thereafter but at a much slower rate. Since the formation of new cracks is

attributed at least in part to the restraint to shrinkage by the tensile reinforcing bars, the decay

of tension stiffening with time is very much associated with the amount of shrinkage.
CHAPTER 6 FINITE ELEMENT
FORMULATION AND MATERIAL
CONSTITUTIVE LAW

6.1 OVERVIEW
Numerical methods are widely adopted to model concrete structures. Nonlinear finite element

methods have been used for many years to analyse reinforced concrete structures and have

proven to provide good accuracy. Various methods have been developed to model the

interaction between concrete and steel reinforcement and the associated tension stiffening (see

Chapter 2). The difficulties of modeling reinforced concrete structures can be attributed to the

complexities of the material, in particular the development of reliable concrete stress-strain

relationships. Nevertheless, the finite element approach incorporating appropriate material

modeling can generally predict structural behaviour and model the complexities in a manner

and at a level that is not available in other approaches. The complicating factors when using

finite element method to model reinforced concrete structure can be categorized as below:

x The nonlinear stress-strain behaviour of concrete under multiple stress state (including

cracking, tension softening, creep and shrinkage, etc.)

x The development of failure criteria for concrete under multiple stress states.
168

x The interaction between concrete and steel as they work together, namely the bond

mechanism at the concrete-steel interface.

In this study, the finite element program RECAP (Foster 1992, Foster and Gilbert 1990,

Foster and Marti 2003) has been adopted and further developed by the author. RECAP

includes a number of 2-D and 3-D concrete models, together with steel and bond elements.

The major contribution by the author is the development of a 2-D bond interface element that

is demonstrated to provide a good model of bond throughout the serviceability loading range.

The bond model is time-dependent and has been calibrated to accurately model the decay of

tension stiffening with time. The corresponding concrete, steel and bond element formulations

and their material constitutive relationships are introduced in this chapter.

6.2 FOUR-NODE ISOPARAMETRIC CONCRETE


ELEMENT
6.2.1 Finite element formulation

The four-node isoparametric element is often used to model concrete. A typical isoparametric

element is shown in Figure 6-1.


169

Figure 6-1: Four-node plane isoparametric element.

In the element, two coordinate systems are established, namely the global coordinate system

xOy and the local coordinate system O. The values of  and  on the edge of the elements

are [ r1 andK r1 . The global coordinate of any point within the element can be related

to the coordinate of each node by the shape functions. The relationship are given by

4
x ¦N x
i 1
i i 6.1a

4
y ¦N y
i 1
i i 6.1b

Where xi and yi (i = 1, 2, 3, 4) are the nodal coordinates; Ni (i = 1, 2, 3, 4) are the shape

functions of each node and given as

1
N1 1  [ 1 K 6.2a
4

1
N2 1  [ 1 K 6.2b
4
170

1
N3 1 [ 1K 6.2c
4

1
N4 1 [ 1K 6.2d
4

A feature of isoparametric elements is that the shape functions that are used to derive element

coordinates are also the ones that are used to obtain element displacements. The relationship

between interior element displacements and displacements at the nodes can be described as

4
u ¦N u
i 1
i i 6.3a

4
v ¦N v
i 1
i i 6.3b

Where ui and vi (i = 1, 2, 3, 4) are the nodal displacements in x and y directions.

Given the displacements, the strains of any point in the element can be derived as the first

derivatives of the displacements


171

­ wu ½
° °
­H x ½ ° wx °
° ° ° wv °
^H ` ®H y ¾ ® ¾
°J ° ° wy °
¯ xy ¿ ° wu  wv °
° wy wx °
¯ ¿
­u1 ½
° °
­ wN 1 wN 2 wN 3 wN 4 ½° v1 ° 6.4
° 0 0 0 0 °°u 2 °
° wx wx wx wx °° °
° wN 1 wN 2 wN 3 wN 4 °°v 2 °
® 0 0 0 0 ¾® ¾
° wy wy wy wy °°u 3 °
° wN 1 wN 1 wN 2 wN 1 wN 3 wN 3 wN 4 wN 4 °° v °
° wy wx °¿°u °
3
¯ wx wy wx wy wx wy
° 4°
°v °
¯ 4¿
>B @>G @

where [B] is called the strain-displacement matrix, in which the shape functions are

differentiated with respect to global coordinates x and y. Since the shape functions are

represented by local coordinates  and , transformation needs to be made between the two

coordinate systems by the Jacobian matrix, given as

ª wx wy º ª 4 wN i 4
wN i º
« w[ w[ » «¦ w[ xi ¦ w[ yi »
>J @ « » « i 41 i 1
» 6.5
« wx wy » « wN i x wN i »
4

«¬¦ ¦ yi
«¬ wK wK »¼ i 1 wK
i
i 1 wK
»¼

The derivatives of the shape functions about the x and y coordinates can be written as

ª wN i º ­ wN i ½
« wx » ° °
« wN » >J @1 °® wwN[ °¾ 6.6
« i» ° i°
«¬ wy i »¼ °¯ wK °¿

and inputted into Equation 6.4 to obtain the strain-displacement matrix [B].
172

In order to construct the global element stiffness matrix ([Dc]), the material constitutive law

matrix obtained from the material coordinate system ([Dc12]) needs to be transformed by using

a strain transformation matrix [T]

ª c2 s2 cs º
« 2 »
>TH @ « s c 2
 cs » 6.7
« 2cs 2cs c 2  s 2 »
¬ ¼

>Dc @ >TH @T >Dc12 @>TH @ 6.8

in which c cos T and s sin T .  is the angle between the global coordinate system and the

material principal strain direction. The material constitutive matrix [Dc12] in the principal

directions is given by Equation 6.23.

Finally applying the principal of virtual work gives the element stiffness matrix as:

>K e @ t e ³ >B @ >Dc @>B @dA


T
6.9
Ae

where te is the thickness of the element.

6.2.2 Concrete element stress-strain relationship

6.2.2.1 Instantaneous stress-strain relationship

The biaxial concrete strength envelope proposed by Foster and Marti (Foster and Marti 2003),

and shown in Figure 6-2, is used. Other biaxial stress-strain relationships such as that

proposed by Kupfer et al (1969) could also have been used. In Figure 6-2, 1c and 2c are the

principal stresses, and fcp is the uniaxial compressive strength of the in-situ concrete.
173

Figure 6-2: Biaxial strength envelop of Foster and Marti (2003).

For concrete under tension, the bilinear softening curve proposed by Petersson (1981) is used

(Figure 6-3a). Cracking of concrete is introduced in the element as soon as the principal

tensile stress exceeds the tensile strength of the concrete. In order to deal with the mesh

sensitivity issue induced by localization of deformation in the vicinity of a crack in a concrete

element, the “crack band theory” approach (Bazant and Oh 1983) is used to depict the

softening branch of the tensile stress-strain relationship in the fracture zone immediately

adjacent to the crack. According to Petersson (1981), the three parameters required to define

the tensile stress-strain relationship are given in Equation 6.10.


174

(a) Concrete in tension (b) Concrete in compression

Figure 6-3: stress-strain relationship of concrete

1 2 18 E c G f
D1 ; D2 D 3  D1 ; D 3 6.10
3 9 5 f ct2 lt

where Ec is the elastic modulus of concrete; lt is the element size; fct is the tensile strength of

concrete and Gf is the fracture energy of concrete under softening which in this thesis is given

by the CEB-FIP model code (1993):

0.7
§f ·
Gf D f ¨¨ cp ¸¸ , ( f cp :MPa and G f :N/mm) 6.11
© 10 ¹

and f is a coefficient related to the aggregate size and may be taken as:

Df 1.25d max  10 u 10 3 6.12

where dmax is the maximum aggregate size in mm.

In this study, the concrete compressive strength in one principal direction is given by:

f c* E f cp 6.13
175

where  is a scaling factor for the compressive curve in one principal direction and it depends

on the stress in the other principal direction, as shown in Figure 6-3b. The concrete strain at

peak stress is also modified by the same scaling factor , so that

H cp* EH cp 6.14

In the compression-compression state,  is greater than 1.0 and depends on the ratio of major

and minor principal stresses D '


V 1c
V 2c

1.0
E2 for 0 d D ' d 0.48
1D ' 6.15a
2.4

1.15 1  5.5 1
E2 for 0.48  D ' d 1.0 6.15b
1  D ' / 5.5

where 2 is the confining scaling factor for minor principal direction. The confining scaling

factor for major principal direction 1 is given by multiplying ’ with 2

E1 D ' E 2 6.16

In the tension-compression state, the major principal stress is tensile and it reduces the

compressive strength in the minor principal direction (Berlarbi and Hsu 1991; Miyakawa and

Kawakami 1987; Vecchio and Collins 1982). In this study, the equation for the scaling factor

 that was developed by Vecchio and Collins (1986) is adopted

1
E d 1.0
H 6.17
0.8  0.34 1
H cp

where 1 is the principal tensile strain.


176

The uniaxial compressive stress-strain relationship proposed by Thorenfeldt et al.

(Thorenfeldt et al. 1987) is adopted here

nK
Vc  f cp 6.18
n  1  K nk

in which:

Ec Hc
n ;K 6.19
E c  E cp H cp

and cp is the strain corresponding to the peak stress and Ecp is the secant modulus at the peak

f cp
of the curve E cp . The parameter k in Equation 6.18 is the decay factor for post peak
H cp

response and is given by Collins and Porasz (1989) as

k 1 .0 for H c d H cp 6.20a

f cp
k 0.67  t 1.0 for H c ! H cp 6.20b
62

6.2.2.2 Constitutive Relationship for Membranes

For finite element implementation, Foster and Marti (Foster and Marti 2003) used the

equivalent uniaxial strains concept from Darwin and Pecknold (1977) to subtract the

Poisson’s effect in the material constitutive matrix and allow the usage of uniaxial stress-

strain curves. The new formulation can be written as:

­ H 1P ½ 1 ª 1 Q 12 º ­H 1 ½
® ¾ «Q »® ¾ 6.21
¯H 2 P ¿ 1  Q 12Q 21 ¬ 21 1 ¼ ¯H 2 ¿
177

where 1 and 2 are the strains in principal directions, 1 and 2 are the equivalent uniaxial

strains in principal directions, 12 and 21 are the Poisson’s ratios. Therefore the stress-strain

relationship now can be expressed as:

­V c1 ½ ª E c1 0 º ­ H 1P ½
® ¾ « 0 ® ¾
E c 2 »¼ ¯H 2 P ¿
6.22
¯V c 2 ¿ ¬

where Ec1 and Ec2 are the secant moduli in the principal directions. The orthogonal elasticity

matrix in the principal directions is:

ª E c1 Q 12 E c1 0 º
>D@c12 1 «Q E Ec 2 0 »
« 21 c1 » 6.23
1  Q 12Q 21
«¬ 0 0 1  Q 12Q 21 Gc12 »¼

The shear modulus in equation 6.23 is given by Attard et al. (Attard et al. 1996):

1
1  Q 12Q 21 Gc12 >Ec1 1 Q 12  Ec 2 1  Q 21 @ 6.24
4

For cracked concrete elements, both Poisson’s ratios in Equation 6.23 are taken as zero, thus

the equation can be written as:

ª E c1 0 0 º
>D@c12 « 0
« Ec 2 0 »» 6.25
«¬ 0 0 Gc12 »¼

6.2.2.3 Rate type creep model

A rate-type creep law based on the application of Kelvin chains has been incorporated into the

finite element program RECAP by Chong et al. (2008). The rate-type law is based on the

solidification theory that has been reviewed in Section 2.6.2.4. The great advantage of rate-
178

type law over the integral-type law is that it does not require huge computer storage for

tracing the previous stress or strain history in the structure, but only the current states of

variables, hence it greatly reduces the computational effort at each time step.

In this model, the creep of concrete is described by N numbers of Kelvin Chain units. Each

unit  contains a linear spring modulus and a viscosity dashpot (Figure 2-16). The constitutive

equation for this Kelvin Chain can be written as:

N
E P J P  K P J P V, J ¦
P
JP
1
6.26

Where, E P and J P are the elastic moduli and viscosities of unit . Integrating the above

equation can generate the constitutive relation between viscoelastic microstrain  at time t and

the constant stress  that is applied at time t’:

KP 1
V ¦ AP (1  e
( t t ') / t P )
J ), WP and AP 6.27
EP EP

where  is the so called retardation time of the -th Kevin chain unit. Hence the creep

compliance function can be given as:

¦ AP (1  e
( t  t ') / W P )
C (t , t ' ) ) 6.28

In equation 6.28,  must be determined in order to solve the ill-posed problem since different

retardation times can give rise to good fits of the experimental creep curves. Bažant and Xi

(1995) formulated the solidification theory with a continuous retardation spectrum. They also

specified the way to convert this continuous spectrum into a discrete spectrum for numerical

models by simply discretizing the continuous retardation spectrum. A typical discretization of


179

the retardation spectrum is shown in Figure 6-4. A in equation 6.28 is the retardation

spectrum for the discretization of the Kelvin chains and is presented in terms of a continuous

retardation spectrum L( ):

AP L(W P ) Ln10'(logW P ) 6.29

L(W ) «
> @
ª  2n 2 (3W ) 2 n 3 n  1  (3W ) n º (3W ) 3
» q2 
¬« >1  (3W ) n@ 3
¼» 2
6.30
«
> @
ª n(n  2)(3W ) n 3 n  1  (3W ) n  n 2 (3W ) 2 n 3 º (3W ) 3
» q2
«¬ >
1  (3W ) n @
2
»¼ 2

Figure 6-4: Discretization of continuous retardation spectrum of solidifying constituent.

Notice in Figure 6-4, the negative infinite area A0 of the retardation spectrum (Bazant et al.

1997) can be obtained by subtracting equation 6.29 from the microscopic creep compliance

function.
180

6.2.2.4 Finite element implementation for creep and shrinkage

Chong et al. (2008) adopted the exponential algorithm proposed by Bazant (Bazant 1982;

Bazant 1988a), modified the incremental quasi-elastic stress-strain relationship proposed by

Bazant and Prasannan (1989b), and incorporate it into the finite element formulation in the

program RECAP.

The instantaneous concrete stress-strain relationship introduced before is based on a smeared

crack and total strain formulation. Therefore, the total stress of every concrete element is only

related to the total strain at every load step and independent of any previous stress-strain

history. Therefore the creep strain that is calculated from the solidification theory should also

be the total creep strain at any specific time step instant.

By equation 2.15, the viscoelastic creep component is expressed in the form:

'J i 1
'H v t i 1 6.31
Q i 1 / 2

where the subscripts i and i+1/2 indicate the reference to time ti and the time in the middle of

a logarithmic time step ti+1/2, respectively, where

t i 1 / 2 t 0  > t i 1  t 0 t i  t 0 @0.5 6.32

and t0 is the age at first loading.

By equation 6.27, the change in viscoelastic microstrain can be obtained as:

N
'J i 1 ¦J
j 1
j i 1  J ji  G'VA0 6.33

in which G is the biaxial volumetric growth matrix and is given as:


181

ª 1 v 0 º
G « v 1 0 »
« » 6.34
«¬ 0 0 2 1 v »¼

The modified form of the viscoelastic microstrain at time ti+1 for the jth Kelvin chain given by

Chong et al. as:

GV i 1 1 Oj
Jj i 1
J j e  'y i
1  e 'yi  G'V 6.35
i
Ej Ej

where

 'y j
't 1 e
'y j ; Oj ; and 'V V i  V i 1 6.36
Wj 'y j

The volume of the solidified matter at mid-time of a logarithmic time step, vi+1/2, is then given

by:

1
ª 1 q º
vi 1 / 2 «  3» 6.37
¬ t i 1 / 2 q 2 ¼

Then the change in the viscous, non-recoverable, component of creep is evaluated from

equation 2.18 and given as:

'H f t i 1 GV i 1 / 2
6.38
't K i 1 / 2

Where V i 1 / 2 V i 1  'V / 2 . Substituting the apparent macroscopic viscosity, defined as

K i 1 / 2 q 41t i 1 / 2 , into equation 6.38 then the change in the viscous strain per finite element

step is derived as:


182

GV i 1 / 2 q 4 't
'H f t i 1 6.39
t i 1 / 2

Consequently the total creep strain can be obtained by adding both the viscoelastic strain and

viscous strain change components into the total creep strain from the previous time step:

H v t i 1 H v t i  'H v t i 1 6.40a

H f t i 1 H f t i  'H f t i 1 6.40b

In this study, since shrinkage is considered to be independent of stress, the shrinkage strain at

any specific time after the commencement of drying is given by:

Ash t
H sh t 6.41
Bsh  t

Thus, the total time-dependent strain at time t due to creep and shrinkage in the concrete is

given as:

H0 t H v t  H f t  H sh t 6.42

For the sake of finite element implementation, 0(t) is treated as an inelastic strain and then

converted into pre nodal forces P0 applied to the finite element nodes. These equivalent nodal

forces are not computed explicitly in the computational algorithm. However, the pre-strains

0(t) are accounted by being deducted from the total strains before the stresses of the elements

are evaluated. Therefore the pre nodal forces are actually indirectly considered.
183

6.3 TWO-NODE TRUSS ELEMENT FOR


REINFORCEMENT
6.3.1 Finite element formulation

The two node truss element is adopted in this study for modeling the reinforcement used in

this study. Assume a reinforcing bar i-j with a cross section area of A, length of l, shown in

Figure 6-5. The angle between the truss and the global x axis is given as .

The longitudinal strain of the truss can be easily represented as long as the nodal deformations

of both nodes are known:

­ui ½
°v °
H
1
> c  s c s@°® i °¾ 6.43
l °u j °
°¯v j °¿

where c and s are the cosine and sine values of ; u i , vi , u j , v j are the nodal deformation

(Figure 6-5).

Figure 6-5: Two node truss element


184

According to Hooker’s law, the stress in the truss element can be calculated as:

­ui ½
°v °
V EH
E
> c  s c s @°® i °¾ 6.44
l °u j °
°¯v j °¿

Using virtual work theory the nodal force and nodal deformation of the truss element are

related by the stiffness matrix given as:

ª c2 cs c2  cs º
« »
AE « cs s 2
 cs  s 2 »
>K @ 6.45
l « c 2  cs c2 cs »
« »
¬«  cs  s2 cs s 2 ¼»

6.3.2 Stress-strain relationship of the reinforcement

An elastic-plastic stress-strain relationship for the reinforcement is adopted for the finite

element analysis. The relationship is shown in Figure 6-6, in which fsy is the yielding stress

and sy is the corresponding yielding strain. Since this research is only focused on the in-

service load range for concrete structures, the steel bars generally have not yielded and the

stress-strain relationship is assumed to be linear and elastic.


185

Figure 6-6: Stress-strain relationship for the truss element.

6.4 ZERO-WIDTH INTERFACE BOND ELEMENT


6.4.1 Finite element formulation

The zero width interface element is designed to model the bond mechanism between the

concrete and the reinforcing bar in a reinforced concrete structure. An interface element was

firstly introduced by Goodman et al. (1968) to model the joint behaviour in systems of rock

blocks and joints. For reinforced concrete structures, the element layout and its connectivity

with concrete and steel elements are shown in Figure 6-7.


186

Figure 6-7: Zero-width bond interface element.

The relative displacement between node set 1 (consisting of nodes1 and 4) and node set 2

(nodes 2 and 3) represents the slip between the concrete and the steel and is given by

'u ti u ti  u ti and 'vti vti  vti 6.46

where the superscript “+” and “-” denote the upper and lower faces of the interface element,

respectively, and the subscripts t and n represent shear and normal movement, respectively.

The relative nodal displacements ui are linked to the continuous displacement field u by

2
'u ¦ N 'ui i >B@^u e ` 6.47
i 1

Where Ni (i=1, 2) are linear shape functions.

The matrix [B] in equation 6.47 relates the continuous field relative displacements u to the

nodal displacements {ue} along the interface and is given by


187

ª N 1 0  N2 0 N2 0 N1 0º
>B@ « 0 N 1 »¼
6.48
¬  N1 0  N2 0 N2 0

The element stiffness matrix is computed by applying the principal of virtual work and is

given by

>K e @ ³ >B@ >Db @>B@dA


T

6.49
Ae

where Ae is the tangential contact surface area between the interface element and the adjacent

materials and [Db] is the constitutive model relating the bond stress to the relative

displacements between the reinforcing steel and concrete, for example, Ae is the surface area

of the reinforcing bar encased in the concrete and [Db] is given by:

ª E bt 0 º
>Db @ « 0 Ebn »¼
6.50
¬

where Ebt is the bond stress-slip modulus for the current stress state and Ebn is the normal

bond stress-slip modulus. To maintain compatibility between the reinforcing steel and the

concrete in the normal direction, a stiff value for Ebn is used in this study. In the tangential

direction the stiffness is obtained from the relevant constitutive law. Ebt is taken as the secant

stiffness in a modified Newton-Raphson solution process.

For the four-node linear bond element, integration of equation 6.49 is undertaken explicitly

giving:
188

ª2 Ebt 0 Ebt 0  Ebt 0  2 Ebt 0 º


«
«
2 Ebn 0 Ebn 0  Ebn 0  2 Ebn »»
« 2 Ebt 0  2 Ebt 0  Ebt 0 »
« »
cL « 2 Ebn 0  2 Ebn 0  Ebn »
>K e @ 6.51
6 « 2 Ebt 0 Ebt 0 »
« »
« 2 Ebn 0 Ebn »
« sym 2 Ebt 0 »
« »
¬« 2 Ebn ¼»

where c is the sum of the bar circumferences of all bars on a layer and L is the length of the

bond element. The bond element stiffness matrix [Kb] in the global coordinate system is

obtained by:

>K b @ >Tb @T >K e @>Tb @ 6.52

where [Tb] is the diagonal bond-displacement transformation matrix:

>Tb @ >Te Te Te Te @ 6.53

ª cos I sin I º
>Te @ « sin I cos I »¼
6.54
¬

where I is the angle of the interface element to the global x-axis.

6.4.2 Constitutive relationship for bond stress-slip

The bond model must correctly predict the relationship between local bond stress and slip,

and must therefore consider the effects that may influence this relationship (section 2.5).

However most of the pull-out tests on which the available bond-slip relationships have been

calibrated have been undertaken on short anchorage lengths with various boundary conditions,

and cracking and steel stress states are usually ignored.


189

In the following, the CEB-FIP bond model (section 2.5.4) is modified by incorporating the

effects of local damage, primary cracking and concrete confinement pressure for modelling

short-term tension stiffening behaviour under in-service loads (Wu and Gilbert 2009a). At

service load levels, when the steel stress-strain relationship is linear-elastic (i.e. the steel stress

is less than the yield stress), slip normally does not exceed S1 in Figure 6-8. The basic bond-

slip relationship (ignoring cracking and steel stress state) is therefore based on the ascending

part of CEB-FIP bond model, which can be expressed as:

D
§s·
Wb W max ¨¨ ¸¸ 6.55
© s1 ¹

in which max is the maximum bond stress equal to 2.5 f cp if good concrete confinement is

provided; s is the slip, and b is the bond stress.

Figure 6-8: CEB-FIP bond stress-slip relationship.

A number of investigators have proposed to incorporate concrete cracking or damage into the

bond constitutive law (Bolander et al. 1992; Lowes et al. 2004; Soh et al. 2003), by reducing
190

the bond stiffness as a result of damage to the surrounding concrete elements. On the basis of

experimental results, Maekawa et al (2003) proposed a bond model that is related to the local

steel strain.

As shown in Figure 2-2, soon after first cracking, the drop in tension stiffening is substantially

attributed to the formation of primary cracks, where at each crack the bond stress drops to

zero and the slip is substantial. After the formation of the primary cracks (at the crack

stabilization stage), the loss of tension stiffening can be best modeled by reducing the bond

stress with increasing local steel stress (to account for cover controlled cracking). Two

governing parameters 1 and 2 are introduced here to take into account the local concrete

damage at primary cracks and the effect of steel stress on cover-controlled cracks, respectively.

O1 1  D c 6.56

O2 = 1.0 when Vs < 250 MPa;

O2 2.14  0.0044V s when 250 MPa d Vs d 500 MPa; 6.57

O2 = 0.0 when Vs > 500 MPa

In Equation 6.56, Dc is the non-local averaged damage parameter of the concrete elements in

the vicinity of the reinforcement and represents the severity of the cracks. According to the

basic concept of continuum damage theory, the magnitude of damage in part of the concrete

material can be represented by an internal variable Dc. In this study, damage occurs as soon as

concrete cracks in the major principal direction (İ1 > İtp). According to the concrete stress-

strain relationship in tension (Figure 6-3a), concrete is undamaged (Dc = 0) before the stress

reaches its peak value fct, and soon after cracking, damage starts to occur until eventually the

material has lost its integrity (when Dc = 1). Therefore, Dc can be given as
191

Dc 0 when H 1 d H tp ;

V 1c
Dc 1 when H tp  H 1 d D 3H tp ;
Ec ˜ H 1 6.58

Dc 1 when D 3H tp  H 1

In Equation 6.57, s is local steel stress in MPa. It can be seen that the bond stiffness

decreases linearly as 250MPa < s <500MPa, and it diminishes after s >500MPa. The basic

assumption made for this equation is that the bond stiffness is reduced linearly with increasing

steel stress at in-service stage. Equations 6.58 have been calibrated based on the short-term

test results presented in Chapter 4 and 5. Figure 6-9 gives a typical graph of the parameter 2

versus steel stress s within the in-service region for normal-ductility ribbed steel bars.

In Equation 6.55, max and s1 are treated as constants, which are related to the bond

confinement condition. The test data in Marvar (Malvar 1992) shows that the bond strength

envelope increases and its radial deformation decreases significantly with the application of

confinement pressure. This test data is adopted here to alter the peak bond strength max

according to the average confinement stress pc of concrete around the steel bar elements. A

parameter 3 is selected as a variable power of max as follows:

pc
O3 1 6.59
f cp
192

2 1.2

1
O2 2.14  0.0044V s
0.8

0.6

0.4

0.2 Crack formation stage Crack stablised

0
0 100 200 300 400 500
Steel stress ı s

Figure 6-9: Parameter 2 versus steel stress s.

Consequently, the modified bond stress-slip law under short-term monotonic load can be

written as:

D
O3 §s·
Wb O1O2W max ¨¨ ¸¸ 6.60
© s1 ¹

In order to incorporate these parameters into the bond stress-slip relationship in the finite

element program, a non-local analysis is undertaken at the end of each load step. As shown in

Figure 6-10, at each integration point in the bond interface elements, the surrounding concrete

damage value and confinement stress within a radius R is averaged by a weighting function

(r), where r is the distance between the bond element integration point and the source

concrete element integration points within the radius R (as shown in Figure 6-10a). Therefore,

the weighted average damage parameter Dc and confinement stress pc can be written as:
193

Dc ³ Z' r D
V
c r , T dV 6.61a

pc ³ Z' r pc r, T dV 6.61b
V

where the normalized weight function ’(r) is:

Zr
Z' r
6.62
V
³ Z r dV

As indicated in Figure 6-10b, the weight function in this study is taken as:

ª § 2r · 2 º
Zr exp « ¨ ¸ » 6.63
«¬ © R ¹ »¼

§ § 2r · 2 ·
Exp¨  ¨ ¸ ¸
¨ ©R¹ ¸
© ¹

a): Non-local region of bond and concrete b): Weight function of non-local
element integration point analysis

Figure 6-10: Non-local average scheme of bond element.

The weighting function is largest at the nearest integration point and decreases as the distance

r increases. The weighted averaged parameters Dc and pc are then substituted into
194

Equations 6.56 and 6.59 to give the governing parameters 1 and 3. In this study, the concrete

crack effective radius R is taken as 2db.

In order to model the long-term behaviour and mechanism of tension stiffening, the

development of a time-dependent bond model is essential. Through the literature review in

Chapter 2, few experimental programs have been undertaken to investigate the long-term or

time-dependent behaviour of bond.

As mentioned earlier, the time-dependent loss of tension stiffening at crack initiation stage in

a tension member is mainly attributed to the formation of new primary cracks caused by the

restraint provided by the steel bars to shrinkage. At this stage, the steel stress usually is less

than 250MPa and the local deterioration of bond is not severe. However, if the applied load

remains constantly after the primary crack pattern has stablised (usually when 250MPa < s

<400MPa), the bond break-down between the primary cracks with time becomes the

dominant factor causing the loss of tension stiffening. As a result, the time-dependent bond

stress is dependent on both internal and external actions. The breakdown of bond with time is

primary the result of restraint to concrete shrinkage by the reinforcing bar, but is also due to

creep.

In this study, the time-dependent bond model is considered to be associated with drying

shrinkage of the surrounding concrete and the steel stress state. In order to correctly model the

interaction of shrinkage and steel stress on the decrease of bond stress with time, a long-term

shrinkage-related parameter 4 is introduced as another multiplier to modify equation 6.60 so

that for the time-dependent analysis


195

D
O3 §s·
Wb O1O2 O4W max ¨¨ ¸¸ 6.64
© s1 ¹

in which 4 is given as:

k1
§H t ·
O4 1  ¨¨ sh ¸¸ and O 4 t 0.5
© H cs ¹
6.65
§ V ·
k1 3.0¨1  s ¸
© 500 ¹

where H cs is the design shrinkage strain after 30 years that is specified in AS3600 (2009), and

V s is the local tensile steel stress in MPa. In this study, Equation 6.64 is only applied when

the structure is under long-term loading as it only accounts for the long-term loss of bond,

which means in short-term analysis the effect of 4 is not taken into account (4 = 1).

Figure 6-11 shows the shrinkage effects on bond given by Equation 6.65. The proposed

shrinkage-related bond model gives rise to a sequence of curves 4 that depict the bond stress

deterioration at different steel stress levels. The influence of concrete shrinkage on bond

breakdown is less significant at lower steel stresses, as shown in Figure 6-11. The lower limit

of 4 is 0.50, which indicates that ultimately reduction of bond stress-slip stiffness due to

restraint to drying shrinkage is assumed to be 50%. The accuracy of this approximation will

be verified in Chapter 7 by using the proposed bond model to model the experimental work.
196

4 1.2

0.8
200MPa 100MPa
0.6 300MPa
400MPa
0.4

0.2

0
0 100 200 300 400 500
Shrinkage strain (×10-6)

Figure 6-11: Bond breakdown with the amount of shrinkage by the proposed model.

6.5 SOLUTION PROCEDURES


Nonlinear problems in the finite element context usually are solved using iterative algorithms.

The selection of a particular algorithm is important since it affects the computational

efficiency and effectiveness.

In the finite element program RECAP, a variety of solution schemes are available. One of the

most prevailing method is called the Standard Newton Raphson Method (SNRM) is adopted.

In the SNRM, the tangential stiffness matrix [K0] is formed at the beginning of the first

iteration (Figure 6-12), and then the displacement can be approximated by:

>G 1 @ >K 0 @1 ^P` 6.66

The strain and stress in each element is then obtained from the initial displacement and the

constitutive law of the material. Next, the corresponding nodal forces {P1} are formed. The

tangential stiffness matrix [K1] according to the displacement [1] is formed and the
197

incremental displacement [2] under load {P1} = {P} - {P1} are calculated by:

>'G 2 @ >K1 @1 ^'P1 ` 6.67

Thus the approximated displacements at the second iteration are given by:

>G 2 @ >G 1 @  >'G 2 @ 6.68

The foregoing steps are repeated as:

>'Pi @ >P@  >Pi @ 6.69a

>'G i 1 @ >K i @1 ^'Pi ` 6.69b

>G i 1 @ >G i @  >'G i 1 @ 6.69c

until the convergence criteria is satisfied (either [i+1] - [i] or {Pi}) is less than a preset

minimum limit. The solution procedure for a finite element model consists of a number of

loading steps, each of which contains several iterations as described above. For the analysis in

this study, convergence is set at 1 percent for displacement and force norms with a maximum

of 100 iterations for one load step. For the time analysis under constant load, the period of

loading is discretized into a predetermined number of time steps, and the analysis is

performed at each successive time instant.

In this study, instead of using tangent stiffness matrix (Figure 6-12), a secant stiffness matrix

is adopted in Equations 6.66 – 6.69. In the analysis of a structure where softening occurs, the

secant stiffness approach offers better numerical stability than the tangent stiffness approach.

This is because a negative tangent stiffness may lead to numerical instability, while the

change of secant stiffness under softening is still gradual and a positive stiffness is guaranteed.
198

Figure 6-12: One load step in Standard Newton Raphson Method.

6.6 SUMMARY
A 2-dimensional continuum based finite element package RECAP has been modified and used

for the analysis of the experimental work in this thesis. The power and accuracy of the finite

element method has made it one of the most attractive approaches to analyse reinforced

concrete structures, both in engineering practice and as a research tool.

This chapter presents the concrete, steel bar and bond interface elements and their constitutive

relationships that have been written into RECAP. The concrete model is a hyperelastic model

that is based on the total secant modulus at each load step. The cracking of concrete is

demonstrated by using the crack band theory that relates the fracture energy of concrete with

the element size in order to solve the mesh sensitivity issues. Creep is modelled with the rate-

type solidification creep theory and incorporated into the program by using an incremental

stress-strain relationship. Creep and shrinkage strains developing with time are treated as
199

inelastic total strains at each time.

The reinforcement bars are modelled assuming an elastic-plastic stress-strain relationship,

where the stress after yielding remains constant.

The bond mechanism between concrete and reinforcing bar is modelled with the zero-width

interface element. A modified bond model is proposed by the author. The bond stress in this

study is not only related to slip, but also some local parameters (including concrete damage,

steel stress, confinement and shrinkage, etc). A non-local analysis is performed at the

beginning of each load step to obtain relevant local bond parameters from the adjacent

concrete and steel elements.

The Newton Raphson Method is used to solve the nonlinear finite element program. The

solution procedure is also presented herein.

The finite element program has been used to analyse the test specimens described in Chapter

4 and 5, and the results are compared with the experimental measurement in the next chapter.

The program has also been verified by modeling some experimental specimens tested by

others. The proposed model is used to undertake a parametric investigation of the factors that

affect the time-dependent behaviour of cracked reinforced concrete and the magnitude of

tension stiffening.
CHAPTER 7 FINITE ELEMENT ANALYSIS
OF EXPERIMENTAL WORKS

7.1 OVERVIEW
The finite element program presented in the previous chapter has been used to analyse a

number of specimens tested experimentally and the results are presented in this Chapter. The

specimens tested by the author in Chapter 4 and 5 are modeled first. Some other specimens

tested in other studies and reported in the literature have also been analysed (Gribniak 2009,

Nejadi and Gilbert 2004b). The numerical results are then compared with the experimental

results in order to verify the finite element model. Good correlation between numerical and

experimental results shows that the finite element program with the proposed bond model is

capable of accurately capturing time-dependent tension stiffening behaviour in reinforced

concrete structures.

7.2 FINITE ELEMENT ANALYSIS OF UNIXIAL


TENSION PRISMS
7.2.1 Finite element configuration

The finite element mesh designed for the analysis of the uniaxial tension prisms (discussed in

Chapter 4) is shown in Figure 7-1. Symmetry requires only one quarter of each prism to be
201

modeled. An element size of 5 mm by 5mm was adopted, so as to reliably model the local

stress redistribution between adjacent primary cracks. The prisms are modeled using 1332

nodes, 1100 concrete elements, 110 two node steel reinforcement truss elements and 110 zero-

width bond interface elements. The concrete elements are isoparametric quadrilateral plane

stress elements with numerical integration performed using 2 × 2 gauss quadrature. The steel

elements are connected to the concrete elements with overlapping nodes by the zero-width

bond interface elements. The finite element solution is dependent on the mesh size. Relatively

small elements are necessary to accurately model the discrete nature of cracking and bond-slip

at the steel-concrete interface. However, for the specimen sizes considered here, a further

reduction in the element size resulted in relatively little increase in accuracy. In the modeling,

randomised tensile strengths are generated in concrete elements with a 2% maximum

difference being assumed. Therefore, first cracking could be initiated in the concrete element

with the lowest assigned tensile strength and the tensile stress in surrounding elements in the

vicinity of the first crack dropped sharply.

The material properties used as input data for the finite element model were those measured

in the experimental program. Relevant equations can be found in section 6.2.2.1. Table 7-1

gives the major instantaneous material parameters for concrete. Notice that the value of 2 and

3 are dependent on both fracture energy and the finite element size. Other property

parameters are calculated implicitly by the computer program.


202

Figure 7-1: Finite element model configuration for uniaxial tension prisms.

Table 7-1: Parameters for the instantaneous constitutive law of concrete (uniaxial tension
specimens).

Specimens fcp fct Ec 1 2 3


STN12
21.56 2.0 22400 0.333 50 230
STN16
STS12
24.73 2.1 21600 0.333 50 230
STS16
LTN12A
21.56 2.0 22400 0.333 50 230
LTN12B
LTN12C
27.73 2.6 20500 0.333 41 183
LTN12D

In order to obtain the appropriate input data for the solidification creep model, the creep

coefficient is transformed into the compliance function J(t,t’), indicating the strain at age t

caused by a unit constant stress applied at age t’. Figure 7-2 plots the compliance function

curves obtained for this study. The empirical material constants q2, q3 and q4 in Equations
203

6.37 and 6.39 were determined by fitting the compliance data into the curves (Table 7-2). The

Dirichlet series was discretized into seven Kelvin chain units for storing the viscoelastic strain

history. The elastic moduli E of the -th Kelvin chain can be obtained by substituting the

predefined  (retardation time) into Equations 6.29 and 6.30. The negative retardation

spectrum area A0 can be calculated by subtracting the sum of 1/E from the microscopic

creep compliance function. In this study, the retardation time  is taken as 0.0001, 0.001, 0.01,

0.1, 1, 10, and 100 days. The corresponding values of E are given in Table 7-3.

1.4E-04
J(t,t') (MPa-1)

1.2E-04
1.0E-04
8.0E-05
6.0E-05
Experiment result
4.0E-05
2.0E-05 fit in curve

0.0E+00
0 10 20 30 40 50 60
t-t' (days)
(a) For the first batch (STN12, STN16, STS12, STS16, LTN12A and LTN12B)
1.2E-04
J(t,t') (MPa-1)

1.0E-04

8.0E-05

6.0E-05

4.0E-05 Experiment result


2.0E-05 fit in curve
Poly. (fit in curve)
0.0E+00
0 10 20 30 40 50 60
t-t' (days)
(b) For the second batch (LTN12C and LTN12D)
Figure 7-2: Compliance function curves for uniaxial tension tests.
204

Table 7-2: Major creep solidification data for uniaxial tension tests

A0 q2 q3 q4
Specimens
(MPa-1) (μ/MPa) (μ/MPa) (μ/MPa)

First batch 13.2 158 3 70

Second batch 66.7 200 2 35

*First batch: STN12, STN16, STS12, STS16, LTN12A and LTN12B; Second batch: LTN12C
and LTN12D.

Table 7-3: The retardation times  and the correlated E for uniaxial tension tests

Eμ μ
Specimens
(MPa) (days)

0.15348 0.0001
0.10448 0.001
0.072 0.01
First batch 0.05026 0.1
0.03555 1.0
0.02548 10.0
0.01850 100.0

0.160 0.0001
0.109 0.001
0.075 0.01
Second batch 0.052 0.1
0.037 1.0
0.026 10.0
0.019 100.0

The Newton Raphson solution scheme is used in the analysis to capture the nonlinear

behaviour of the structure. Because this study is only concerned with the tension stiffening

effect under in-service conditions, the analysis was terminated when the steel stress reached
205

the yield stress (500MPa), where tension stiffening is assumed to have disappeared.

7.2.2 Finite element results

7.2.2.1 Short-term results

Figure 7-3 shows the numerical and experimental results for the axial force versus average

strain diagram for the short-term tension prisms. The responses obtained from the finite

element model using the proposed bond model and the CEB-FIP bond model (only for STN12

and STN16) are shown in the figure. The average strain is calculated from the finite element

model by dividing the axial deformation between the two monitored nodes A and B, as shown

in Figure 7-1, by the distance between them.

In general, the proposed bond model gives excellent correlation with the experimental

response, especially at the cover-control crack stage. At first cracking (P = Pcr), there is an

abrupt change of stiffness and the stiffness continues to degrade under increasing deformation

as further cracks occur. As the load increases, the tension stiffening strain gradually reduces.

By contrast, the CEB-FIP bond model tends to overestimate tension stiffening under

increasing loads.

To further investigate the difference between proposed bond model and CEB-FIP bond model,

tension stiffening factor t of STN12 and STN16 are calculated. Figure 7-4 shows the tension

stiffening factor versus average strain curves for STN12 and STN16. The tension stiffening

factor predicted using the proposed bond model also shows good correlation with the test

results, with t decreasing as the average strain increases. The results predicted using the

CEB-FIB bond model tend to increase with increasing strain after all the cracks have formed

and this is contrary to the experimental observations. Furthermore, this overestimation of


206

stiffness is more significant in the specimen with the smaller reinforcement ratio (STN12).

The tension stiffening factor obtained from the experiments declines to almost zero as the

bare bar steel stress approaches the yield stress of the steel.
Load (kN)

60

Load (kN)
60
50
50
40 40
30 30
20 Experimental result
Proposed bond model 20
Experimental result
CEB-FIP bond model
10 Bare bar response 10 Proposed bond model
Bare bar response
0 0
0 500 1000 1500 2000 2500 -500 0 500 1000 1500 2000 2500
Average strain (×10-6)
Average strain (×10-6)

(a) STN12 (b) STS12


Load (kN)

120
Load (kN)

120
Experimental result
100 100 Proposed bond model
80 Bare bar response
80
60 60
40 Experimental result
Proposed bond model 40
20 CEB-FIP bond model
Bare bar response 20
0
0
0 500 1000 1500 2000 2500
-500 0 500 1000 1500 2000 2500
Average strain (×10-6)
Average strain (×10-6)

(c) STN16 (d) STS16

Figure 7-3: Comparison between experimental and finite element results for the short-term
uniaxial tension tests (Load versus average strain).
207

Tension stiffening factor ȕt


1.2 Experimental result 1.2
Tension stiffening factor ȕt

Experimental result
1.0 proposed bond model 1.0 Proposed bond model
CEB bond model CEB bond model
0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
0 500 1000 1500 2000 2500 0 500 1000 1500 2000 2500
Average strain (×10-6) Average strain (×10-6)

(a) STN12 (b) STN16

Figure 7-4: Comparison between experimental and finite element results for the short-term
uniaxial tension tests (tension stiffening factor versus average strain).

The significant difference between the proposed and CEB-FIP bond models is caused by the

fact that the latter directly links bond stress to slip which keeps increasing as external load

increases. However, in reality, bond stress is also related to the local concrete damage due to

cracking, which has been considered in the proposed bond model.

The analysis output from the finite element model was converted into GID postprocessor in

order to show the local results in a graphical format. Figure 7-5 and 7-6 show the output from

the GID postprocessor, including tensile strain contours, the variation of steel force, bond

stress and slip distributions along each member at different load levels. The principal tensile

strains at the crack locations are much greater than the strain corresponding to the onset of

cracking (about 0.0001). In between the cracks; all the concrete strains remain in the elastic

range (< 0.0001) and the concrete remains uncracked (intact). Also at the cracks the steel

force reaches its maximum, and mid-way between cracks, it is at a minimum. The

numerically-generated graphs show that, with increasing load, the steel force tends to be more

uniform along the bar (due to the degradation of bond) as the tension stiffening effect
208

decreases. The proposed bond model gives rise to reduced bond stress with increasing slip as

the load is increased within the service load range ( 250MPa < s < 400MPa). However, the

variation of steel force measured during the tests is not smooth (Section 4.6), because the

breakdown in bond is quite local at the location of cover-controlled cracks and this is reflected

by the relatively large differences in strains measured at adjacent strain gauges. Therefore, the

proposed bond model reduces the average bond stress with increasing steel stress without

capturing the very local effect of damage on bond caused by cover-controlled cracking.
209

(a) Tensile strain contour of STN12 (at 50kN) (b) Tensile strain contour of STS12 (at 50kN)

Force (kN)
60
Axial force (kN)

60
50 50
40 40
30 30
20 20
40kN
10 40kN
50kN 10
50kN
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6
Distance from the middle section (m) Distance from the middle section (m

(c) Steel force distribution of STN12 (d) Steel force distribution of STS12

6.0 6.0 40kN


Bond stress (MPa)

4.0 50kN
Bond stress (MPa)

4.0
2.0 2.0

0.0 0.0

-2.0 0 0.1 0.2 0.3 0.4 0.5 0.6 -2.0 0 0.1 0.2 0.3 0.4 0.5 0.6

-4.0 -4.0
40kN
-6.0 50kN -6.0
Distance from the middle section(m)
Distance from the middle section (m)

(e) Bond stress distribution of STN12 (f) Bond stress distribution of STS12
Slip (mm)

0.3 0.3
40kN
0.2 50kN
0.2
Slip (mm)

0.1 0.1

0.0
-0.1 0 0.1 0.2 0.3 0.4 0.5 0.6
0 0.1 0.2 0.3 0.4 0.5 0.6
-0.1
-0.2 40kN
50kN -0.2
-0.3 Distance from the middle section(m)
Distance from the middle section (m)

(g) Slip distribution of STN12 (h) Slip distribution of STS12

Figure 7-5: Finite element results for STN12 and STS12 (local response).
210

(a) Tensile strain contour of STN16 (90kN) (b) Tensile strain contour of STS16 (90kN)
Force (kN)

100

Force (kN)
100

80 80

60 60
76kN 76kN
90kN 90kN
40 40
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6
Distance from the middle section (m)
Distance from the middle section (m)
(c) Steel force distribution of STN16 (d) Steel force distribution of STS16

4.0 76kN 6.0 76kN


90kN 90kN
4.0
Bond stress (MPa)
Bond stress (MPa)

2.0
2.0
0.0 0.0
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6
-2.0
-2.0
-4.0
-4.0 -6.0
Distance from the middle section (m)
Distance from the middle section (m)

(e) Bond stress distribution of STN16 (f) Bond stress distribution of STS16

0.2
0.2 76kN 76kN
90kN 90kN
0.1
0.1
Slip (mm)
Slip (mm)

0.0 0.0
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6
-0.1 -0.1

-0.2 -0.2
Distance from the middle section (m) Distance from the middle section (m)

(g) Slip distribution of STN16 (h) Slip distribution of STS16

Figure 7-6: Finite element results for STN16 and STS16 (local response).
211

7.2.2.2 Long-term results

Figure 7-7 plots the tension stiffening strain versus time curves of the long-term uniaxial

tension tests generated by the finite element models (with both the CEB-FIP bond creep

model and proposed bond model). The experimental results of LTN12A and LTN12C (Figure

7-7 (a) and (b)) indicate that during the long-term loading period, tension stiffening strain

decreases gradually, which can also be accurately demonstrated by the results of the proposed

shrinkage-related bond model. The CEB bond creep model in fact predicts responses that

generally indicate a larger instantaneous tension stiffening strain (as discussed for the short-

term analysis) and an increase of tension stiffening strain with time. The restraint to concrete

shrinkage gradually builds up the tensile stress in the concrete and the shortening of the entire

member, however, it also induce cover-controlled cracks and a breakdown of bond and

consequently member stiffness. The CEB bond creep model only relates the deterioration of

bond stress with creep (time period after first loading) and fails to consider the effects of

shrinkage-induced cover-controlled cracks. Therefore, it gives rise to an unreasonable

response for the uniaxial tension members under long-term sustained tensile loads. The

proposed shrinkage-related bond model takes account of this effect and hence provides the

more reasonable results.

The finite element results of LTN12B and LTN12D are quite different from those of LTN12A

and LTN12C in terms of the development of tension stiffening strain with time. The former

were under relatively small sustained loads that initially produce one or no cracks and the

steel stresses were relatively small. The latter were fully-cracked under sustained loads and

steel stresses were relatively large. The loss of tension stiffening strains in LTN12B and

LTN12D is the result of new crack formed with time due to the restraint to shrinkage by the

reinforcing bar. The dots in Figure 7-7 (c) and (d) indicates the time at which a new crack
212

occurred in the models, and a sharp drop of tension stiffening strain is observed thereafter.

Both the proposed bond model and CEB bond model capture the new crack formation

properly. However, the experimental results show a steadier decrement of tension stiffening

strain compared with the finite element results. The formation of new cracks during the tests

appears to be a more gradual process, while in the finite element analysis an abrupt change

occurs as the new crack form (typical at one numerical step). Nevertheless, in the long-term,

after the development of time-dependent cracking the experimental and finite element results

are in good agreement.

The loss of tension stiffening in the long-term tension members can also be seen in the local

responses shown in Figures 7-8 and 7-9. For LTN12A or LTN12C, no additional cracks are

monitored by the model but clear decreases of local steel stresses and bond stresses are found

(Figure 7-8 (c), (d), (e), and (f)). For LTN12B or LTN12D, one new crack formed during the

sustained loading period (no crack had formed under instantaneous loads), which causes a

significant redistribution of local stresses and loss of tensile stress in the concrete around the

cracks. The finite element model is a valuable research tool to undertake a detailed

investigation of the mechanisms of tension stiffening and consequent reduction of tensile

stresses in the concrete with time.


213

Tension stiffening strain (×10 -6


LTN12A
Tension stiffening strain (×10 -6

700 600
Proposed bond model Experiment result
600 CEB-FIP bond model 500 CEB bond creep model
Proposed bond model
500
400
400
300
300
200
200
100
100
Concrete age (days) Concrete age (days)
0 0
33 43 53 63 73 83 28 43 58 73 88

(a) LTN12A (b) LTN12C


Tension stiffening strain (×10 -6

Tension stiffening strain (×10 -6


1000 900
new crack
800
800 700
600
600 500
400
400 Experimental results Experimental results
300
CEB bond creep model CEB bond creep model
200 Proposed bond model
200 Proposed bond model
100
Concrete age (days) Concrete age (days)
0 0
30 50 70 90 110 28 38 48 58 68 78

(c) LTN12B (d) LTN12D


Figure 7-7: Comparison between experimental and finite element results for the long-term
uniaxial tension tests (tension stiffening strain versus time).
214

(a) Tensile strain contour of LTN12A after 60 days (b) Tensile strain contour of LTN12C after 60 days
Force (kN)

60

Force (kN)
60
50
50
40 40
30 30
20 Start of test 20
Start of test
10 After 60 days 10 After 60 days
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6
Distance from the middle section (m) Distance from the middle section (m)

(c) Steel force distribution of LTN12A (d) Steel force distribution of LTN12C

6.0 Start of test 6.0 Start of test


After 60 days After 60 days
4.0 4.0
Bond stress (MPa)

Bond stress (MPa)

2.0 2.0

0.0 0.0
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6
-2.0 -2.0

-4.0 -4.0

-6.0 -6.0 Distance from the middle section (m


Distance from the middle section (m)

(e) Bond stress distribution of LTN12A (f) Bond stress distribution of LTN12C

0.2 Start of test 0.3


Start of test
After 60days
0.2 After 60 days
0.1
Slip (mm)

0.1
Slip (mm)

0.0
0.0
0 0.1 0.2 0.3 0.4 0.5 0.6
0 0.1 0.2 0.3 0.4 0.5 0.6
-0.1 -0.1

-0.2
-0.2
Distance from the middle section (m) -0.3 Distance from the middle section (m)

(g) Slip distribution of LTN12A (h) Slip distribution of LTN12C

Figure 7-8: Finite element results for LTN12A and LTN12C (local response).
215

(a) Tensile strain contour of LTN12B (after 60 days) (b) Tensile strain contour of LTN12D (after 60 days)

30
Force (kN)

Force (kN)
Start of test 30 Start of test
After 60 days After 60 days
20 20

10 10

0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6
Distance from the middle section (m) Distance from the middle section (m)

(c) Steel force distribution of LTN12B (d) Steel force distribution of LTN12D

6.0 6.0 Start of test


Start of test
4.0 After 60 days After 60 days
4.0
Bond stress (MPa)
Bond stress (MPa)

2.0 2.0

0.0 0.0
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6
-2.0 -2.0

-4.0 -4.0

-6.0 Distance from the middle section (m) -6.0 Distance from the middle section (m)

(e) Bond stress distribution of LTN12B (f) Bond stress distribution of LTN12D

0.2 0.2
Start of test Start of test
After 60days After 60 days
0.1 0.1
Slip (mm)
Slip (mm)

0.0 0.0
0 0.1 0.2 0.3 0.4 0.5 0.6 0 0.1 0.2 0.3 0.4 0.5 0.6

-0.1 -0.1

-0.2 Distance from the middle section (m)


-0.2 Distance from the middle section (m

(g) Slip distribution of LTN12B (h) Slip distribution of LTN12D

Figure 7-9: Finite element results for LTN12B and LTN12D (local response).
216

7.3 FINITE ELEMENT ANALYSIS OF FLEXURAL


SPECIMENS
7.3.1 Finite element configuration

The finite element models for both beams and slabs are depicted in Figure 7-10. Due to

symmetry, only half of the span of each member is modeled. Four-node isoperimetric

elements, two-node truss elements and interface elements are also used to model the concrete,

reinforcing bars and bond interface, respectfully. An element size 10mm by 10mm is adopted,

which leads to a large number of nodes and elements of one model. Larger element size can

be adopted to reduce the number of nodes and elements and improve the computational speed.

However, in this analysis, in order to represent a more discrete-like cracking pattern in the

numerical results, smaller element sizes are chosen. The steel plates at the supports are

modeled by using very stiff linear elastic elements. A vertical roller is placed under the steel

plate to model the boundary condition at the support, while horizontal rollers are placed on

the outer right edge of the model at mid-span. The concrete elements used in the models are

assigned with a randomized tensile strength equal to the tensile strength of concrete ± 2%.

The Newton-Raphson solution scheme is also used for the finite element analysis of the

beams and slabs. For the short-term analysis, small load steps are adopted to capture the

nonlinear performance at the onset of cracking. The analysis is terminated at the load when

the steel is about to yield. For the long-term analysis, the load is maintained constant after

instantaneous loading; and the time analysis is undertaken at each time instant until the end of

the test. The mid-span deflection of the specimens is the vertical displacement at Node A in

Figure 7-10.
217

Figure 7-10: Finite element mesh configuration for flexural tests.

The major parameters of concrete instantaneous constitutive law for the flexural tests are

given in Table 7-4. Figure 7-11 demonstrates the calculated compliance function J(t,t’) in

accordance with the measured creep coefficient and the data used to calculate the numerical

model. The calculated creep solidification parameters A0, q2, q3 and q4 are given in Table 7-5.

The corresponding elastic moduli E P and retardation time W P of the Kelvin chain units are

given in Table 7-6.


218

Table 7-4: Parameters for concrete instantaneous constitutive law (flexural tests)

Specimens fcp fct Ec 1 2 3

BSTN2-16
26.0 2.2 25000 0.333 36.6 163
BSTN3-16
BSTS2-16
40.0 2.4 32000 0.333 41.1 185
BSTS3-16
SSTN4-12 28.6 1.8 25000 0.333 50.0 226.6
SSTS4-12 32.6 2.0 29000 0.333 68.0 320.6
SLTN4-12A
28.6 1.8 25000 0.333 50.0 226.6
SLTN4-12B

Table 7-5: Major creep solidification data for flexural tests.

A0 q2 q3 q4
Specimens -1
(MPa ) (μ/MPa) (μ/MPa) (μ/MPa)

Beams 21.3 100 1.5 25

Slabs 50.2 100 1.5 35

Table 7-6: The retardation times  and the correlated E of the Kevin chain units for flexural
tests.

Eμ μ Eμ μ
Specimens Specimens
(MPa) (days) (MPa) (days)
0.21772 0.0001 0.34758 0.0001
0.15734 0.001 0.23601 0.001
0.10813 0.01 0.16219 0.01
Beams 0.07526 0.1 Slabs 0.11289 0.1
0.05308 1.0 0.07963 1.0
0.03794 10.0 0.05691 10.0
0.02748 100.0 0.04122 100.0
219

1.0E-04

J(t,t') (MPa-1) 8.0E-05

6.0E-05

4.0E-05

Experiment result
2.0E-05
fit in curve
0.0E+00
0 10 20 30 40
t-t' (days)

(a) Beams

1.4E-04
J(t,t') (MPa-1)

1.2E-04
1.0E-04
8.0E-05
6.0E-05
Experiment result
4.0E-05
2.0E-05 fit in curve
0.0E+00
0 20 40 60 80 100 120
t-t' (days)
(a) Slabs

Figure 7-11: Compliance function curves for flexural tests.

7.3.2 Finite element results

7.3.2.1 Short-term results

Figure 7-12 illustrates the moment versus mid-span deflection diagrams for the short-term

flexural members. The figure compares the results from the tests and the finite element

models using the proposed bond model, with excellent agreement observed for each specimen.

Since the fully-cracked responses are the same for both experimental and finite element

results, the effect of tension stiffening observed in the experimental results can be accurately
220

simulated by the finite element results. In addition, the finite element model is capable of

predicting the initial deflections before loading due to the gradual shrinkage warping of the

member, and also the decreased cracking moment caused by restraint to initial shrinkage

(Figure 7-12 (b), (d) and (f)). In general, the model accurately simulates the tension stiffening

behaviour in a global sense.

Locally, as can be seen in Figure 7-13, 7-14 and 7-15, the models provide an insight into the

mechanisms of tension stiffening for the short-term flexural members. Similar to the

responses for the uniaxial tension members (Figure 7-5 and 7-6), the tensile strain contours

illustrate the cracking patterns of the flexural members. At cracked sections, the steel forces

are at maximum, and then drop to a minimum in-between the cracks, as shown in the steel

force distribution diagrams. Before loading, the restraint to shrinkage by the reinforcing bar

results in tensile force in the concrete and compressive force in the steel bar. Therefore, the

steel forces near the left support where the moment is low in the pre-shrunk members

(BSTS2-16, BSTS3-16 and SSTS4-12) are negative. This tensile force in the concrete due to

restraint to shrinkage, working together with the applied moment, may be large enough to

cause additional primary cracks. The distributions of bond stress and slip are also affected by

the formation of new cracks under increasing loads, as shown in the figures. Slip reaches its

peak and bond stress drops to zero at the cracks. As the moments keep increasing and no more

cracks are formed, slip continues to increase due to the widening of existing cracks. Whereas

the bond stress in-between the cracks seems not to vary too much, which can be indicated by

the fact that the gradient of local steel force does not change too much as well.
221

Moment (kN.m)

Moment (kN.m)
60 60
C Experimental result
50 C 50
Fully-cracked
FE result
40 40

30 30

20 Experimental result 20
FE result
10 10
Fully-cracked response
0 0
0 1 2 3 4 5 6 0 1 2 3 4 5 6
Mid-span deflection (mm) Mid-span deflection (mm)

a) BSTN2-16 (b) BSTS2-16


Moment (kN.m)

70 Moment (kN.m) 70

60 60

50 50

40
40
30
Experimental result 30
20 Experimental result
FE result 20
FE result
10
Fully-cracked response Fully-cracked response
10
0 O
0 1 2 3 4 5 0
Mid-span deflection (mm) 0 1 2 3 4 5 6
Mid-span deflection (mm)

(c) BSTN3-16 (d) BSTS3-16


Moment (kN.m)

25 25
Moment (kN.m)

20 20

15 15

10 10
Experimental result
FE result Experimental result
5 5 FE result
Fully-cracked response
Fully-cracked response
0 0
0 3 6 9 12 15 18 0 3 6 9 12 15 18
Mid-span deflection (mm) Mid-span deflection (mm)

(e) SSTN4-12 (f) SSTS4-12


Figure 7-12: Comparison between experimental and finite element results for the shot-term
flexural tests (moment versus mid-span deflection).
222

(a) Tensile strain contour of BSTN2-16 at 40kN.m (b) Tensile strain contour of BSTS2-16 at 40kN.m
Tensile bar force (kN)

80
crack crack crack crack crack

Tensile bar force (kN)


at 28.8kN.m at 28.8kN.m crack crack crack crack crack
70 at40kN.m
at 40kN.m at 51.2kN.m
60 at 51.2kN.m
50

40
30

20
10

0
-10
0.1 0.3 0.5 0.7 0.9 1.1 1.3
Distance from left support to mid span (m) 0.1 0.3 0.5 0.7 0.9 1.1 1.3
Distance from left support to mid span (m)

(c) Steel force distribution of BSTN2-16 (d) Steel force distribution of BSTS2-16
6 6 at 28.8kN.m
at 28.8kN.m
at 40kN.m
at 40kN.m at 51.2kN.m
4 4
at51.2kN.m
Bond stress (Mpa)

Bond stress (Mpa)

2 2

0 0
0.1 0.3 0.5 0.7 0.9 1.1 1.3 0.1 0.3 0.5 0.7 0.9 1.1 1.3
-2 -2

-4 -4

-6 -6
Distance from left support to mid span (m) Distance from left support to mid span (m)

(e) Bond stress distribution of BSTN2-16 (f) Bond stress distribution of BSTS2-16
0.2 0.2 at 28.8kN.m
at 28.8kN.m
at 40kN.m at 40kN.m
at 51.2kN.m at 51.2kN.m
0.1 0.1
Slip (mm)

Slip (mm)

0.0 0.0
0.1 0.3 0.5 0.7 0.9 1.1 1.3 0.1 0.3 0.5 0.7 0.9 1.1 1.3

-0.1 -0.1

-0.2 -0.2
Distance from left support to mid span (m)
Distance from left support to mid span (m)

(g) Slip distribution of BSTN2-16 (h) Slip distribution of BSTS2-16

Figure 7-13: Finite element results for BSTN2-16 and BSTS2-16 (local response).
223

(a) Tensile strain contour of BSTN3-16 at 60.8kN.m (b) Tensile strain contour of BSTS3-16 at 60.8kN.m
70

Tensile bar force (kN)


at 40kN.m crack crack crack crack crack at 40kN.m crack crack crack crack crack
60 at 51.2kN.m 70 at51.2kN.m
at 60.8kN.m
Tensile bar force (kN)

at 60.8kN.m
50
50
40

30 30
20
10
10

0
-10
0.1 0.3 0.5 0.7 0.9 1.1 1.3
Distance from left support to mid span (m) 0.1 0.3 0.5 0.7 0.9 1.1 1.3
Distance from left support to mid span (m)

(c) Steel force distribution of BSTN3-16 (d) Steel force distribution of BSTS3-16
6 at 40kN.m 6
at 40kN.m
at 51.2kN.m at 51.2kN.m
at 60.8kN.m 4 at 60.8kN.m
4
Bond stress (Mpa)
Bond stress (Mpa)

2
2

0
0
0.1 0.3 0.5 0.7 0.9 1.1 1.3
0.1 0.3 0.5 0.7 0.9 1.1 1.3
-2
-2
-4
-4
-6 Distance from left support to mid span (m)
Distance from left support to mid span (m)
-6

(e) Bond stress distribution of BSTN3-16 (f) Bond stress distribution of BSTS3-16
0.2 0.2 at 40kN.m
at 40kN.m
at 51.2kN.m at 51.2kN.m
at 60kN.m
at 60.8kN.m
0.1 0.1
Slip (mm)

Slip (mm)

0.0 0.0
0.1 0.3 0.5 0.7 0.9 1.1 1.3 0.1 0.3 0.5 0.7 0.9 1.1 1.3

-0.1
-0.1

Distance from left suppport to mid span (m) -0.2


-0.2 Distance from left support to mid span (m)

(g) Slip distribution of BSTN3-16 (h) Slip distribution of BSTS3-16

Figure 7-14: Finite element results for LTN12B and LTN12D (local response).
224

(a) Tensile strain contour of SSTN4-12 at 20kN.m (b) Tensile strain contour of SSTS4-12 at 18kN.m
60 60
at 12kN.m at 12kN.m
crack crack crack crack crack crack crack crack crack crack
at 16kN.m at 16kN.m
50
Tensile bar force (kN)

50 at 20kN.m at 20kN.m

Tensile bar force (kN)


40
40
30
30
20
20
10

10 0

0 -10
0.1 0.3 0.5 0.7 0.9 1.1 1.3
0.1 0.3 0.5 0.7 0.9 1.1 1.3
Distance from left support to mid span (m)
Distance from left support to mid span (m)
(c) Steel force distribution of SSTN4-12 (d) Steel force distribution of SSTS4-12
6 at 12kN.m 6 at 12kN.m
at 16kN.m at 16kN.m
at 20kN.m at 20kN.m
4 4
Bond stress (Mpa)

Bond stress (Mpa)

2 2

0 0
0.1 0.3 0.5 0.7 0.9 1.1 1.3 0.1 0.3 0.5 0.7 0.9 1.1 1.3
-2 -2

-4 -4

-6
Distance from left support to mid span (m) -6
Distance from left support to mid span (m)

(e) Bond stress distribution of SSTN4-12 (f) Bond stress distribution of SSTS4-12
0.2 at 12kN.m 0.2 at 12kN.m
at 16kN.m at 16kN.m
at 20kN.m at 20kN.m

0.1 0.1
Slip (mm)

Slip (mm)

0.0
0.0
0.1 0.3 0.5 0.7 0.9 1.1 1.3
0.1 0.3 0.5 0.7 0.9 1.1 1.3

-0.1
-0.1

-0.2
Distance from left support to mid span (m) -0.2
Distance from left support to mid span (m)

(g) Slip distribution of SSTN4-12 (h) Slip distribution of SSTS4-12


Figure 7-15: Finite element results for SSTN4-12 and SSTS4-12 (local response).
225

7.3.2.2 Long-term results

Figure 7-16 plots the mid-span deflection versus time results of SLTN4-12A and SLTN4-12B

from the experiment and finite element models. Generally the finite element model shows

good ability to capture the test responses that the mid-span deflection increases significantly

in the first 15 days under loads, which implicitly means that the change of tension stiffening

deflection can also be well simulated by the models.

Figure 7-17 shows the development of the crack patterns of the slabs in terms of tensile strain

contour derived by the finite element models. It can be seen that in the first 15 days, one

additional crack is formed in the half length of both SLTN4-12A and SLTN4-12B (Figure 7-

17 (a) and (b)), which is deemed to be the major factor for the loss of stiffness. Thereafter, the

models give no more cracks till the end of the test and the loss of tension stiffening with time

can be mainly attributed to the deterioration of local bond stresses.


Deflection (mm)

20
Deflection (mm)

15
15
10
10
first 15days Experimental result
5 5 first 15days Experimental result
FE result
FE result
0 0
0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70
Time after first loading (days) Time after first loading (days

(a) SLTN4-12A (b) SLTN4-12B

Figure 7-16: Comparison between experimental and finite element results for the long-term
flexural tests (mid-span deflection versus time).
226

(a) Tensile strain contour of SLTN4-12A after load (b) Tensile strain contour of SLTN4-12B

(c) Tensile strain contour of SLTN4-12A after 15 (d)Tensile strain contour of SLTN4-12B after 15
days days

Figure 7-17: Finite element results for SLTN4-12A and SLTN4-12B (local response).

So far, the finite element program, with the proposed bond model, introduced in Chapter 6, is

able to accurately predict time-dependent tension stiffening in both macro and micro scales.

The model appropriately takes account of the effect of shrinkage and creep in the concrete,

together with the basic mechanisms (cracking and bond deterioration) that cause the loss of

tension stiffening. In order to further verify that the finite element model can be applied in a

universal structural engineering context to predict time-dependent tension stiffening in

reinforced concrete structures, the experimental works conducted by Nejadi and Gilbert

(2004b), and Gribniak (2009) were used to be analysed.

7.4 NEJADI AND GILBERT (2004)


Nejadi and Gilbert (2004b) constructed 6 beams and 6 slabs and tested them under sustained

long-term loads, in order to investigate the effects of shrinkage on crack width. Herein, the

author uses the experimental data to investigate the change of tension stiffening with time in

their tested slabs and to further verify the proposed shrinkage-related bond model.
227

Figure 7-18: Details of cross section for the slabs (Nejadi and Gilbert 2004).

Figure 7-18 shows the cross-sectional dimensions of the slabs constructed by Nejadi and

Gilbert. The slabs were simply supported over a span of 3.5m and each subjected to a constant

sustained distributed load q over the entire span. Table 7-7 tabulates the design details for

each slab. The specimens were all cast on the same day from the same batch of concrete and

moist cured for 14 days prior to the tests. The sustained loading commenced on the 14th day

and then the loads were held constant for 394 days. The constant loads were created by

placing pre-cast concrete blocks on the top of the slabs, together with the self-weight. The

shrinkage and creep characteristics of the concrete were also monitored during the tests, as

well as the instantaneous concrete and steel material properties. The measured material

properties were fully documented by Nejadi and Gilbert (2004). During the test, dial gauges

were used to monitor the mid-span deflections of the slabs, and the measured results are used

here to derive tension stiffening deflections.


228

Table 7-7: Details of slab specimens (bar diameter, db = 12 mm).

Slab No. UDL, cb cs s Mcr Ms fs Mu M s/


of q mm mm mm kNm kNm MPa kNm Mu
bars kN/m (%)
S1-a 2 2.9 25 40 308 4.64 6.81 252 13.6 50.1
S1-b 2 1.9 25 40 308 4.64 5.28 195 13.6 38.9
S2-a 3 4.9 25 40 154 4.74 9.87 247 19.8 49.7
S2-b 3 2.9 25 40 154 4.74 6.81 171 19.8 34.3
S3-a 4 5.8 25 40 103 4.84 11.35 216 25.7 43.7
S3-b 4 3.9 25 40 103 4.84 8.34 159 25.7 32.4
* Mu is the predicted ultimate failure moment; fs is the steel stress at the mid-span.

7.4.1 Finite element configuration

The finite element mesh used for the slabs of Nejadi and Gilbert were similar with that shown

in Figure 7-10, with only half length of each slab being modeled due to its symmetry. Each

slab model is composed of 2832 isoparametric concrete elements (10mm by 10mm size), 177

steel truss elements and 177 bond interface elements. The uniformly constant distributed loads

were applied as nodal loads at every node on the top of the slabs.

The instantaneous properties for concrete are given as

fcp = 24.8MPa; fct = 2MPa; Ec = 24950MPa; v = 0.2; Es = 200000MPa; fsy = 500MPa; and Gf =
75N/m.

and the creep and shrinkage data is given in Table 7-8.

Table 7-8: Creep coefficient and shrinkage strain (x10-6) (Nejadi and Gilbert 2004).

Age 14 16 21 27 53 96 136 200 242 332 394


(days)
Mc 0 0.14 0.36 0.48 0.92 1.15 1.29 1.40 1.50 1.64 1.71

Hsh 0 14 109 179 403 591 731 772 784 816 825
229

According to Table 7-8, with first loading and the onset of drying both at 14 days, creep

coefficient c and shrinkage sh strain gradually increased with time. By the end of the tests,

the measured c and sh are 1.71 and 825 x10-6 respectively. For the finite element model,

parameters Ash and Bsh are taken as 950 and 45days for the shrinkage model in Equation

6.41. The corresponding solidification creep parameters are calculated as q2 = 186.5/MPa, q3

= 1.0/MPa, and q4 = 23.67/MPa. The negative infinite retardation area is taken as A0 =

52.8MPa-1. Eight Kelvin chain units were used and the calculated retardation time  and

elastic modulus E for each Kelvin chain are given in Table 7-9.

Table 7-9: Kelvin chain properties for solidification creep model for Nejadi and Gilbert’s
slabs.

E (MPa) (Days)

0.0848 0.0001;
0.0721 0.001
0.0621 0.01
0.0541 0.1
0.0478 1
0.0428 10
0.0388 100
0.0356 1000

7.4.2 Finite element results

Figure 7-19 shows the numerical and experimental mid-span versus time responses of s1-a

and s2-a in Nejadi and Gilbert’s test. The finite element model results are in good agreement

with the measured experimental results as shown. It appears that the mid-span deflections of

the slabs increased significantly within the first 50 days then tended to be steadier thereafter.
230

In the long-term, the mid-span deflections of the slabs increased by up to 150% of the

instantaneous values. Shrinkage warping was concerned as the major contribution to the time-

dependent development of deformations.


Deflection (mm)

35
35

Deflection (mm)
30
30
25
25
20
20
15
15
10 Experimental result Experimental result
10
5 FE result FE result
5
0 0
0 50 100 150 200 250 300 350 400 0 50 100 150 200 250 300 350 400
Time after first loading (days) Time after first loading (days)

(a) s1-a (b) s2-a

Figure 7-19: mid-span deflection versus time for s1-a and s2-a.

In order to analyse the contribution of tensile concrete to the stiffness of the slabs and how it

changes with time, the fully-cracked deflections at mid-span were calculated in 15, 50, and

394 days after first loading. The relaxation method presented in Appendix B was used to

calculate the responses of the fully-cracked sections and the fully-cracked deflections of the

slabs with time. Then the tension stiffening deflection at different loading stages can be

obtained by subtracting the actual mid-span deflections from the fully-cracked deflections.

Table 7-10 summarises the calculated tension stiffening deflections from both experimental

and numerical results at different stages of the loading history for the slabs considered here.

It is seen in Table 7-10 that for most of the slabs tension stiffening deflections dropped

dramatically in the first 15 days after instantaneous loading by 30% - 50%. Thereafter, the

decrease of tension stiffening deflections slowed down. This decrease of tension stiffening is

quite similar to the observation of the slabs tested by the author, although the rate of decrease

of tension stiffening was slightly different. The total tension stiffening deflections over
231

sustained loading period also varied among the slabs. For the slabs under consideration, the

loss of tension stiffening deflections by the end of the test varied from 46% - 94%. The major

mechanism to cause the loss of tension stiffening was shrinkage-induced cracks developing

with time.

Table 7-10: Calculated tension stiffening deflections for Nejadi and Gilbert’s slabs.

Instant t = 15 days t = 50 days t = 394 days


M
Specimen ts mm
kN.m
EXP FEM EXP FEM EXP FEM EXP FEM
s1-a 6.81 9.26 9.16 5.68 6.30 5.84 4.58 4.79 3.22

s1-b 5.28 9.98 10.11 6.99 7.50 7.49 6.98 5.34 5.48

s2-a 9.87 5.25 5.07 2.25 1.91 2.34 2.41 0.3 1.44

s2-b 6.81 7.34 7.03 4.95 4.87 5.26 3.90 3.2 1.97

s3-a 11.35 4.94 4.75 1.71 2.39 2.06 1.93 1.36 0.88

s3-b 8.34 6.45 6.46 4.00 3.31 4.58 2.70 2.58 2.00

*M is the moment at mid span.

For the slabs subjected to higher applied moments (s1-a, s2-a and s3-a), the decrement of

tension stiffening deflections with time seems more significant than the slabs subjected to

lower applied moments (s1-b, s2-b, and s3-b). For instance, the experimental tension

stiffening deflection of s3-a decreased by 65.4% after 15 days and by 72.5% at the end of the

test, while the corresponding changes for s3-b were 38.0% after 15 days and 60.0% at the end.

Notice this phenomenon was not quite remarkable in the tests undertaken by the author. A

comparison between the results for s1-a (Ast = 220mm2) and s2-b (Ast = 330mm2) shows that

under the same loading level, the reinforcing ratio has little effects on the rate of change of
232

tension stiffening deflections. Tension stiffening deflection of s1-a decreased by 38.7% after

15 days and 48.3% at the end, while that of s2-b decreased by 32.6% after 15 days and 56.4%

at the end.

The crack patterns of s1-a and s2-a at instantaneous loads, after 50 and 394 days are indicated

by the tensile strain contours in Figure 7-20. The graphs indicate that new primary cracks

were formed during the sustained loading period. For the first 50 days under loading, there

were 3 and 2 new cracks discovered in the numerical models of s1-a and s2-a, respectively,

and then after 394 days, 1 more crack had developed in each specimen. Apparently the figure

shows that the time-dependent new cracks were normally formed at lower-moment regions at

some distance away from the existing cracks. Restraint to shrinkage is regarded as the major

factor triggering the time-dependent cracking and causing the loss of tension stiffening.

The deterioration of bond stress, in addition to shrinkage-induced primary cracking, must be

considered as another factor contributing to the loss of tension stiffening in the slabs. Derived

from the finite element results, the bond stress distributions at different loading stages (after

50 days and 394 days) along the specimens for s1-a and s2-a are shown in Figure 7-21. The

proposed shrinkage-related bond model gives rise to a gradual time-dependent reduction of

bond stresses in-between the existing cracks in the slabs. Regardless of the influence of new

cracks on bond, the maximum bond stress in-between the existing cracks in s1-a dropped

from 3.48MPa to 3.01MPa from 50 days after first loading to 394 days after first loading,

while for s2-a, it dropped from 3.52MPa to 3.11MPa during the same time interval.
233

At instantaneous loading At instantaneous loading

After 50 days After 50 days

After 394 days After 394 days

(a) s1-a (b) s2-a


Figure 7-20: Tensile strain contours for s1-a and s2-a.

6 After 50 days 6 After 50 days


After 394 days After 394 days
4 4
Bond stress (Mpa)

Bond stress (Mpa)

2 2

0 0
0.3 0.5 0.7 0.9 1.1 1.3 1.5 1.7 0.3 0.5 0.7 0.9 1.1 1.3 1.5 1.7
-2 -2

-4 -4

-6 -6
Distance from left support to mid span (m) Distance from left support to mid span

(a) s1-a (b) s2-a

Figure 7-21: Bond stress distributions for s1-a and s2-a.


234

7.5 GRIBNIAK (2009)


Gribniak (2009) undertook an experimental program to investigate the effects of shrinkage on

tension stiffening and short-term deformations of lightly-reinforced members. Eight beams

(four couples of twin specimens), each with a single span of 3m, were tested under the four-

point loading. The design dimensions of the beams are shown in Figure 7-22.

Figure 7-22: Longitudinal (a) and cross-sectional (b) reinforcement of the test beams;
notation of cross-section (c) (Gribniak 2009).

The experimental program comprised of two series of beams. In the first series (S-1 and S-2),

four N10 bars were used as the tensile reinforcement, whereas in the second series (S-3 and S-

4), two N14 bars were used. Within the same series, a twin specimen designated with extra

‘R’ was produced. ‘R’ means the twin specimen that had heavy top reinforcement of three

N18 bars (instead of two N6 bars). Heavy compressive reinforcement is provided in an

attempt to eliminate the shrinkage curvature on the specimens. However the longitudinal

restraint to shrinkage within the specimen will be significantly greater. Table 7-11 summaries

the details of the tested beam specimens.

Free shrinkage measurements were performed on prisms of 100 × 100 × 400mm and 280 ×

300 × 350mm in size. The latter was in fact the fragments of the beams, so that the size effects
235

on shrinkage measurements can be eliminated in this situation. Steel gauge studs, with the

base 200mm, were glued or embedded in the concrete to measure the companion free

shrinkage strains.

Table 7-11: Details of beam specimens (Gribniak 2009).

Series Beam D b dst dsc Ast Asc Age fcp


2
mm mm days MPa
S-1 300 280 276 23 309 56.6 48 47.3
1 S1-R 299 281 276 28 309 749 47 47.3
S-2 300 280 282 25 309 56.6 28 48.7
S2-R 300 282 279 31 309 749 29 48.2
S-3 300 277 277 26 303 56.6 31 41.1
2 S3-R 299 281 278 28 303 749 32 41.2
S-4 300 277 274 26 303 56.6 32 54.2
S4-R 301 283 281 29 303 749 34 54.2

7.5.1 Finite element configuration

The finite element configuration of the beams is similar to that shown in Figure 7-10, whereas

in this case not only tensile reinforcement but compressive and transverse reinforcement are

also provided. The tensile steel truss elements are connected to the concrete elements with the

bond interface elements, while the compressive and transverse steel truss elements are

embedded to the concrete elements by sharing the same nodes. The tensile strength and the

elastic modulus of the concrete were calculated from the measured compressive strength

using the formulas specified in AS3600 (2009) and given in Table 7-12.

The other instantaneous material properties for concrete and steel bars are given as:

 = 0.2; Es = 212000MPa; fsy = 566MPa (N10) and 542MPa (N14); and Gf = 92.5N/m.
236

Table 7-12: Calculated tensile strength and elastic modulus of the concrete at the tests.

Series Beam fcp fct Ec


MPa
S-1 47.3 2.75 34770
1 S1-R 47.3 2.75 34770
S-2 48.7 2.79 35282
S2-R 48.2 2.78 35100
S-3 41.1 2.56 32412
2 S3-R 41.2 2.57 32451
S-4 54.2 2.94 37221
S4-R 54.2 2.94 37221

In the experiment, free shrinkage measurements were initiated 24 hours after casting or 3-4

days after casting. Table 7-13 gives the shrinkage deformations of the 280 × 300 × 350mm

prisms at the tests of each beam and the corresponding shrinkage parameters (Ash and Bsh)

used for the model. As for creep, creep strains were also measured in companion tests on 100

× 100 × 400 mm prisms at a stress level corresponding to 40% of the strength. The

measurement of creep strain was initiated on the 7th day after casting. Figure 7-23 illustrates

the measured creep coefficient versus time. The author used this creep coefficient data and

converted it to the compliance function to derive the solidification creep parameters for the

finite element model. The parameters are taken as: q2 = 75/MPa, q3 = 1.2/MPa, and q4 =

15/MPa. The negative infinite retardation area of the solidification spectrum is taken as A0 =

20.0MPa-1. Ten Kelvin chain units were used and the calculated retardation time  and elastic

modulus E for each Kelvin chain are given in Table 7-14.


237

Figure 7-23: Set-up of creep tests and variation of creep coefficient with time
(Gribniak 2009).

Table 7-13: Shrinkage deformations of 280 × 300 × 350mm prisms (Gribniak 2009) and the
corresponding modeling parameters.

Beam t0 (days) Shrinkage Ash Bsh


strain
S-1 4 1,946×10–4 1000 186
S-1R 4 1,882×10–4 1000 194
S-2 3 1,526×10–4 1000 133
S2-R 3 1,557×10–4 1000 146
S-3 4 1,370×10–4 1000 170
S3-R 4 1,396×10–4 1000 173
S-4 4 1,720×10–4 1000 136
S-4R 4 1,770×10–4 1000 136
t0 is the time when free shrinkage measurement commenced.
238

Table 7-14: Kelvin chain properties for solidification creep model for Gribniak’s beams.

E (MPa) (Days)

0.46344 0.0001;
0.31468 0.001
0.21626 0.01
0.15053 0.1
0.10617 1
0.07589 10
0.05495 100

7.5.2 Finite element results

The experimental and numerical mid-span deflection versus mid-span moment responses of

two of the beams are shown in Figure 7-24. Excellent agreement is achieved using the finite

element model. As it is known that with free shrinkage prior to loading, initial deflections

were induced. This initial deflection is very dependent on the amount of shrinkage and the

number of tensile and compressive steel bars. As the beam sections remained uncracked

before loading, shrinkage in the sections with larger compressive reinforcement can induce a

reversed curvature compared with those with little compressive reinforcement, thus the beams

with designation R (heavy compressive reinforcement) might even exhibit upwards

deflections (negative) before loading, even though this initial deflections could be very small

and may be eliminated due to self weight of the specimens (Figure 7-24b). In the test,

shrinkage induced initial curvature or deflection was not measured. Therefore, the initial

negative deflections derived from the finite element models were added to the experimentally

measured deflections to give the more accurate responses.


239

On the fully-cracked sections, since the tensile concrete is ignored, the amount of

compressive reinforcement has little effect on the shrinkage-induced curvature or deflection.

By using the method presented in Appendix B, the initial mid-span deflections of each beam

assumed fully-cracked are calculated and the fully-cracked responses are also plotted in

Figure 7-24.

50 Experimental response
50 Experimental response
FE response
Fully-cracked response FE response
40
Moment (kN.m)

40 Fully-cracked response

30 Moment (kN.m) 30

20 20

10 10
Mid-span deflection (mm) Mid-span deflection (mm)
0 0
0 2 4 6 8 10 12 -1 1 3 5 7 9 11

(a) s-1 (b) s-1R

Figure 7-24: Moment versus mid-span deflection response of s-1 and s-1R.

Table 7-15 summarises the cracking moments Mcr obtained from both the experimental and

numerical results for all the eight beams. The finite element model has predicted Mcr with

good accuracy compared with experimental results. It can be clearly seen that the beams with

heavy top bars had increased cracking moment compared to the beams with small top bars.

This can be illustrated by looking at the local responses of the beams. Figure 7-25 shows the

finite element results of stress and strain distributions at the mid-span sections of the first

series of beams. The y axis is the distance from the bottom to the top of the sections, whereas

the x axis is the strain or stress values just before loading. Due to shrinkage and the restraint

to free shrinkage by the reinforcing bars, stress and strain develop at the section. The restraint

forces on the concrete are significantly affected by the amount of reinforcement as known.
240

For the beams heavily reinforced with compressive bars (S-1R and S-2R), the shrinkage

strains on the top levels of the sections are greatly restrained by the reinforcement. Thus larger

restraining tensile force is applied on the concrete on the top levels of S-1R and S-2R,

compared with that of S-1 and S-2. Therefore the stress (strain) distributions of S-1R and S-

2R are actually in reversed directions compared with S-1 and S-2. The shrinkage induced

tensile stress on the bottom sides of the beams is also reduced before loading commenced.

Therefore, s-1R and s-2R exhibited higher cracking moments (Table 7-15).

Table 7-15: Measured and modeled Cracking moments for Gribniak’s beams.

Mcr (kN.m)
Beam
EXP FEM
S-1 16.8 17.0
S-1R 19.8 20.0
S-2 15.9 17.0
S2-R 17.9 21.0
S-3 15.8 16.0
S3-R - 19.0
S-4 13.9 15.0
S-4R 16.0 17.0
241

300 300
Distance from the bottom (mm)

Distance from the bottom (mm)


250 Compressive bar level 250
Compressive bar level
200 200

150 S-1 150


S-2
S-1R S-2R
100 100

50 Tensile bar level 50 Tensile bar level

Strain (×10-6) Strain (×10-6)


0 0
-200 -190 -180 -170 -160 -150 -140 -130 -170 -160 -150 -140 -130 -120 -110 -100

(a) s-1 and s-1R (strain) (b) s-2 and s-2R (strain)
300 300
Distance from the bottom (mm)

Distance from the bottom (mm)

250 Compressive bar level 250


Compressive bar level

200 200

S-1 150 S-2


150 S-2R
S-1R
100 100
Tensile bar level
Tensile bar level
50
50
stress (MPa)
stress (MPa) 0
0
-0.2 0.0 0.2 0.4 0.6 0.8 1.0
-0.3 -0.1 0.1 0.3 0.5 0.7 0.9

(c) s-1 and s-1R (stress) (d) s-2 and s-2R (stress)
Figure 7-25: Stress and strain distributions at the mid-span sections of Gribniak’s first series
beams just before loading (FEM).

In order to investigate the effects of compressive reinforcement on tension stiffening in

Gribniak’s test, the tension stiffening deflection versus mid-span moment diagrams for the

first series beams are plotted in Figure 7-26. The figure shows that the tension stiffening

deflections in S-1R is slightly greater than that in S-1 after first cracking, whereas little

difference is found between the results of S-2 and S-2R. It seems that even though the

provision of compressive reinforcement can reduce the cracking moment, influence on the

short-term tension stiffening still is not very clear. In chapter 8, a parametric study on the

effect of compressive reinforcement on tension stiffening is undertaken and a further


242

discussion is made.
Tension stiffening deflection (mm

Tension stiffening deflection (mm


5 5

4 4
s-1 s-2
3 s-1R 3 s-2R

2 2

1 1

0 0
0 10 20 30 40 50 60 0 10 20 30 40 50 60
Moment (kN.m) Moment (kN.m)

(a) s-1 and s-1R (b) s-2 and s-2R


Figure 7-26: Tension stiffening deflection versus mid-span moment of Gribniak’s first series
beams (experimental results).

7.6 SUMMARY
The finite element program incorporated with the proposed bond model has been used to

analyse the specimens tested by the author. Each specimen was decomposed into numbers of

elements, including concrete, steel truss and bond interface elements for the analysis. The

details of the configuration of each finite element model were presented. Shrinkage and creep

of the specimens measured at the tests were carefully examined and modeled, as well as the

instantaneous properties of concrete and steel bars.

The finite element modeling results are presented in two scales, namely the global scale and

the local scale. The global scale results investigated tension stiffening behaviour in terms of

the tension stiffening strain or tension stiffening deflection under both instantaneous loads and

sustained loads. But then the local scale results focused on the mechanisms of tension

stiffening through the concrete-steel interactions. Excellent correlation was achieved between
243

the finite element results and experimental results.

To further verify the accuracy and applicability of the finite element program, the long-term

and short-term tests conducted by Nejadi and Gilbert (2004) and Gribniak (2009) were

modeled by the author. Both of the tests had considered and measured shrinkage and creep.

Again good correlation was achieved. For the long-term slabs tested by Nejadi and Gilbert,

both the experiment and the model predicted a rapid decrease of tension stiffening in the first

15 days by 30%-50%, and then a more gradual decrease to 54% - 6% of the instantaneous

values. The major factor of reducing tension stiffening is attributed to the formation of new

crack due to shrinkage of the concrete and secondary bond deterioration. The testing and

modeling results for Gribniak’s beams show that the provision of compressive reinforcement

can greatly resist the effect of shrinkage on the uncracked beam section and reduce cracking

moments.

In general, the finite element model has been shown to accurately capture the tension

stiffening behavior under both short-term and long-term loading. The model considers the

effects of shrinkage and creep in the concrete, which can significantly reduce the effect of

tension stiffening.
CHAPTER 8 PARAMETRIC STUDY

8.1 OVERVIEW
In the forgoing chapters, it has been demonstrated that the magnitude of tension stiffening at

any time depends on the amount of shrinkage in the concrete before and after cracking. The

numerical model introduced in Chapter 6 has shown to accurately and rationally model the

mechanisms of tension stiffening and the correlation with experimental results is good.

However, the physical experiments conducted by the author are necessarily limited and a

comprehensive laboratory based study of various parameters affecting tension stiffening has

not been possible. For example, the time-dependent tension stiffening of the beams may be

influenced by the transverse reinforcements if provided. To undertake such a study effectively

and efficiently, it is necessary to conduct a series of computer experiments. In this chapter, a

number of numerical specimens are devised and modeled, and a parametric study is

undertaken using the 2-D finite element model described in Chapter 6. Both beam and slab

specimens are considered. The major parameters considered in the parametric study are:

a. Concrete cover cb;

b. Concrete tensile strength fct;

c. Concrete elastic modulus Ec;


245

d. Tensile reinforcement area Ast;

e. Tensile bar diameter db;

f. Magnitude of creep ijc;

g. Magnitude of shrinkage sh;

h. Compression reinforcement area Asc; and

i. Quantity of transverse reinforcement (stirrups).

In each series of numerical experiments, only a single parameter is varied and all the other

parameters are kept constant. Both short-term and long-term analyses are undertaken. For the

short-term analysis, shrinkage is allowed to develop prior to loading. For the long-term

analysis, shrinkage is only allowed to develop after first loading, when the loads are sustained

for a prescribed period. The variation of tension stiffening with time of each specimen is then

evaluated in terms of the change in the tension stiffening deflection. It should be remembered

that for a given specimen under consideration, the increase of tension stiffening deflection

indicates the reduction of beam or slab deflection. Therefore the examination of the effects of

these parameters on tension stiffening deflection reflects their effects on member behaviour

under service loads. Results of the parametric studies are discussed and conclusions are drawn

concerning the effects of the various parameters.

8.2 NUMERICAL EXPERIMENTS


8.2.1 Benchmark material properties

A benchmark specimen is used as a reference for both the beam and slab specimens. The

material properties for the benchmark specimens are:

fcp = 25MPa; fct = 2MPa; Ec = 25000MPa; v = 0.2; Es = 200000MPa; fsy = 500MPa; and Gf =
246

75N/m.

The creep and shrinkage characteristics of the benchmark specimens are given in Tables 8-1.

Table 8-1: Creep coefficient and shrinkage strain (x10-6) for the benchmark specimens.

Age (days) 14 20 25 30 40 50 75 100 114

Mc 0 0.25 0.55 0.75 1.06 1.33 1.74 1.90 2.0


Hsh (×10 )
-6
0 39.4 69.9 98.5 150.4 196.4 291.3 365.3 400.0

Table 8-2: Kelvin chain properties for solidification creep model.

E (MPa) (Days)
0.348 0.0001;
0.236 0.001
0.162 0.01
0.113 0.1
0.080 1
0.057 10
0.041 100

Shrinkage is assumed to commence at the 14th day and accumulate up to 400×10-6 after 100

days. Accordingly parameters Ash and Bsh are taken as 960×10-6 and 140days for the shrinkage

model (Equation 6.41). The benchmark creep coefficient after 100 days is taken as 2.0. The

corresponding solidification creep parameters are q2 = 100/MPa, q3 = 1.5/MPa, q4 =

40.0/MPa, and A0 = 37.3MPa-1. Table 8-2 summarises the retardation time  and elastic

modulus E for each Kelvin chain for the benchmark creep.

8.2.2 Beam specimens

Figure 8-1 shows the beam specimens analysed in the numerical experiments. All the beams

are simply-supported on the top of stiff plates. Each plate is of 100mm length and 250mm

width, located at 100mm from each end of the beam. Each beam is 250mm wide, with a span
247

L of 3m. The effective depth of the tensile reinforcement in the beam is dst = 280mm. The

bottom cover to the tensile reinforcement cb is treated as a variable in the experiment. Some

of the beams also contain compressive reinforcements Asc at the top of the members (with an

effective depth dsc = 30mm) and stirrups over the length of the beams. A uniformly distributed

load q is applied on the top of the beams so that the concrete at the bottom of the beam is in

tension. For the short-term tests, q is increased monotonically until the steel is about to yield.

For the long-term tests, q is sustained for a period of 100 days. The level of the sustained load

is typical of the design load at the serviceability limit state.

Figure 8-1: Construction of beam specimens in the numerical experiment.

In the finite element model, only half the length of the beam is modeled, as shown in Figure

8-2. The uniformly distributed applied load is applied as nodal loads on the concrete elements

on the top of the beams. It is not necessary to adopt too fine a mesh layout. In these numerical

experiments, 25mm by 25mm isoparametric concrete elements are adopted to model the

concrete, together with reinforcing bar truss elements and bond interface elements. The

compressive reinforcing and stirrup truss elements are directly attached to the surrounding

concrete element nodal points. The benchmark beam model consists of 658 nodal points, 552

concrete elements, 46 tensile reinforcing truss elements and 46 interface elements.


248

Figure 8-2: Finite element model of beam specimens in the numerical experiment.

8.2.3 Slab specimens

The slab specimens are also simply-supported over a length of 3m, with uniformly distributed

loads imposed on the top of the slabs (Figure 8-3). Each slab has a cross section 400mm wide

and D depth. The effective depth of the tensile reinforcing bars dst is taken as 130mm. The

mesh layout for the slabs is shown in Figure 8-4. The quadrilateral concrete element size is

also taken as 25mm × 25mm. The benchmark slab specimen model consist of 376 nodal

points, 276 concrete elements, 46 tensile reinforcing truss elements and 46 bond interface

elements.
249

Figure 8-3: Construction of slab specimens in the numerical experiment.

Figure 8-4: Finite element model of slab specimens in the numerical experiment.

8.2.4 Experimental variables

The various parameters investigated in each series of numerical test are provided in Tables 8-3

and 8-4. The letter (‘A’ to ‘I’) represents the different testing series. Each series contains 2 – 4

specimens. The parameters investigated in each test series are highlighted with the dark

shading in the table. Normally it is specimen No.1 in each series that is defined as the

benchmark specimen. It should be noted that the maximum bond stress (max) alter according

to the various concrete strengths selected in the test series B. The way to modify the values of

max will be introduced in the following sections.


250

Table 8-3: Various parameters for the beam specimens in the numerical experiment.

cb fct Ec Ast No. of db Hsh Asc


Series Specimen Mc Stirrup
(mm) (MPa) (MPa) (mm2) bars (mm) (×10-6) (mm2)
1 20 2 25000 400 2 16 2 400 - -
A 2 30 2 25000 400 2 16 2 400 - -
3 40 2 25000 400 2 16 2 400 - -
1 20 2 25000 400 2 16 2 400 - -
2 20 1.5 25000 400 2 16 2 400 - -
B
3 20 2.5 25000 400 2 16 2 400 - -
4 20 3 25000 400 2 16 2 400 - -
1 20 2 25000 400 2 16 2 400 - -
2 20 2 15000 400 2 16 2 400 - -
C
3 20 2 35000 400 2 16 2 400 - -
4 20 2 45000 400 2 16 2 400 - -
1 20 2 25000 400 2 16 2 400 - -
D 2 20 2 25000 309 2 14 2 400 - -
3 20 2 25000 226 2 12 2 400 - -
1 20 2 25000 309 2 14 2 400 - -
E
2 20 2 25000 309 4 10 2 400 - -
1 20 2 25000 400 2 16 2 400 - -
2 20 2 25000 400 2 16 0.5 400 - -
F
3 20 2 25000 400 2 16 1 400 - -
4 20 2 25000 400 2 16 1.5 400 - -
1 20 2 25000 400 2 16 2 400 - -
2 20 2 25000 400 2 16 2 200 -
G
3 20 2 25000 400 2 16 2 300
4 20 2 25000 400 2 16 2 500
1 20 2 25000 400 2 16 2 400 226 -
2 20 2 25000 400 2 16 2 400 400 -
H
3 20 2 25000 400 2 16 2 400 509 -
4 20 2 25000 400 2 16 2 400 628 -
1 20 2 25000 400 2 16 2 400 226 -
2 20 2 25000 400 2 16 2 400 226 2 8@100
I
3 20 2 25000 400 2 16 2 400 226 2 10@100
4 20 2 25000 400 2 16 2 400 226 2 12@100
251

Table 8-4: Various parameters for the slab specimens in the numerical experiment.

cb fct Ec Ast No. of db Hsh


Series Specimen Mc
(mm) (MPa) (MPa) (mm2) bars (mm) (×10-6)
1 15 2 25000 452 4 12 2 400
A 2 20 2 25000 452 4 12 2 400
3 25 2 25000 452 4 12 2 400
1 15 2 25000 452 4 12 2 400
2 15 1.5 25000 452 4 12 2 400
B
3 15 2.5 25000 452 4 12 2 400
4 15 3 25000 452 4 12 2 400
1 15 2 25000 452 4 12 2 400
2 15 2 15000 452 4 12 2 400
C
3 15 2 35000 452 4 12 2 400
4 15 2 45000 452 4 12 2 400
1 15 2 25000 452 4 12 2 400
D 2 15 2 25000 314 4 10 2 400
3 15 2 25000 603 4 16 2 400
1 15 2 25000 452 4 12 2 400
E
2 15 2 25000 470 6 10 2 400
1 15 2 25000 452 4 12 2 400
2 15 2 25000 452 4 12 0.5 400
F
3 15 2 25000 452 4 12 1 400
4 15 2 25000 452 4 12 1.5 400
1 15 2 25000 452 4 12 2 400
2 15 2 25000 452 4 12 2 200
G
3 15 2 25000 452 4 12 2 300
4 15 2 25000 452 4 12 2 500

8.3 DISCUSSION OF TEST RESULTS


The results of each test series are presented together so that the effect of particular variable in

that series can be clearly investigated. The mid-span deflections of the specimens are directly

obtained from the finite element program and the results are used to derive tension stiffening

deflections. The results of tension stiffening and its variation with time are reported in terms

of tension stiffening deflection.

8.3.1 Bottom concrete cover, cb – Series A

The influence of bottom concrete cover on time-dependent tension stiffening deflection is


252

investigated in Series A, which includes 3 beams and 3 slabs under short-term loading, and

the same numbers of beams and slabs under long-term loading. Figure 8-5 shows the moment

versus mid-span deflection responses of the short-term tests. Notice that since the effective

depths of the tensile steel bars for each beam series or slab series are identical, the calculated

fully-cracked responses are also identical. The numerical results show that deformation of the

members occurs before loading due to initial shrinkage strain (-400×10-6). Shrinkage prior to

loading also brings down the cracking moment from Mcr to Mcr.sh (from 18kN.m to 13.4kN.m

for beams and from 7.3kN.m to 6.2kN.m for slabs) for the specimens with shrinkage. The

results indicate that the variation of bottom concrete cover has little effect on Mcr.sh provided

the initial shrinkage for each specimen is the same. It also appears that the short-term mid-

span deflections of the beams or slabs with different concrete cover are quite similar.
Moment (kN.m)

50 30
Moment (kN.m)

40
20
30
C
Cbb == 20mm
20mm C b = 15mm
S-A-S-1
20 C b = 30mm
B-A-S-2 10 C b = 20mm
S-A-S-2
C b = 40mm
B-A-S-3 C b = 25mm
S-A-S-3
10

Mid-span deflection (mm) Mid-span deflection (mm)


0 0
0 3 6 9 12 15 0 6 12 18 24 30 36

(a) Short-term beams (b) Short-term slabs


Figure 8-5: Instantaneous moment versus mid-span deflection responses of series A short-
term specimens.

Figure 8-6 plots the tension stiffening deflection versus mid-span moment responses of the

series A short-term test specimens. Notice that at zero moment, tension stiffening deflection is

greater than zero due to initial shrinkage. The results show that with different bottom concrete

cover, tension stiffening deflections at the same moment level are slightly affected by the
253

bottom concrete cover. The magnitudes of tension stiffening deflections are slightly greater in

the specimens with larger bottom covers. For instance, for the beams specimens, at M =

26.9kN.m, the tension stiffening deflection for the beam specimen with cb = 20mm equals to

1.68mm, which is less than that in the beam specimen with cb = 40mm that equals to 1.94mm.

However, the difference between the beams with cb = 30mm and cb = 40mm is not that

significant as shown in Figure 8-6a. For the slabs specimens, at M = 16.3kN.m, the tension

stiffening deflection changes from 2.45mm to 2.60mm, as cb increases from 15mm to 25mm.

It could be perceived that with even larger bottom cover, the magnitude of tension stiffening

can be further increased.


Tension stiffening deflection (mm

4 7
Tension stiffening deflection (mm

6
3 5

4
2
3

CCb
b = 2 CCb
b =
=15mm
1 =20mm
20mm 20mm
CB-A-S-2
b = 30mm
CS-A-S-2
b = 20mm
1 CS-A-S-3
CB-A-S-3
b = 40mm b = 25mm

0 0
0 10 20 30 40 50 60 0 5 10 15 20 25 30
Mid-span moment (kN.m) Mid-span moment (kN.m)

(a) Short-term beams (b) Short-term slabs


Figure 8-6: Tension stiffening deflection versus mid-span moment responses of series A
short-term specimens.

Figure 8-7 plots the development of long-term mid-span deflections for beams and slabs

under sustained loads in test series A. The beams and slabs are under sustained uniformed

distributed loads, which give rise to a constant mid-span moment M = 25.8kN.m and

10.7kN.m for beams and slabs respectively. Similar to the short-term specimens, the mid-span

deflections for the long-term test specimens do not vary significantly as the concrete cover
254

varies. With time, the development of mid-span deflection of each beam or slab is similar,

with the curves for the slabs are almost the same, as shown in Figure 8-7b. For the beams

subjected to long-term sustained loads, the developments of mid-span deflections of

specimens Beam 1 and Beam 2 (cb = 20mm and cb = 30mm) are quite similar, whereas beam

4 (cb = 40mm) exhibits greater long-term stiffness under the same applied moment. The finite

element models have given rise to less numbers of primary cracks in Beam 4 than Beam 1 and

Beam 2, which are regarded as the major reason for the improved stiffness in Beam 4.

10 20
Mid-span deflection (mm)

Mid-span deflection (mm)

8 16

6 12

4 B-A-L-1
C b = 20mm 8 C b = 15mm
S-A-L-1
B-A-L-2
C b = 30mm C b = 20mm
S-A-L-2
B-A-L-3
C = 40mm S-A-L-3
C b = 25mm
2 b
4

Time (days) Time (days)


0 0
0 20 40 60 80 100 0 20 40 60 80 100

(a) Long-term beams (b) Long-term slabs


Figure 8-7: Mid-span deflection versus time responses of series A long-term specimens.

Table 8-5 summarises the mid-span deflections ms(t), and the calculated tension stiffening

deflections ts(t) at different stages under loading (t = 0, 15 days and 100 days). It is found

that by increasing the bottom concrete cover in reinforced concrete beams, the time-dependent

tension stiffening deflections are increased. As shown in the table, for the beam specimens, at

t = 15 days, ts(t) increases from 1.27mm to 1.71mm as the bottom cover increases from

20mm to 40mm. The rate of decrease of tension stiffening with time is also decreased with the

increased bottom cover. For example, for the beam with cb = 20mm, ts(t)/ tsi = 0.64 on t = 15

days and 0.38 on t = 100 days, whereas for the beam with cb = 40mm, the corresponding
255

values are 0.72 and 0.56. However, this phenomenon is not apparent for the slab specimens.

Even though the magnitude of time-dependent tension stiffening deflection ts(t) slightly

increases with increased bottom cover, the rate of decrease of tension stiffening seems

unaffected in the slabs with different bottom covers.

Table 8-5: Tension stiffening deflections of series A long-term specimens.

Deflections (mm)
Cb Instantaneous t = 15 days t = 100 days
Specimen
(mm) ts(t)/ ts(t)/
msi tsi ms(t) ts(t) ms(t) ts(t)
tsi tsi
Beam 1 20 3.69 1.99 5.58 1.27 0.64 8.49 0.75 0.38
Beam 2 30 3.61 2.08 5.51 1.34 0.65 8.42 0.82 0.40
Beam 3 40 3.31 2.37 5.14 1.71 0.72 7.91 1.33 0.56
Slab 1 15 6.99 3.61 10.9 2.31 0.64 17.2 1.48 0.41
Slab 2 20 6.79 3.81 10.76 2.45 0.64 17.12 1.56 0.41
Slab 3 25 6.73 3.87 10.73 2.48 0.64 17.13 1.55 0.40

It is generally agreed that concrete cover is an important factor in the development of the

crack pattern and crack spacing. With less bottom cover, more cracks will form and hence

break down the local bond stress at the interface between concrete and steel bars. The

magnitude of tension stiffening will drop off with the breakdown of bond at the cracked

sections. Therefore, the time-dependent stiffness of a reinforced concrete member can be

implicitly improved by larger concrete cover. For example, the analytical model have shown

that Beam 3 (cb = 40mm) exhibits less cracks than Beam 1 (cb = 20mm) and Beam 2 (cb =

30mm). Study conducted by Salem, et al (1999) drew the same conclusion in their study of

the effect of concrete cover on tension stiffening of reinforced concrete members. They

concluded that spitting cracks due to lack of concrete cover can greatly affect the bond effect

and reduce the stiffness of a reinforced concrete member.


256

8.3.2 Concrete tensile strength, fct – Series B

The test specimens in Series B are assigned with the different concrete tensile strengths fct. A

group of four beams and four slabs was tested under short-term loading, and another group

with the identical number of beams and slabs was subjected to long-term loads. In this case, it

is also necessary to adjust the corresponding peak bond stress max for the modeling input,

since the bond strength is dependent on the concrete strength (max equals to 2.5 times the

square root of fcp ). Thus, max increases with the increased fct. Table 8-6 shows the values of

max for each specimen in this series. In addition, according to the change of fct, the parameters

1 2 and 3 defining the descending branch of Petersson’s equation (Equation 6.10) are also

modified.

Table 8-6: max for the series B test specimens (All in MPa).

Specimens fct max


Beam 1
2.0 12.5
Slab 1
Beam 2
1.5 9.7
Slab 2
Beam 3
2.5 15.6
Slab 3
Beam 4
3.0 18.8
Slab 4

Figure 8-8 illustrates the moment versus mid-span deflection responses of the short-term

specimens of the series B tests. An increase of tensile strength obviously leads to increased

cracking moment Mcr and decreased deflections after cracking. Specimens with lower tensile

strength exhibit more cracks (smaller crack spacing), which results in reduced stiffness and

less tension stiffening. Figure 8-9 shows the tensile strain contour of the beams with fct =

2.5MPa and fct = 3.0MPa at M = 30kN.m, where the darkened areas demonstrate the crack

locations (tensile strain greater than the cracking strain). There are 7 primary cracks found in
257

the former and only 5 in the latter.


Moment (kN.m)

50 30

Moment (kN.m)
40
20
30
f ct = 1.5MPa
S-B-S-1 2.0MPa
f ct = 1.5MPa
B-B-S-1 f ct = 2.0MPa
1.5MPa
20 S-B-S-2
f ct = 2.0MPa
B-B-S-2 10 fS-B-S-3
ct = 2.5MPa
f ct = 2.5MPa
B-B-S-3 fS-B-S-4
ct = 3.0MPa
10 f ct = 3.0MPa
B-B-S-4
Mid-span deflection (mm) Mid-span deflection (mm)
0 0
0 3 6 9 12 15 0 6 12 18 24 30 36

(a) Short-term beams (b) Short-term slabs


Figure 8-8: Instantaneous moment verse mid-span deflection responses of series B short-term
specimens.

(a) Beam 3 (fct = 2.5MPa) (b) Beam 4 (fct = 3.0MPa)


Figure 8-9: Tensile strain contour of series B short-term specimens at M = 30kN.m.

In reality, the elastic modulus Ec and the strength of concrete are normally interdependent

properties. However, this test series tries to focus on the influence of fct and Ec was held

constant for all specimens. Based on the formulas provided in Appendix B, the fully-cracked

response is independent of fct but dependent on Ec, therefore the fully-cracked deflections for

the beams or slabs are identical. The fully-cracked deflections of each specimen were

calculated, as well as the tension stiffening deflections. The results for each short-term test

specimen are plotted in Figure 8-10. The magnitude of tension stiffening deflection is

obviously greater in the specimen with largest tensile strength. For instance, at a typical post-

cracking moment M = 16.3kN.m, tension stiffening deflection ts = 2.45mm, 2.96mm and
258

3.89mm for the slabs with fct = 2.0MPa, 2.5MPa and = 3.0MPa, respectively.

Tension stiffenign deflection (mm)


Tension stiffenign deflection (mm)

5 8

fB-B-S-1
ct = 1.5MPa fS-B-S-1
ct = 1.5MPa
2.0MPa
4
fB-B-S-2
ct = 2.0MPa
6 fS-B-S-2
ct = 2.0MPa
1.5MPa
fB-B-S-3
ct = 2.5MPa fS-B-S-3
ct = 2.5MPa
3
fB-B-S-4
ct = 3.0MPa fS-B-S-4
ct = 3.0MPa
4
2
2
1

0 0
0 10 20 30 40 50 60 0 5 10 15 20 25 30
Moment (kN.m) Moment (kN.m)

(a) Short-term beams (b) Short-term slabs


Figure 8-10: Tension stiffening deflection verse mid-span moment responses of series B
short-term specimens.

Figure 8-11 shows the development of mid-span deflections with time for the long-term

specimens in series B. The specimens were subjected to the same amount of loading as those

in Series A long-term tests. The specimens with smaller tensile strength undergo larger

deformation with time. Table 8-7 tabulates the mid-span deflections and tension stiffening

deflections for the long-term specimens at instantaneous loading, 15 days and 100 days after

first loading. The mid-span deflections by the end of the test are about 2 - 3 times of the

instantaneous deflections. With larger tensile strength, the proportion of final deflection to

instantaneous deflection ms(t)/ msi is also larger. For instance, the values of ms(t) / msi (t =

100 days) for Beam 1, Beam 3 and Beam 4 (fct = 2.0MPa, 2.5MPa and 3.0MPa) are 2.30, 2.73

and 3.48, respectively. This is attributed to the fact that at the instantaneous moment (M =

25.8kN.m), the specimen with the larger tensile strength suffers less cracking and thus the

instantaneous deflection is smaller. Time-dependent cracking occurs as time progresses and

this reduces the stiffness of the member and increases the long-term deflections significantly.
259

10 20
Mid-span deflection (mm)

Mid-span deflection (mm)


8 16

6 12

f ct = 1.5MPa fS-B-L-1
ct = 1.5MPa
B-B-L-1
4 f ct = 2.0MPa 8 fS-B-L-2
ct = 2.0MPa
B-B-L-2
fS-B-L-3
ct = 2.5MPa
f ct = 2.5MPa
B-B-L-3
2 4 fS-B-L-4
ct = 3.0MPa
f ct = 3.0MPa
B-B-L-4
Time (days) Time (days)
0 0
0 20 40 60 80 100 0 20 40 60 80 100

(a) Long-term beams (b) Long-term slabs


Figure 8-11: Mid-span deflection versus time responses of series B long-term specimens.

Table 8-7: Tension stiffening deflections of series B long-term specimens.

Deflections (mm)
Instantaneous t = 15 days t = 100 days
Specimens fct (MPa)
ts(t)/ ts(t)/
msi tsi ms(t) ts(t) ms(t) ts(t)
tsi tsi
Beam 1 2.0 3.69 1.99 5.58 1.27 0.64 8.49 0.75 0.38
Beam 2 1.5 4.60 1.08 6.33 0.52 0.48 8.92 0.32 0.30
Beam 3 2.5 2.88 2.80 4.96 1.89 0.68 7.88 1.36 0.49
Beam 4 3.0 1.91 3.77 3.71 3.14 0.83 6.65 2.59 0.69
Slab 1 2.0 6.99 3.61 10.9 2.31 0.64 17.2 1.48 0.41
Slab 2 1.5 7.90 2.70 11.72 1.49 0.55 17.91 0.77 0.29
Slab 3 2.5 6.34 4.26 10.38 2.83 0.66 16.65 2.03 0.48
Slab 4 3.0 5.52 5.08 9.52 3.70 0.73 15.76 2.92 0.58

With regard to the change of tension stiffening deflection with time, after t = 15 days or 100

days, both the magnitude of tension stiffening deflection ts(t) and the ratio of time-dependent

tension stiffening deflection to instantaneous tension stiffening deflection ts(t)/tsi increase as

the tensile strength increases. For example at t = 100 days, the values of ts(t) for Beam 1,

Beam 3 and Beam 4 (fct = 2.0MPa, 2.5MPa and 3.0MPa) are 0.75mm, 1.36mm and 2.59mm,

respectively. Meanwhile, the values of and ts(t)/tsi for these specimens are 0.38, 0.49 and

0.69, respectively. Therefore, the specimen with larger tensile strength exhibits greater tensile

stiffening deflection and slower rate of change in tension stiffening with time. Note that in this
260

study, the tensile strength of concrete is treated as a constant while in reality it is also a time-

dependent parameter, which normally increases with time. It is reasonable to assume that the

decreasing behaviour of tension stiffening could be more unpredictable if the time-varying

nature of tensile strength is considered.

8.3.3 Concrete elastic modulus, Ec – Series C

Series C consists of specimens designed with different values for the elastic modulus of

concrete Ec. Four beams and four slabs are chosen for both the short-term and long-term

numerical experiments. The design details are given in Tables 8-3 and 8-4.

Figure 8-12 shows the moment versus mid-span deflection responses for the short-term

specimens in series C. Before loading, the initial deflections of each specimen caused by

shrinkage warping vary. The specimens with larger concrete elastic modulus exhibit slightly

smaller initial deflections. For the short-term beams and slabs under consideration, the initial

deflections before loading are:

Beams: 0.66mm (Ec = 15000MPa), 0.55mm (Ec = 25000MPa), 0.49mm (Ec = 35000MPa),

and 0.46mm (Ec = 45000MPa);

Slabs: 1.50mm (Ec = 15000MPa), 1.24mm (Ec = 25000MPa), 1.11mm (Ec = 35000MPa) and

1.04mm (Ec = 45000MPa).

The difference in the shrinkage-induced deflection is because the time-dependent curvature

due to shrinkage at an uncracked section is inversely proportionally to the elastic modulus of

concrete. Namely the larger is the elastic modulus of concrete, the smaller is the initial

curvature or the initial deflection. After cracking, the deformation responses of the beam or
261

slab specimens with Ec are shown in Figure 8-12. The slope of the post-cracking response

increases as Ec increases, with the specimens with Ec = 15000MPa clearly showing the

greatest post-cracking range.


Moment (kN.m)

50 30

Moment (kN.m)
40
20
30

E c = 15000MPa
BES1 ESES1
c = 15000MPa
20 ESES2
E c = 25000MPa
BES2 c = 25000MPa
10
E c = 35000MPa
BES3 ESES3
c = 35000MPa
10 E c = 45000MPa ESES4
c = 45000MPa
BES4
Mid-span deflection (mm) Mid-span deflection (mm)
0 0
0 3 6 9 12 15 0 6 12 18 24 30 36

(a) Short-term beams (b) Short-term slabs


Figure 8-12: Instantaneous moment verse mid-span deflection responses of series C short-
term specimens.

Since the fully-cracked response is dependent on the magnitude of concrete elastic modulus,

in this series, the fully-cracked deflections vary among different specimens. Figure 8-13 plots

the tension stiffening deflection versus mid-span moment responses for the short-term

specimens considered herein. For the beam specimens, the influence of concrete elastic

modulus on the quantity of short-term tension stiffening deflection can hardly be told. Beam 3

and Beam 4 (Ec = 35000MPa and 45000MPa) show similar amount of tension stiffening,

while both Beam 1 and Beam 2 (Ec = 25000MPa and 15000MPa) exhibit slightly greater

tension stiffening deflection than the formers. The slabs also show a similar phenomenon.

Tension stiffening deflections in Slab 3 and Slab 4 (Ec = 35000MPa and 45000MPa) are quite

similar, while Slab 1 and Slab 2 (Ec = 25000MPa and 15000MPa) possess greater values of

tension stiffening deflections. Note that this difference only becomes significant as the

reinforcing steel are about to yield. Before the applied moment is greater than 2 – 3 times of
262

the cracking moment, the effect of concrete elastic modulus on short-term tension stiffening

deflection is only slight.

Tension stiffenign deflection (mm)


Tension stiffenign deflection (mm)

4 8
EBES1
c = 15000MPa

EBES2
c = 25000MPa ESES1
c = 15000MPa
3 EBES3 6 ESES2
c = 25000MPa
c = 35000MPa

EBES4
c = 45000MPa ESES3
c = 35000MPa

4 ESES4
c = 45000MPa
2

1 2

0 0
0 10 20 30 40 50 60 0 5 10 15 20 25 30
Moment (kN.m) Moment (kN.m)

(a) Short-term beams (b) Short-term slabs


Figure 8-13: Tension stiffening deflection verse mid-span moment responses of series C
short-term specimens.

For the long-term specimens in this series, all the beams and slabs were initially loaded to

25.8kN.m and 10.7kN.m (moment at mid-span), respectively, and held constant with time.

Figure 8-14 shows the development of mid-span deflections during the loading history for the

long-term beams and slabs in this series. No significant difference of mid-span deformations

with time is observed for the specimens with the concrete elastic modulus ranging from

25000MPa to 45000MPa. However, both the instantaneous and time-dependent deformations

of Beam 2 and Slab 2 (Ec = 15000MPa) are slightly larger than the other specimens.

Table 8-8 gives the time-dependent tension stiffening deflections of the long-term test

specimens in this series. With increased concrete elastic modulus, the magnitude and rate of

change of the time-dependent tension stiffening deflection decreases. For instance, ts(t)/ tsi

for Slab 1 (Ec = 25000MPa), Slab 3 (Ec = 35000MPa) and Slab 4 (Ec = 45000MPa) at the age

t = 15 days are 0.64, 0.47 and 0.29, respectively. The time-dependent loss of bond stress and
263

tensile stress in the concrete at the level of reinforcing steel does not directly depends on the

elastic modulus of concrete Ec. However, the magnitude and the rate of change of tension

stiffening is related to Ec, because the fully-cracked response is highly affected by Ec. The

time-dependent fully-cracked response is weaker in the specimens with smaller Ec, therefore

the tension stiffening deflection is greater in these specimens.

20
Mid-span deflection (mm)

10

Mid-span deflection (mm)


8 16

6 12

EBEL1 SEL1
E c = 15000MPa
c = 15000MPa
4 8 SEL2
E c = 25000MPa
EBEL2
c = 25000MPa
SEL3
E c = 35000MPa
EBEL3
c = 35000MPa
2 4 SEL4
E c = 45000MPa
EBEL4
c = 45000MPa

Time (days) Time (days)


0 0
0 20 40 60 80 100 0 20 40 60 80 100

(a) Long-term beams (b) Long-term slabs


Figure 8-14: Mid-span deflection versus time responses of series C long-term specimens.

Table 8-8: Tension stiffening deflections of series C long-term specimens.

Deflections (mm)
Instantaneous t = 15 days t = 100 days
Specimens Ec (MPa)
ts(t)/ ts(t)/
msi tsi ms(t) ts(t) ms(t) ts(t)
tsi tsi
Beam 1 25000 3.69 1.99 5.58 1.27 0.64 8.49 0.75 0.38
Beam 2 15000 4.20 2.10 5.90 1.60 0.76 8.79 1.20 0.57
Beam 3 35000 3.60 1.75 5.53 0.82 0.47 8.42 0.58 0.34
Beam 4 45000 3.38 1.78 5.59 0.52 0.29 8.41 0.28 0.16
Slab 1 25000 6.99 3.61 10.9 2.31 0.64 17.2 1.48 0.41
Slab 2 15000 8.72 3.49 12.26 2.37 0.68 18.5 1.85 0.53
Slab 3 35000 6.51 3.39 10.66 1.64 0.48 16.93 1.27 0.37
Slab 4 45000 6.08 3.34 10.35 1.53 0.46 16.79 0.81 0.24
264

8.3.4 Tensile reinforcement area, Ast – Series D

The effect of the quantity of tensile reinforcement Ast on tension stiffening is investigated by

the Series D tests. There are 3 beams and 3 slabs modeled in both the short-term tests and

long-term tests in this series.

The global responses of the short-term tests are plotted in Figure 8-15. With the same amount

of initial creep and shrinkage in the concrete, little difference of stiffness is found before first

cracking of the members. Only slight variation of initial deflections due to shrinkage warping

is found (Figure 8-15), which indicates that shrinkage warping is dependent on the amount of

reinforcement, because the restraining force at the level of the steel bars is proportional to the

reinforcement area. Therefore, shrinkage warping on the uncracked beams or slabs is more

significant for members with larger reinforcement area. Hence the greater the tensile

reinforcement area is, the larger the initial deflection prior to loading. For example, the initial

mid-span deflections for the beams with Ast = 400mm2, 309mm2 and 226mm2 are 0.55mm,

0.45mm and 0.36mm, respectively.

After cracking, the stiffness in the members with less tensile reinforcement is obviously

smaller than the stiffness of members with larger tensile reinforcement area. The flatten out of

the moment deflection responses for Beam 3 and Slab 2 as the moment increases is caused by

yielding of the tensile reinforcement.


265

Moment (kN.m)
Moment (kN.m)

50 30 2
AS-C-S-1
st = 314mm
2
AS-C-S-2
st = 452mm
40 2
AS-C-S-3
st =603mm
20
30

2
B-C-S-1
A st = 400mm
20
B-C-S-2
A st = 309mm
2 10
2
10 B-C-S-3
A st = 226mm

Mid-span deflection (mm) Mid-span deflection (mm)


0 0
0 3 6 9 12 15 18 0 6 12 18 24 30 36

(a) Short-term beams (b) Short-term slabs


Figure 8-15: Instantaneous moment verse mid-span deflection responses of series D short-
term specimens.

Figure 8-16 shows the tension stiffening deflection versus mid-span moment response for the

short-term tests in series D. It appears that with decreased tensile reinforcement area Ast (or

reinforcing ratio ȡ), the magnitude of tension stiffening deflection is more significant both

before and after cracking. This is also true in terms of the contribution of tension stiffening to

the total stiffness of the member. For instance, at M = 26.9kN.m the values of fc/ms (ratio

between fully-cracked deflection and mid-span deflection) for the beams with ȡ = 0.005,

0.004 and 0.003 are 1.52, 1.67 and 1.73, respectively. Therefore, in a member with small

reinforcement ratio, such as slab, the contribution of tension stiffening to the total member

stiffness is more significant, and the importance of accurately modeling tension stiffening in

deflection calculation increases. Besides, the figure shows that at the time when yielding is

about to occur, tension stiffening deflection tends to drop off dramatically. The short-term slab

tests conducted by Gilbert (2007) drew a similar conclusion.


266

Tension stiffenign deflection (mm)


Tension stiffening deflection (mm

5 2 8
A st = 400mm
B-C-S-1
2 2
A st = 309mm
B-C-S-2 S-B-S-1
A st = 314mm
4 2 2
B-C-S-3
A st = 226mm 6 S-B-S-2
A st = 452mm
2
A st =603mm
S-B-S-3
3
4
2

2
1

0 0
0 10 20 30 40 50 60 0 5 10 15 20 25 30
Mid-span moment (kN.m) Moment (kN.m)

(a) Short-term beams (b) Short-term slabs


Figure 8-16: Tension stiffening deflection verse mid-span moment responses of series D
short-term specimens.

Figure 8-17 shows the development of the long-term deflections of the series D specimens

under constant sustained loads (with mid-span moment M = 25.8kN.m for beams and M =

10.7kN.m for slabs) for a period of 100 days. Similar to the short-term deflections, the long-

terms deflections are greater in the members with smaller tensile reinforcement areas.

The calculated tension stiffening deflections ts(t) for the long-term specimens in this series

are given in Table 8-9. It seems that the magnitudes of ts(t) are also greater in the specimens

with smaller reinforcing ratio, either after t = 15 days or t = 100 days. For example, after t =

15 days ts(t) equals 2.94mm, 2.31mm and 1.82mm for the slabs with ȡ = 0.005, 0.0075 and

0.01, respectively. However, the proportion of ts(t) to tsi is not as affected by the

reinforcement ratio, as seen in the table. As discussed in Chapter 7, the experimental results of

the long-term slabs tested by Nejadi and Gilbert (2004) also demonstrated that the rate of

change of tension stiffening is not significantly affected by the reinforcing ratio. The proposed

bond model only relates the long-term decay of bond stress to the quantity of shrinkage and

the level steel stress, but not the reinforcement area. However, the magnitude of local steel

stress is also dependent on the reinforcement area. At the same level of bending moment M,
267

the sections in the members with larger reinforcement areas have smaller steel stresses, which

can lead to a smaller rate of change of bond stress with time (Equation 6.65). In addition, with

larger reinforcement area, the shrinkage induced compressive force on the reinforcing bars is

also larger. Thus if the crack spacing in a specimen has already stabilised under instantaneous

loads, the effect of reinforcement area on the rate of change of tension stiffening with time is

implicitly expressed by the change of steel stress with time.

12 25
Mid-span deflection (mm)

Mid-span deflection (mm)


10 20

8
15
6
2 2
A st = 400mm
B-C-L-1 10 A st = 314mm
S-C-L-1
2 2
4 A st = 309mm
B-C-L-2 A st = 452mm
S-C-L-2
2 2
A st = 226mm
B-C-L-3 5 S-C-L-3
A st =603mm
2
Time (days) Time (days)
0 0
0 20 40 60 80 100 0 20 40 60 80 100

(a) Long-term beams (b) Long-term slabs


Figure 8-17: Mid-span deflection versus time responses of series D long-term specimens.

Table 8-9: Tension stiffening deflections of series D long-term specimens.

Deflections (mm)
Instantaneous t = 15 days t = 100 days
Specimen ȡ
ts(t)/ ts(t)/
ms tsi ms ts(t) ms ts(t)
tsi tsi
Beam 1 0.005 3.69 1.99 5.58 1.27 0.64 8.49 0.75 0.38
Beam 2 0.004 4.75 2.32 6.69 1.58 0.68 9.71 1.04 0.45
Beam 3 0.003 5.93 3.26 8.50 1.99 0.61 11.86 1.23 0.38
Slab 1 0.0075 6.99 3.61 10.9 2.31 0.64 17.2 1.48 0.41
Slab 2 0.005 9.63 4.49 13.97 2.94 0.66 20.61 1.98 0.44
Slab 3 0.01 5.52 3.01 9.23 1.82 0.60 15.33 1.08 0.36
268

8.3.5 Tensile bar diameter, db – Series E

The numerical models in Series E were established to look into the significance of tensile bar

diameter db on time-dependent tension stiffening. There are only 2 beams and 2 slabs that are

modeled for both the long-term and short-term tests in this series. According to Table 8-3 and

8-4, the beams or slabs are designed with the similar amount of tensile reinforcement area but

different diameters of bars. Therefore, the value of c for the bond interface elements in

equation 6.51 is the only variable parameter in this testing series.


Tension stiffening deflection (mm

6
Moment (kN.m)

50
S-D-S-1
412
214
5
40 410
610
S-D-S-2
4
30
3
20 214
B-D-S-1
410
2
B-D-S-2
10 1
Mid-span deflection (mm)
0 0
0 3 6 9 12 15 18 0 10 20 3
Mid-span moment (kN.m

(a) Short-term beams (b) Short-term slabs


Figure 8-18: Instantaneous moment verse mid-span deflection responses of series E short-
term specimens.

The short-term global moment versus mid-span deflection responses are shown in Figure 8-15.

The results indicate that there is very slight difference for the members reinforced with bars of

different diameters, either before cracking or after cracking. Since the reinforcement area and

the depth for the reinforcing bars are similar for each pair of specimens, the fully-cracked

responses shall be similar as well.

As shown in Figure 8-19, the quantity of short-term tension stiffening deflection also exhibits

little difference between the members with different bar diameters. It can be concluded by the
269

numerical test results that the bar diameter has a very minor effect on the quantity of short-

term tension stiffening in reinforced concrete beams or slabs. This observation is similar to

what was found by Sooriyaarachchi et al. (2005), as mentioned in Chapter 2.

Tension stiffening deflection (mm


Tension stiffening deflection (mm

5 6
B-D-S-1
214 S-D-S-1
412
214
4 5
410
B-D-S-2 410
610
S-D-S-2
4
3
3
2
2
1 1

0 0
0 10 20 30 40 50 60 0 10 20 30
Mid-span moment (kN.m) Mid-span moment (kN.m)

(a) Short-term beams (b) Short-term slabs


Figure 8-19: Tension stiffening deflection verse mid-span moment responses of series E
short-term specimens.

Similar responses are also found in the results of long-term tests in this series. The time-

dependent tension stiffening deflections ts(t) seems not to be influenced by bar diameters,

both in terms of its magnitude and its proportion of the instantaneous tension stiffening

deflections tsi (Figure 8-20 and Table 8-10).


270

12 25

Mid-span deflection (mm)


Mid-span deflection (mm)

10 20

8
15
6
214
B-D-L-1
10 412
S-D-L-1
410
B-D-L-2
4 610
S-D-L-2
5
2
Time (days) Time (days)
0 0
0 20 40 60 80 100 0 20 40 60 80 100

(a) Long-term beams (b) Long-term slabs


Figure 8-20: Mid-span deflection versus time responses of series E long-term specimens.

Table 8-10: Tension stiffening deflections of series E long-term specimens.

Deflections (mm)
No. of Instantaneous t = 15 days t = 100 days
Specimen db
bars ts(t)/ ts(t)
ms tsi ms ts(t) ms ts(t)
tsi / tsi
Beam 1 2 14 4.63 2.43 6.72 1.55 0.64 9.73 1.02 0.42
Beam 2 4 10 4.49 2.57 6.68 1.59 0.62 9.66 1.09 0.43
Slab 1 4 12 6.99 3.61 10.9 2.31 0.64 17.2 1.48 0.41
Slab 2 6 10 6.62 3.67 10.6 2.23 0.61 16.88 1.46 0.40

8.3.6 Magnitude of creep, ijc – Series F

As discussed previously, creep has a significant impact on the time-dependent behaviour of

reinforced concrete beams. On an uncracked section in bending, tensile creep increases the

tensile stress in the steel bars and reduces the tensile stress in the concrete, hence reducing

tension stiffening. In this test series, four beams and four slabs are considered for both the

short-term and long-term tests, while the same numbers of specimens for the long-term tests.

The only variable parameter is the creep coefficient ijc. In order to adjust the creep

solidification model to obtain the desired ijc, only the value of q4 needs to be varied. For

example, q4 for B-E-S-1 (ijc = 2.0) is taken as 40.0/MPa, and for B-E-S-2 (ijc = 0.5) as

10.0/MPa. All the other parameters are kept constant.


271

Figure 8-21 shows the global moment versus deflection responses for the short-term

specimens in series F. In this case, creep only occurs together with shrinkage prior to loading.

For the short-term beams and slabs under consideration, the initial deflections before loading

are:

Beams: 0.43mm (ijc = 0.5), 0.47mm (ijc = 1.0), 0.51mm (ijc = 1.5), and 0.55mm (ijc = 2.5);

Slabs: 0.96mm (ijc = 0.5), 1.05mm (ijc = 1.0), 1.15mm (ijc = 1.5) and 1.24mm (ijc = 2.5).

Moment (kN.m)
Moment (kN.m)

50 30
c = 0.5
S-E-S-2
c = 1.0
S-E-S-3
40
c = 1.5
S-E-S-4
20 c = 2.0
30 Series1
c = 0.5
B-E-S-2
c = 1.0
B-E-S-3
20 c = 1.5
B-E-S-4
10
c = 2.0
Series2
10
Mid-span deflection (mm) Mid-span deflection (mm)
0 0
0 3 6 9 12 15 18 0 6 12 18 24 30 36

(a) Short-term beams (b) Short-term slabs


Figure 8-21: Instantaneous moment verse mid-span deflection responses of series F short-
term specimens.

It seems that with same amount of shrinkage, increased initial deflections can be induced by

increasing the quantity of creep in the concrete, even though its effect is not very significant.

This is because the instantaneous stress and strain at zero moment is quite small and only

gradually accumulates with increasing shrinkage. The redistribution of stress and strain on the

uncracked sections due to creep is much less significant compared with that caused by

shrinkage. Therefore, the short-term responses of the members are quite similar both before

and after cracking, as seen in Figure 8-21.


272

Figure 8-22 plots tension stiffening deflection versus mid-span moment responses for the

short-term members in series F. In this case, on the fully-cracked sections, only the concrete

in the compression zone carries small stresses and tensile creep is not applicable. Therefore

the initial deflections of the fully-cracked response for the members with various creep

coefficients are similar. The values of instantaneous tension stiffening deflections for the

members in Series F are not significantly affected by creep.


Tension stiffenign deflection (mm)

Tension stiffenign deflection (mm)


4 6

c = 0.5
S-B-S-1 5 c = 0.5
S-B-S-1
3 c = 1.0
S-B-S-2 c = 1.0
S-B-S-2
4
S-B-S-3
c = 1.5 c = 1.5
S-B-S-3
S-B-S-4
c = 2.0
3
S-B-S-4
c = 2.0
2

2
1
1

0 0
0 10 20 30 40 50 60 0 10 20 30
Moment (kN.m) Moment (kN.m)

(a) Short-term beams (b) Short-term slabs


Figure 8-22: Tension stiffening deflection verse mid-span moment responses of series F
short-term specimens.

Figure 8-23 shows the mid-span deflection versus time responses of series F long-term

specimens. The beams and slabs were loaded up to 25.8kN.m and 10.7kN.m then held

constant for 100 days. Clearly the members with increased creep during the sustained loading

also exhibit increased long-term deflections. However, the amount of calculated time-

dependent tension stiffening deflections ts(t) as given in Table 8-11 seem to be little affected

by concrete creep. The increase of actual deflection due to creep is largely due to the increase

of curvatures on the cracked sections. Despite the fact that the relative increase of curvature

on a cracked section due to creep is less than that on an uncracked section, the magnitude of

curvature on the cracked section is much larger and dominates. On the uncracked sections in-
273

between the cracks, creep increases the depth of compressive zone and reduce the area of

concrete in tension, meanwhile the total tensile force increases due to the shortening of lever

arm. This may leads to an increase of tensile stress in the concrete and in the tensile steel bars.

However bond deterioration and tensile creep in the concrete break down the concrete tensile

stress with time which cancel out the effect of compressive creep. Therefore, the quantity of

creep has only a very minor impact on the long-term behaviour of tension stiffening.

25 c = 0.5
S-E-L-2
12 c = 0.5
B-E-L-2
Mid-span deflection (mm)

c = 1.0
B-E-L-3 Mid-span deflection (mm) c = 1.0
S-E-L-3
c = 1.5
S-E-L-4
10 c = 1.5
B-E-L-4 20
c = 2.0
Series2
c = 2.0
Series2
8
15

6
10
4
5
2
Time (days)
Time (days)
0
0
0 20 40 60 80 100
0 20 40 60 80 100

(a) Long-term beams (b) Long-term slabs


Figure 8-23: Mid-span deflection versus time responses of series F long-term specimens.

Table 8-11: Tension stiffening deflections of series F long-term specimens.

Deflections (mm)
Instantaneous t = 15 days t = 100 days
Specimen ijc
ts(t)/ ts(t)
msi tsi ms ts(t) ms ts(t)
tsi / tsi
Beam 1 2.0 3.69 1.99 5.58 1.27 0.64 8.49 0.75 0.38
Beam 2 0.5 3.77 1.92 4.91 1.34 0.70 6.90 0.87 0.45
Beam 3 1.0 3.69 2.00 5.21 1.24 0.62 7.52 0.76 0.38
Beam 4 1.5 3.80 1.89 5.46 1.18 0.63 8.06 0.70 0.37
Slab 1 2.0 6.99 3.61 10.90 2.31 0.64 17.2 1.48 0.41
Slab 2 0.5 7.02 3.58 9.47 2.40 0.67 13.9 1.33 0.37
Slab 3 1.0 7.03 3.57 10.02 2.32 0.65 15.1 1.33 0.37
Slab 4 1.5 6.95 3.66 10.44 2.33 0.64 16.1 1.40 0.38
274

8.3.7 Magnitude of shrinkage, sh – Series G

As previously mentioned, shrinkage has a significant effect on both the short-term and long-

term behaviour of tension stiffening. Series G of the numerical experiments examines this

effect. Four identical beams and four identical slabs are chosen for both the short-term and

long-term numerical tests, with the only difference as the level of shrinkage. The design

details are given in Table 8-3 and 8-4. The numerical way to modify the values of shrinkage

strain in the finite element models consists in replacing the values of Ash in Equation 6.41. For

example, the final shrinkage strains for Beam 1 and Beam 2 are -400  and -200 , the

corresponding Ash are -960  and -480 .

As for the short-term analysis, since shrinkage only occurs before loading and bond

deterioration due to shrinkage is not significant, the long-term decay parameter of bond stress

Ȝ4 in Equation 6.64 is not applicable (taken as 1.0) in the finite element models. The global

moment versus deflection responses are plotted in Figure 8-24. The curves in the figure are

generally parallel with each other. At the same moment level, the member with greater initial

shrinkage exhibits greater deflection. This is attributed to shrinkage-induced initial deflection

in each specimen. An increase of shrinkage prior to loading results in a larger initial deflection

and a smaller cracking moment. The initial deflections of the short-term beams and slabs in

this series are as follows:

Beams: 0.30mm (sh = -200 ), 0.42mm (sh = -300 ), 0.55mm (sh = -400 ), and 0.67mm

(sh = -500 );

Slabs: 0.62mm (sh = -200 ), 0.93mm (sh = -300 ), 1.24mm (sh = -400 ), and 1.55mm

(sh = -500 ).


275

Moment (kN.m)
50
Moment (kN.m)

30

40

20
30
B-F-S-2
sh = 200
S-F-S-2
sh = 200
B-F-S-3
sh = 300 
20 S-F-S-3
sh = 300 
sh = 400 
Series2 10 Series1
sh = 400 
Series4
sh = 500 
 sh = 500 
Series4
10
Mid-span deflection (mm) Mid-span deflection (mm)
0 0
0 3 6 9 12 15 18 0 6 12 18 24 30 36

(a) Short-term beams (b) Short-term slabs


Figure 8-24: Instantaneous moment verse mid-span deflection responses of series G short-
term specimens.

The fully-cracked responses after loading are also parallel with each other as the other

parameters (such as elastic modulus) remain the same during the short-term loading period.

Figure 8-25 plots the tension stiffening deflections at different loading levels for the series G

short-term specimens. The results demonstrate that before cracking, the magnitude of tension

stiffening of the member is increased with increasing initial shrinkage. However, as the

moment increases and cracks start to form, the significance of initial shrinkage on tension

stiffening deflection becomes less obvious. At the descending branches of the curves in Figure

8-25, it appears that tension stiffening deflection is in fact greater in the members with less

initial shrinkage.
276

Tension stiffenign deflection (mm)


Tension stiffenign deflection (mm)

4 6

 shc=
= 200
0.5
Series1 5  shc=
= 200
0.5
Series1
3  shc=
= 300
1.0 
Series2  shc=
= 300
1.0 
Series2
4  sh
 shc=
= 400
1.5 
Series3 c=
= 400
1.5 
Series3
Series4
 shc=
= 500
2.0   Series4
shc =
= 500
2.0 
2 3

2
1
1

0 0
0 10 20 30 40 50 60 0 10 20 30
Moment (kN.m)
Moment (kN.m)

(a) Short-term beams (b) Short-term slabs


Figure 8-25: Tension stiffening deflection verse mid-span moment responses of series G
short-term specimens.

The analytical results indicate that for the specimens with significant initial shrinkage, such as

Beam 4 and Slab 4 (İsh = -500 ), tension stiffening and its contribution to the total stiffness

of the specimens are less than the others at higher bending moments. A careful examination of

the tensile strain contours (Figure 8-26a) shows that these specimens actually have more

primary cracks formed, which leads to a further decrease of tension stiffening. This is because

the restraint to initial shrinkage by the steel bars has induced significant tensile stress on the

concrete at the steel bar level before loading. As illustrated by Figure 8-26b, the maximum

tensile stress at the mid span section of Slab 4 (due to shrinkage and self-weight) prior to

loading is 1.75MPa, which is quite close to the cracking stress (2.0MPa). It is possible that

with more shrinkage, cracking can even occur before loading. Comparatively the maximum

tensile stress on the bottom side of Slab 2 is only about 0.7MPa.


277

Distance from the bottom of the section (mm)


160

120

Slab 2 (İsh = -200 ) 80

Slab 2 Slab 4

40

Steel bar level

Slab 4 (İsh = -500 ) 0


-1 -0.5 0 0.5 1 1.5 2
Concrete stress (MPa)

(a) Tensile strain contours at 20.8kN.m (b) Concrete stress at mid-span before loading
Figure 8-26: Mid-span deflection versus time responses of series G long-term specimens.

It has been mentioned that shrinkage is the major factor for the long-term increase of

deflection and loss of tension stiffening. Figure 8-27 shows the development of mid-span

deflections of series F long-term specimens. The specimens were subjected to the same

constant loads as in series E tests. The mid-span deflections are very dependent on the total

amount of shrinkage developed during the sustained loading period. The larger the quantity of

shrinkage is, the greater the final deflection. The decay of long-term tension stiffening of the

specimens is given in Table 8-12. The decrease in tension stiffening deflections ts(t) depends

on the quantity of shrinkage. For instance, ts(t) of Beam 1 (İsh = -400 ), eventually drops to

0.38 of its instantaneous value tsi, while for Beam 4 (İsh = -500 ) ts(t) / tsi drops to 0.26.

When the developed shrinkage strain is less than -400 , the effect of shrinkage on the decay

of tension stiffening is not so great. This is because with sh = -500 , more primary cracks

form by the end of the tests, which leads to significant decrease of tension stiffening with time.

Since the proposed bond model relates the bond stress to shrinkage, bond stress decreases

more rapidly in the specimens with greater amount of shrinkage. Thus tension stiffening may
278

also decrease more rapidly in these specimens.

B-F-L-2
sh = 200 25  sh = 200
S-F-L-2
12
Mid-span deflection (mm)

Mid-span deflection (mm)


 sh = 300   sh = 300 
S-F-L-3
B-F-L-3
10 Series2  sh = 400 
Series2
sh = 400  20
Series4  sh = 500 
Series4
sh = 500 
8
15
6
10
4
5
2
Time (days) Time (days)
0 0
0 20 40 60 80 100 0 20 40 60 80 100

(a) Long-term beams (b) Long-term slabs


Figure 8-27: Mid-span deflection versus time responses of series G long-term specimens.

Table 8-12: Tension stiffening deflections of series G long-term specimens.

Deflections (mm)
Instantaneous t = 15 days t = 100 days
Specimen sh ()
ts(t)/ ts(t)/
msi tsi ms ts(t) ms ts(t)
tsi tsi
Beam 1 -400 3.69 1.99 5.58 1.27 0.64 8.49 0.75 0.38
Beam 2 -200 3.65 2.03 5.48 1.17 0.58 7.63 0.83 0.41
Beam 3 -300 3.72 1.97 5.54 1.20 0.61 8.06 0.79 0.40
Beam 4 -500 3.81 1.87 5.82 1.10 0.59 9.14 0.49 0.26
Slab 1 -400 6.99 3.61 10.90 2.31 0.64 17.20 1.48 0.41
Slab 2 -200 6.86 3.74 10.46 2.35 0.63 15.06 1.88 0.50
Slab 3 -300 7.12 3.48 10.85 2.25 0.65 16.16 1.68 0.48
Slab 4 -500 7.13 3.47 11.37 2.04 0.59 18.65 0.90 0.26

8.3.8 Compressive reinforcement area, Asc – Series H

The analysis on the experimental data provided by Gribniak (2009) in Chapter 7 shows that

the provision of compressive reinforcement can reduce the tensile stress due to shrinkage in

the tensile concrete and hence reduce the reduction in the cracking moment caused by early

shrinkage. Therefore, compressive reinforcement can contribute to the stiffness of the


279

specimens under short-term loading. The beams in series H were designed with different

quantities of compressive reinforcement Asc. The details of the design parameters are given in

Table 8-3. Four of the specimens were under short-term loading after being dried for 100 days

(to allow shrinkage strain to accumulate to -400×10-6 and the other four were under sustained

long-term loading (M = 25.8kN.m) for 100 days.


Moment (kN.m)

50
Initial
40 Specimen deflections
(mm)
30 Beam 1 0.24
2
Beam 1 (A sc = 226mm )
B-G-S-1
2
20 Beam 2 (A sc = 402mm )
B-G-S-2 Beam 2 0.05
2
B-G-S-3
Beam 3 (A sc = 509mm )
10 B-G-S-4
Beam
2
4 (A sc = 628mm ) Beam 3 -0.05

Mid-span deflection (mm) Beam 4 -0.15


0
0 3 6 9 12 15

Figure 8-28: Instantaneous moment verse mid-span deflection responses of series H short-
term specimens.

Figure 8-28 demonstrates the moment versus mid-span deflection responses for the short-term

tested specimens in series H. Test specimens with heavy compressive reinforcement (for

example Beam 4, Asc = 628mm2) exhibit less initial deflection (or even upwards deflections)

and larger cracking moments. Furthermore, the post-cracking deflections of these specimens

are also smaller compared with the ones reinforced with light compressive steels. The method

to predict short-term deflection of beams in the latest revision of Australia Standard AS3600-

2009 has taken into account the influence of compressive reinforcement by reducing the

shrinkage-induced tensile stress according to the compressive reinforcement area.


280

Figure 8-29 demonstrates the tension stiffening deflection versus mid-span moment responses

for the short-term specimens in this series. Before cracking, the magnitude of tension

stiffening deflection varies for the beams with different compressive reinforcement area.

However, there is no regular pattern that can be observed to see how tension stiffening

deflection is affected by compressive reinforcement prior to cracking. After cracking, the

difference of tension stiffening deflection among the specimens tends to be less significant, as

can be seen in Figure 8-29.


Tension stiffening deflection (mm

4 Beam 1 (A sc = 226mm )
2
B-G-S-1
 sh = -400
2
Beam 2 (A=sc-200
B-G-S-2 = 402mm )
sh
2
3 Beam 3 sh
B-G-S-3 (A=- = 509mm )
sc 300
2
B-G-S-4
Beam 4 sh
(A=- = 628mm )
sc 500

0
0 10 20 30 40 50 60
Mid-span moment (kN.m)

Figure 8-29: Tension stiffening deflection verse mid-span moment responses of series H
short-term specimens.

Figure 8-30 shows that the long-term deflections of the beams and how they are affected by

the provision of compressive reinforcement. The beams and slabs were under the same

constant loads as those in the previous test series. It can be seen that the long-term

development of deformation is resisted by the provision of compressive steel. This is

attributed to the fact that compressive reinforcement can restrain creep in the compressive
281

zone and reduces the effect of shrinkage warping, and hence increases the time-dependent

stiffening of the beams. Table 8-13 demonstrates that with increased quantities of compressive

steel, the loss of tension stiffening deflection with time is reduced. This is especially the case

for Beam 3 and Beam 4. Since the top compressive reinforcement can help reduce the tensile

stress developed on the bottom of the beam due to shrinkage, the shrinkage-induced cracking

is also reduced. Therefore, the loss of tension stiffening in the long-term is less significant in

these specimens.

2
Beam 1 (A sc = 226mm )
B-G-L-1
12
Mid-span deflection (mm)

2
Beam 2 (A sc = 402mm )
B-G-L-2
2
10 B-G-L-3
Beam 3 (A sc = 509mm )
2
B-G-L-4
Beam 4 (A sc = 628mm )
8

2
Time (days)
0
0 20 40 60 80 100

Figure 8-30: Mid-span deflection versus time responses of series H long-term specimens.

Table 8-13: Tension stiffening deflections of series H long-term specimens.

Deflections (mm)
Asc Instantaneous t = 15 days t = 100 days
Specimens
(mm2) ts(t)/ ts(t)/
msi tsi ms ts(t) ms ts(t)
tsi tsi
Beam 1 226 3.56 2.01 5.17 1.29 0.64 7.41 0.71 0.35
Beam 2 402 3.58 1.92 4.94 1.32 0.69 6.75 0.83 0.43
Beam 3 509 3.47 1.99 4.71 1.46 0.73 6.28 1.05 0.53
Beam 4 628 3.57 1.85 4.71 1.35 0.73 6.18 0.90 0.49
282

8.3.9 Quantity of transverse reinforcement – Series I

The influence of the quantity of transverse reinforcement on the time-dependent behaviour of

tension stiffening is examined in the Series I tests. The idealization of the transverse

reinforcement in the finite element models is the one-dimensional truss element that shares

the same nodes with the connected concrete elements. The developments of mid-span

deflections in both the short-term and long-term models are shown in Figure 8-31 and 8-32. It

seems that the influence of transverse reinforcement on the deflections on the beams is very

minimal. Since the calculated fully-cracked response of the beams is only dependent of the

area of tensile and compressive reinforcement, the difference of tension stiffening deflections

among the beams is small. It may be concluded that the inclusion of transverse reinforcement

does not induce any unpredictable effects on the behaviour of time-dependent tension

stiffening.
Moment (kN.m)

50

40

30
B-H-S-1
Beam 1 (No stirrup)
20 B-H-S-2
Beam 2 (2 8@100)
B-H-S-3
Beam 3 (2 10@100)
Beam 4 (2 12@100)
B-H-S-4
10

Mid-span deflection (mm)


0
0 3 6 9 12 15

Figure 8-31: Instantaneous moment verse mid-span deflection responses of series I short-
term specimens.
283

12

Mid-span deflection (mm)


Beam 1 (No stirrup)
B-H-L-1
B-H-L-2
Beam 2 (2 8@100)
10 B-H-L-3
Beam 3 (2 10@100)
B-H-L-4
Beam 4 (2 12@100)
8

2
Time (days)
0
0 20 40 60 80 100

Figure 8-32: Mid-span deflection versus time responses of series I long-term specimens.

8.4 SUMMARY
The effects of a variety of parameters on time-dependent tension stiffening have been

investigated by means of numerical experiments conducted with the finite element program.

Night series of tests have been undertaken, with one specific parameter being valued and

investigated in each series. The finite element program is able to predict the tension stiffening

effect and is ideal to generate the large amount of numerical results presented herein. Tension

stiffening has been conveniently represented by the tension stiffening deflection, as the

difference between the calculated mid-span deflection and the calculated fully-cracked

deflection.

It has been concluded that certain parameters, such as concrete tensile strength, concrete

elastic modulus, tensile reinforcement area and compressive reinforcement area, can

substantially affect the magnitude of tension stiffening. The magnitude of tension stiffening is

larger in members with less tensile reinforcement area or reinforcing ratio, but its rate of

change under long-term loading seems little affected by this parameter. Tension stiffening is
284

also sensitive to the concrete tensile strength. The greater the concrete strength is, the greater

the contribution of concrete to the global stiffness of the members. Moreover, the long-term

rate of change of tension stiffening decreases as the tensile strength of concrete increases. It

has been concluded that a increase in concrete elastic modulus increases the stiffness of a

reinforced concrete specimen and the rate of change of tension stiffening for the specimens

with larger concrete elastic modulus under long-term loads is greater. The provision of

compressive reinforcement can reduce the deflection of the structure by reducing the effect of

shrinkage warping on the structure. It also reduces the maximum shrinkage-induced tensile

stress that can cause further cracking and helps resist the decay of tension stiffening with time.

Tension stiffening can also be improved by providing larger bottom concrete cover. The study

demonstrates that the rate of change of tension stiffening deflection in a reinforced concrete

beam decreases as the bottom cover increases, although this effect was not observed in the

case of slabs.

The influence of tensile bar diameter and transverse reinforcement on time-dependent tension

stiffening is not significant. The numerical results indicate that very minimal change of

tension stiffening can be observed with either a change of tensile bar diameter or the amount

of transverse reinforcement.

Creep and shrinkage are the main parameters that vary with time. Shrinkage is the most

significant parameter that affects time-dependent tension stiffening. Shrinkage warping of the

flexural members results in initial deflection prior to loading and a decrease in the cracking

moment. The specimens with larger initial shrinkage exhibited slightly less tension stiffening

deflection under short-term loading. In the long-term, new primary cracks due to the restraint

to shrinkage are regarded as the major factor for the loss of tension stiffening. The larger the
285

shrinkage strain, the larger the loss of tension stiffening. Creep on the other hand has

relatively minor impact on the short-term and long-term behaviours of tension stiffening,

compared to shrinkage. According to the numerical results, the short-term and long-term

tension stiffening deflections of the specimens with different amount of creep are not

significantly different.
CHAPTER 9 CONCLUSIONS AND
RECOMMENDATIONS

9.1 SUMMARY
The present study aims to contribute to a theoretical understanding of tension stiffening in

reinforced concrete structures and its variation with time. It has undertaken experimental and

numerical studies to achieve this objective. Tension stiffening under both long-term and short-

term loading has been investigated. Shrinkage and creep, of the concrete have been taken into

the account in the analysis. These time-dependent effects have been largely ignored in the

previous studies of tension stiffening. Numerical works were undertaken using the finite

element program RECAP (Foster 1992, Foster and Gilbert 1990, Foster and Marti 2003). A

bond model has been proposed so as to simulate the effect of bond deterioration and its effect

on tension stiffening. The finite element program was used to model laboratory specimens

tested by the author and others. To first demonstrate the versatility of the finite element

program and demonstrate other possible parameters that may affect time-dependent tension

stiffening, a number of numerical experiments were conducted and the results were analysed

and discussed.
287

9.2 CONCLUSIONS
The mechanisms and factors affecting tension stiffening have been investigated in this study

and the results have been carefully presented. It is been demonstrated that initial shrinkage

before loading has significant effects on the structural behaviour. For uniaxial tension tests,

shrinkage causes initial shortening of the member, while for flexural tests, shrinkage induces

initial deflection. Ignoring this deformation prior to loading can lead to misinterpretation of

the experimental results. The experimental program conducted in this research carefully

examines the effects of initial shrinkage prior to loading, and the results are significant.

Before cracking, the restraint to shrinkage by the reinforcing bars produces significant tensile

force in the concrete. Therefore, members with initial shrinkage had much lower cracking

loads than the members without shrinkage. It has also been demonstrated by the experiments

that after cracking initial shrinkage have relatively minor influence on tension stiffening under

increasing loads. Nevertheless, for the flexural members considered in the numerical

experiments, it seems that initial shrinkage can actually reduce the tension stiffening

deflection after cracking. This is caused by the restraining force to shrinkage that is applied to

the concrete which can trigger additional cracks under the increasing moment. For the short-

term uniaxial tension members, larger crack width was observed on the members with initial

shrinkage. However, this is not the case for the short-term flexural members, where crack

width was not significantly influenced by initial shrinkage. It is also observed that shrinkage

prior to loading may induce more cracks and reduce crack spacing, which has been found in

both the laboratory and numerical experiments of the flexural members.

Compared with shrinkage, creep has a relatively minor effect on short-term behaviour of

tension stiffening. This is observed in the numerical experiment. The short-term specimens
288

with different initial creep strains exhibited almost identical behaviour before and after

cracking. Creep before loading can slightly increase the initial deflection triggered by

shrinkage warping. However, after cracking the significance of creep on the behaviour of a

short-term specimen is very limited.

In the long-term tests where the load was maintained constant after cracking of the member,

the experimental results show that the contribution of the tensile concrete decreases with time.

The analysis of the long-term slab tests indicates that tension stiffening dropped off

dramatically in the first 15 days under loading, and then it decreased at a slower rate thereafter.

The test results also indicate that the long-term decreasing rate of tension stiffening is not

remarkably affected by the loading levels. A similar decreasing rate pattern was found from

the analysis of earlier test results (Nejadi and Gilbert 2004). The restraint to shrinkage by the

reinforcing bars can induce the time-dependent formation of primary cracks and cover-

controlled cracks, and hence is regarded as the major mechanism for the loss of tension

stiffening with time. On the other hand, in the uniaxial tension tests, tensile creep can shed the

tensile stress carried by the intact concrete to the steel bars and cause a gradual softening of

the load-deformation response. In flexural members, compressive creep causes the lowering

of neutral axis and slightly increases of total tensile force carried by the tensile steel bars.

A bond model has been proposed and incorporated into the finite element program. The bond

model was calibrated using the results obtained from the experimental program. In the finite

element models, concrete, steel and bond mechanism are represented by three different types

of elements. Creep in the concrete is modeled in a rate-type shape by incorporating the

solidification creep theory (which had previously been developed in RECAP) in terms of a

series of Kevin chain units. Shrinkage was introduced by a simple age-related function. Both
289

the creep and shrinkage strains were treated as inelastic strains in the finite element analysis.

The degradation of tension stiffening due to cracking is mainly characterized by a non-local

bond stress-slip relationship, in which bond stress is not only dependent on slip but also

concrete damage and steel stress states. For the long-term analysis, bond stress in the

proposed bond model also decreases with respect to the amount of shrinkage, as shrinkage is

considered as the major mechanism to cause long-term bond damage. The finite element

program was then adopted to simulate the experimental works and gave excellent results in

terms of both global and local responses compared with the experimental responses.

Finally, the parametric study has led to the following extra conclusions:

i. The time-dependent behaviour of tension stiffening can be greatly affected by the

amount of tension reinforcement area or the reinforcing ratio. The numerical study has

shown the tension stiffening contribution is more significant in the light-reinforced

members both under short-term and long-term loadings. With time, the decreasing rate

of tension stiffening does not vary too much among the members with different tensile

reinforcement area.

ii. The provision of compressive reinforcement can give rise to a stiffer response for the

beams by reducing the curvature resulting from restraint to shrinkage by the tensile

reinforcing bars and reduces the tensile stresses in the cracked concrete and the

development of bond degradation. Therefore it enhances both the short-term and long-

term tension stiffening effects.

iii. The tensile strength of concrete has a direct impact on time-dependent tension

stiffening. The increased tensile strength of concrete can also improve the capacity of
290

concrete to carry tension and resist shrinkage-induced cracking. Thus both short-term

and long-term tension stiffening are increased in the concrete with higher tensile

strength. For the long-term specimens, the rate of change of tension stiffening seems

improved by this parameter as well.

iv. The long-term rate of change of tension stiffening is dependent on the elastic modulus

of concrete. Namely the larger the concrete elastic modulus, the greater the rate is.

Moreover, the time-dependent tension stiffening deflection decreases as the elastic

modulus of concrete increases.

v. The bottom cover of concrete can increase the amount of tension stiffening in the

reinforced concrete beams. It also reduces the rate of change of tension stiffening

under sustained loading, especially for the beam specimens.

vi. Certain other parameters, such as transverse reinforcement, tensile bar diameter, etc.

have very limited effects on time-dependent tension stiffening.

The long-term decay of tension stiffening is a significant factor that should be considered in

the structural design, since stiffness of the structure is often the governing design criteria.

Shrinkage is the major factor leading to the loss of tension stiffening with time. In addition,

early shrinkage before loading can also contribute to the long-term increase of total deflection.

Thus the design approach that is used to predict long-term stiffness of reinforced concrete

beams and slabs must incorporate shrinkage. In Appendix A, an effective stiffness based

method is proposed by the author to predict the short-term and long-term stiffness of a

uniaxial tension prism. The method is based on the formula given by Eurocode 2 (1992) but

modified by incorporating shrinkage and creep effects to consider the time-dependent


291

stiffness of a cracked reinforced concrete element. A number of tension prisms presented in

Chapter 4 were chosen to be analysed by this method and good correlation has been observed.

The method may then be extended for calculating the time-dependent stiffness of beams or

slabs.

9.3 RECOMMENDATIONS FOR FUTURE


RESEARCH
Future research into the following areas is recommended:

i. Additional experimental short-term tests of reinforced concrete prisms or flexural

members with different amount of initial shrinkage are required to confirm the

conclusions of the significance of shrinkage on the time varying behaviour of

reinforced concrete under service loads.

ii. Experimental work could also be undertaken to investigate the time-dependent tension

stiffening in two-way slabs under serviceability conditions. Numerical work could be

undertaken to extend the 2-D finite element model to cover 3-D simulations for the

analysis of tension stiffening. The three-dimension effects (such as concrete

confinement and radial stress) on the bond mechanism can be more accurately

modeled in this case.

iii. Further numerical studies to investigate the influence of different reinforcement types

(or example FRP bar, plain bar, etc.) on tension stiffening could be conducted. The

bond stress-slip relationship will need to be calibrated for different bar types.
292

iv. Various other parameters (loading history, bar surface characteristics, boundary

conditions, etc.) on time-dependent tension stiffening could be assessed.

v. The proposed methods to predict long-term deflection or elongation of uniaxial

tension prisms given in Appendix A considering both shrinkage and creep can be

further simplified and extended to the analysis of beams and slabs in practice,

culminating in the development of reliable and rational approaches for the analysis

and design of reinforced concrete at the serviceability limit states.


APPENDIX A AN ANALYTICAL METHOD TO
PREDICT TIME-DEPENDENT TENSION STIFFENING
IN A UNIAXIAL TENSION PRISM
An analytical method to predict time-dependent stiffness of a uniaxial tension member under

sustained loads has been proposed by Gilbert and Wu (2009). Recommendations are made for

including tension stiffening in the design of reinforced concrete elements at the serviceability

limit states. The method is presented here and might be further extended to be applied for the

analysis of flexural members under sustained moments.

Before cracking

Consider the uniaxial loaded concrete prism shown in Figure 2-1. When the applied load P is

less than the time-dependent cracking load, i.e. P < Pcr.sh, the member remains uncracked

throughout. Using the age-adjusted effective modulus method, the average member strain at

time t, s.avg(t) and the steel and concrete stresses in the uncracked section (s(t) and c(t),

respectively ) were given by (Gilbert 1988) as follows:

P § nU · § n*  n · H sh
H s ,avg (t ) ¨¨ ¸¸¨¨1  ¸¸ + A.1
© 1  nU ¹© n (1  n U ) ¹ 1  n U
* *
E s As

Vs t E s H s.ave t and Vc t P / Ac  V s t U A.2

where sh is the shrinkage strain that occurs before time t; is the reinforcement ratio (As /

Ac); n is the modular ratio (Es / Ec), n* is the effective modular ratio (Es / Ee); n * is the age-

adjusted effective modular ratio ( E s / E e );Es and Ec are the elastic moduli of the steel and

concrete, respectively; Ee is the effective modulus of concrete given E e E c / 1  M ; E e is

the age-adjusted effective modulus of the concrete given by E e E c / 1  FM ; is the creep


294

coefficient for concrete for the time interval 't (between first loading and time t); and F is an

aging coefficient used to account for the gradual reduction in the creep coefficient with age

when the stress is gradually applied over the time interval 't and is often approximated by

F 0.8 (Gilbert 1988).

The shrinkage induced tensile stress in the concrete on the uncracked cross-section ( V sh 0 ) is

obtained by setting P = 0kN in Equations A.1 and A.2 and the reduced cracking load after

shrinkage (Pcr.sh) is reached when c(t) in Equation A.2 equals the tensile strength of the

concrete f’t. Noting that sh is negative in the context:

H sh E s U
V sh 0  and Pcr . sh Ac (1  nU )( f c  V sh 0 ) A.3
1  n *U

After cracking

Cracking and cracking induced deterioration of bond caused by slip at the concrete-steel

interface are the physical mechanisms that are primarily responsible for the reduction in

tension stiffening under either increasing load or due to restraint to shrinkage and temperature

differentials. Without cracking, incremental increases in stress caused by load or restraint

would cause an increase in tension stiffening, but the formation of new cracks and the

resulting bond slip results in a subsequent overall (net) drop in tension stiffening.

The average strain caused by the applied load P in a cracked tension member can be estimated

using an effective rigidity (EcAe), which is between the rigidities of the fully cracked section

ignoring the concrete (EcnAs) and of the uncracked section (EcAt). The area At is the

transformed area of the uncracked cross-section (Ac+nAs). The approach is similar to that
295

outlined in Eurocode 2 (1992), but modified to include the effects of time-dependent slip at

the concrete-steel interface caused by shrinkage and by time-dependent cracking. The average

strain caused by load in an axially loaded tension member after cracking at time t may be

approximated by

H s.avg ]H s.cr  1  ] H s.uncr A.4

where s.cr is the steel strain at a crack, i.e. ignoring the concrete in tension; s.uncr is the steel

strain on the uncracked transformed section; and  is a coefficient accounting for load level,

the degree of cracking and the amount of shrinkage. For deformed bars under short-term

loading, Eurocode 2 (1992) suggest that  = 1 – (Pcr / P)2. However, to provide better

agreement with the instantaneous and time-dependent test data obtained in this study and

reported subsequently, the equation for  has been modified as follows:

D
] 1  Pcr .sh / P A.5

where Pcr..sh is the load required to cause cracking, suitably modified to account for the tension

induced by restraint to shrinkage (and given by Equation A.3) and the index depends on the

level of shrinkage that has occurred prior to time t (sh) and may be taken as

D 2.5 when H sh d 100 u 10 6 A.6a

D 4 .0 when H sh ! 300 u 10 6 A.6b

If the magnitude of shrinkage strain is between 100 and 300 microstrain, then may be

obtained by linear interpolation.


296

In terms of an effective cross-sectional area, Ae, the average strain caused by load at time t is

obtained from equation A.4 as:

P P P
(H s.avg ) i ]  (1  ] ) A.7
E c Ae E c nAs E c At

Substituting Equation A.5 into A.7 and rearranging gives:

nAs
Ae D
§ nA ·§ P · A.8
1  ¨¨1  s ¸¸¨ cr . sh ¸
© At ¹© P ¹

If the member has been subjected to drying shrinkage, (sh), the initial compressive strain on

the uncracked section ( H sh / 1  n * U must be added to the average strain caused by load to

find the total average member strain (s.avg). That is:

P H sh
H s.avg  A.9
E c Ae (1  n * U )

At any load P > Pcr.sh, the instantaneous tension stiffening strain tsi is the difference between

the strain in the bare bar ( P / AsEs) and the average member strain, s.avg (see Figure 3-1):

P P H sh
H tsi (  ) A.10
E s As E c Ae (1  n * U )

At each crack the reinforcement carries the entire load P, but the average force carried by the

reinforcement over a gauge length containing several cracks is


297

Ns E s H s.avg A.11

and the average force in the concrete and the average concrete stress are therefore

Nc P  Ns and Vc N c / Ac A.12

When the concrete on a cracked member shrinks by an amount sh over a time period t, the

restraint provided by the reinforcement introduces additional tensile stress in the concrete

between the primary cracks. However, this shrinkage induced tension causes additional time-

dependent cracking, and time-dependent slip at the concrete-steel interface and, consequently,

a net reduction in tension stiffening with time. To model the increase in cracking with time,

Eurocode 2 (1992) suggests that, for long term loading, the tension stiffening strain reduces to

50% of its initial value (i.e. 0.5tsi), irrespective of the level of loading. In reality, early

shrinkage causes a reduction in the cracking load and this affects the value of ] obtained

from Equation A.5 and consequently the value of Ae given by Equation A.8.

Working Examples

In the tests conducted by the author, two of the specimens (STN12 and STS12) were tested

under short-term monotonically increasing load (the short-term tests), while the third

specimen (LTN12A) was loaded to a typical service load level and then subjected to a

constant sustained loads for a period of about 50 days (the long-term test).

Short-term tests

Specimen STN12 was subjected to a monotonically increasing load soon after the end of

moist curing. The elastic modulus of concrete at the time of testing at age 32 days was Ec =
298

22400 MPa (and with Es = 200000 MPa, n = 8.93) and the tensile strength of concrete was ft

= 2.04 MPa. Prior to loading, the measured drying shrinkage strain was -28 x 10-6. The creep

coefficient associated with this initial period of shrinkage was M = 0.3 and therefore E e =

18060 MPa ( n * = 11.1). From Equation A.3, the initial concrete tensile stress caused by

restraint to shrinkage prior to loading and the reduced cracking load are Vsh0= 0.06 MPa and

Pcr.sh = 21.6 kN. From Equation A.6, D= 2.5.

Specimen STS12 was identical to specimen STN12 except that it was tested 3½ weeks later at

age 57 days when the drying shrinkage had increased to -249 x 10-6. For STS12, at the time of

testing at age 57 days, Ec = 21600 MPa (i.e. n = 9.26) and ft = 2.15 MPa. The creep coefficient

associated with the initial period of shrinkage was M = 1.13 and therefore E e = 11350 MPa

and n * = 17.6. From Equation A.3, the initial concrete tensile stress caused by restraint to

shrinkage prior to loading and the reduced cracking load are Vsh0= 0.47 MPa and Pcr.sh = 18.3

kN. From Equation A.6, D= 3.62.

The average strain versus applied load curves measured using LVDTs throughout the short-

term tests are plotted in Figure A-1, together with the theoretical curves obtained using

Equation A.9. The agreement between the experimental and theoretical curves is excellent.
299

70
Pcr = 21.5 kN (experimental)
60 = 21.6 kN (theoretical – Eq. A.3)
Axial Load (kN)

Experimental
50 Hts
Theoretical
40 Bare bar P (kN) Hs.avg (exp) Hs.avg (theory)
30 33.9 1097x10-6 1031x10-6
40.0 1531x10-6 1401x10-6
Pcr 50.0 2083x10-6 1941x10-6
20
10 -6
Average Axial Strain (x10 ) (Hs.avg)
0
-500 0 500 1000 1500 2000 2500 3000
-25
(a) STN12

70
Pcr = 13.0 kN (experimental)
60 = 18.3 kN (theoretical – Eq. A.3)
Axial Load (kN)

50 Experimental
Theoretical
40 P (kN) Hs.avg (exp) Hs.avg (theory)
Bare bar 33.9 1098x10-6 1143x10-6
-6
30 40.0 1455x10 1466x10-6
-6
50.0 1994x10 1951x10-6
20
10
Average Axial Strain (x10 )
-6 (Hs.avg)
0
-500 0 500 1000 1500 2000 2500 3000
-209
(b) STS12

Figure A-9-1: Axial load versus average axial strain for STN12 and STS12.

Table A-1: Experimental and theoretical average concrete stresses after cracking

Applied Load, P Average concrete stress (MPa) Applied Load, P Average concrete stress (MPa)
on STN12 (kN) Experimental Theoretical on STS12 (kN) Experimental Theoretical
40 0.84 0.84 40 0.84 0.69
50 0.65 0.62 50 0.81 0.60

From the steel strain gauge readings, the average steel stress was obtained, and knowing the

applied load at each instant, the average concrete stress was determined for each specimen.

The measured and theoretical concrete stresses are compared in Table A-1 for several values

of applied load after the primary crack pattern had developed. Agreement is considered to be

good.
300

Long-term tests

Specimen LTN12A was moist cured until two days before the test started (at age 33 days).

The specimen was initially loaded in 5 kN increments up to a load P = 40 kN. The load was

then sustained for a period of 50 days (i.e. from age 33 days to age 83 days). The measured

elastic modulus of concrete at the time of first loading was Ec = 22400 MPa (n = 8.93) and the

tensile strength of concrete was ft = 2.04 MPa. The measured shrinkage strain in the concrete

just before first loading was -52 x 10-6 and the creep coefficient for this initial period of

drying (up to age 33 days) was 0.3 (i.e. E e = 18060 MPa, n * = 11.1). From Equation A.3, the

initial concrete tensile stress caused by restraint to shrinkage prior to loading and the reduced

cracking load are Vsh0= 0.11 MPa and Pcr.sh = 21.1 kN.

From Equation A.6, D= 2.5 and the theoretical load verses average strain curve at first loading

is shown in Figure A-2, together with the experimental curve obtained at first loading.

Agreement at the time of first loading is again excellent. Immediately after the application of

P = 40.0 kN, the measured average axial strain was 1409 x 10-6 while the calculated value was

1400 x 10-6.

The shrinkage strain at the end of the sustained load period was -310 x 10-6. The creep

coefficient associated with the sustained load period was 1.38. Therefore for the period

between age 33 days and age 83 days, E e = 10650 MPa ( n * = 18.8). From Equation A.3, the

concrete tensile stress caused by restraint to shrinkage after 50 days and the reduced cracking

load are Vsh0= 0.58 MPa and Pcr.sh = 15.9 kN. From equation A.6, D= 4.0 and at a sustained

load of P = 40 kN, the calculated average strain at age 83 days (obtained from Equation A.9)

is 1475 x 10-6 and this agrees reasonably well with the measured member strain at this time of
301

1556 x 10-6.

60
Age (days) = 33 83
Axial Load, P (kN)

50
'Hexp = 147 x 10-6
40

30
Experimental
20
Theoretical
10 -6
Bare bar
Average Strain (x10 )
0
-250 0 250 500 750 1000 1250 1500 1750 2000
Figure A-2: Axial load versus average axial strain forLTN12A.
APPENDIX B TIME-DEPENDENT ANALYSIS OF A
CRACKED OR UNCRACKED BEAM SECTION
An alternative approach for analysing a time-dependent reinforced concrete beam section is

given by Gilbert (1988). The method is also called a relaxation procedure (Bresler and Selna

1964). The basic theory of this method is to assume that the total strain throughout the section

remains constant during any time interval, meanwhile creep and shrinkage strain components

vary. Consequently the instantaneous concrete strains shall be changed, as well as concrete

stresses.

Uncracked beam section

Consider a singly-reinforced concrete beam section in Figure B-1. The beam section has a

cross section of b (width) × D (depth) and is reinforced with Ast bottom tensile bars. The

effective depth of the tensile reinforcement is dst. Assume that the section is imposed with a

relatively small moment M so that it remains uncracked throughout. By using the method of

transformed section, the area of the steel reinforcement is transformed into an equivalent

concrete area, AstEs /Ec = nAst, where n is the modular ratio between steel and concrete, Es /Ec.

Figure B-1: A singly- reinforced beam section under bending.


303

Figure B-2 shows the transformed section and the strain distribution of the uncracked section

without creep and shrinkage taken into account.

Figure B-2: Transformed section and initial strain distribution.

By assuming that the plane section remains plane, at moment M, the instantaneous strain

distribution at any point with a distance y from the top fibre can be given as

Hi H oi  yN i B.1

where oi is the instantaneous strain at the top fibre and i is the instantaneous curvature of the

section. By assuming the linear-elastic stress-strain relationship, the instantaneous stress

distribution can be given as

Vi EcH i E c H oi  yN i B.2

By integrating the first moment of the stress block about the top fibre, the bending moment is

calculated as

Mi ³
 V i ydA  E c H oi B  E c N i I B.2

where B is the first moment of the transformed area about the top fibre, and I is the second
304

moment of area about the top fibre.

Therefore the values of oi and i can be obtained through the following expressions

BM i
H oi B.3
E c AI  B 2

AM i
Ni B.4
E c AI  B 2

where A is the area of the transformed section. The relaxation approach used here is to allow

the stress along the section to change with time due to the change of instantaneous strain.

Thus the equilibrium of the section is not satisfied. To regain the equilibrium along the section,

a change of axial force N and bending moment M must be applied. The time-dependent

variation of top fibre strain and curvature produced by M and N (called restraining forces)

can be obtained from the following equations.

Be M  I e N
 o B.5
E e Ae I e  Be2

Ae M  B e N
 B.6
E e Ae I e  Be2

where A e is the area of the age-adjusted transformed section; B e and I e are the first and

second moments of the area of the age-adjusted transformed section about the top fibre; E e is

the age-adjusted effective modulus (Equation 2.3).

The restraining actions on the section due to the change of creep and shrinkage strain can be

obtained as

N E e >M Ac  oi  Bc i   sh Ac @ B.7
305

M E e >M  Bc  oi  I c i   sh Bc @ B.8

where Ac, Bc and Ic are the properties of the concrete section about the top fibre. The loss of

stress in the concrete at any point with a distance y from the top fibre due to relaxation is

 relax >
E e  M  oi  y i   sh @ B.9

The stress that is needed to restore the equilibrium with the applied actions is:

 restore >
E e 'H o  y'N @ B.10

The actual change of concrete stress with the time interval is the sum of relax and restore

  relax   restore B.11

The change of stress in the steel bar is given by

 s E s 'H o  d'N B.12

Fully-cracked beam section

The analysis of fully-cracked section is also based on the same assumption that cross section

remains plane and the total strain of the cross section remains unchanged with time. Another

assumption is made in this case that the depth of concrete compression zone dn remains

constant with time and is calculated from the short-term analysis. This assumption indicates

that the short-term and time-dependent change of stress or strain can be calculated separately

and added together by the means of superposition. Little error was found under this

assumption.

Under the sustained moment M (greater than cracking moment Mcr), cracking occurs
306

underneath the neutral axis of the section and concrete losses its ability to carry any tensile

stress. Equation B.7 and B.8 can be adopted to calculate the restraining actions to restore

equilibrium of the section when the relaxation method is used. However, the section

properties, such as A, B, and I, (and Ac, Bc, and Ic) in the equations are calculated according to

the concrete compression zone dn rather than the entire depth D. The change of strain

distribution with time can be calculated by equation B.5 and B.6, where A e , B e and I e are the

properties of the fully-cracked age-adjusted transformed section.

Accordingly, the change of concrete stress with any time interval throughout the depth of the

section can be determined by Equations B.9, B.10, and B.11. The change of steel stress can

also be found by using Equation B.12.

WORKING EXAMPLES

Some of the tested beams and slabs presented in Chapter 5 are chosen to be analysed by the

method described above.

BSTS2-16

For the short-term specimens (BSTS2-16), the fully-cracked response needs to be changed

according to the shrinkage prior to loading, which herein are calculated using the analytical

method. The calculated fully-cracked responses were then used to analyse instantaneous

responses of tension stiffening in the experiment.

The relevant properties BSTS2-16 are:

Ec = 25000MPa (day 14); Es = 204000MPa; c = 0.92;  = 0.8; sh = -220 × 10-6; (before

loading) Ast = 401.92mm2; b = 200mm, D = 400mm, dst = 342mm.


307

For a fully-cracked beam section, the depth of the compression concrete zone dn must be

derived before the properties of the cracked section can be determined. In this study, dn is

calculated by considering the equilibrium of the concrete and steel stresses through the

section. For BSTS2-16,

dn 90.0mm

The properties of the fully-cracked transformed section are calculated subsequently as

A 21217mm 2 ; B 1909882mm 3 ; I 4.25 u 10 8 mm 4

and the properties of the uncracked part of the cracked concrete cross section (compressive

zone) with respect to the top fibre are

Ac 18003mm 2 ; Bc 810230mm 3 ; I c 4.86 u 10 7 mm 4

The age-adjusted effective modulus during the time interval of drying is (Equation 2.23)

25000
Ee 14401MPa
1  0.8 u 0.92

In terms of the age-adjusted transformed section, the properties are calculated as:

Ae 23183mm 2 ; B e 2719228mm 3 ; I e 7.01 u 10 8 mm 4 and

Equation B.7 and B.8 give the restraining actions on the cracked cross-section due to creep

and shrinkage when the applied moment is zero.

N >
14401u 0.92 18003 u 0  810230 u 0  220 u 10 6 u 18003 @ 57.0kN

M >
14401u 0.92 810230 u 0  4.86 u 10 7 u 0  220 u 10 6 u 810230 @ 2.57 kN .m

Therefore the variation of time-dependent strain at the top fibre and curvature are obtained
308

using equation B.5 and B.6.

E e §¨ A e I e  B e ·¸ 1.28 u 1017 N.mm


2 4
© ¹

2719228 u (2.57 u 10 6 )  7.01u 10 8 u 57.0 u 10 3


 o 259 u 10 6 ;
1.28 u 10 17

23183 u  2.57 u 10 6  2719228 u 57.0 u 10 3


 7.48 u 10 7
1.28 u 10 17

Integrating the increment of curvature  across the entire span (2.4m) of the beam gives the

change of mid-span deflection during the drying period of the beam, assuming all the sections

are fully-cracked.

G sh 0.cr 0.5mm (highlighted in Figure 5-20)

After being dried for a period of time, BSTS2-16 was subjected to monotonic applied moment.

At a specific applied moment M (M > Mcr.sh), the fully-cracked response needs to be shifted

with an amount of sh0.cr (Figure 3-5), which indicates that the fully-cracked deflection is the

sum of sh0.cr and the one calculated without considering initial drying shrinkage (short-term

fully-cracked deflection). Thereafter tension stiffening deflection for the test beam can be

calculated by subtracting the measured mid-span deflection from the calculated fully-cracked

deflection, shown as line B’E’ in Figure 3-5.

SLTN4-12A

This analytical method can also be adopted to derive the long-term fully-cracked responses of

a flexural member. The slab SLTN4-12A tested by the author and presented in Chapter 5 is

chosen as a working example in this case. The basic properties for specimen SLTN4-12A are

Ec = 25000MPa (Day 15); Es = 200000MPa; c = 1.47;  = 0.8; sh = -400 × 10-6 (70 days
309

after first loading), Ast = 452mm2; b = 800mm, D = 140mm, dst = 114mm, and Mi = 17.9kN.m

(in the constant moment region).

Under instantaneous moment, the depth of compressive zone for the cracked cross-section of

SLTN4-12A is calculated as

dn 28mm

The properties of the fully-cracked transformed section about the top fibre are calculated as

A 25935mm 2 ; B 723565mm 3 ; I 5.28 u 10 7 mm 4

The properties of the concrete compressive zone about the top fibre are:

Ac 22319mm 2 ; Bc 311341mm 3 ; I c 5.79 u 10 6 mm 4

At first loading, the instantaneous top fibre strain and curvature in the constant moment

region can be obtained using Equations B.3 and B.4.

723565 u 17.9 u 10 6
H oi 612 u 10 6
25000 25935 u 5.28 u 10 7  723565 2

25935 u 17.9 u 10 6
Ni 2.2 u 10 5
25000 25935 u 5.28 u 10 7  723565 2

The age-adjusted effective modulus during the sustained loading period is

25000
Ee 11489MPa
1  0.8 u 1.47

Therefore by using equation B.7 and B.8 and assuming applied moment remains constant

with time, the restraining actions in the constant moment region are calculated as
310

N >
11489 u 1.47 22319 u 612 u 10 6  311341 u 2.2 u 10 5  400 u 10 6 u 22319 @ 218kN

M >
11489 u 1.47  311341 u 612 u 10 6  5.79 u 10 6 u 2.2 u 10 5  400 u 10 6 u 311341 @ 2.5kN .m

In terms of the age-adjusted transformed section, the properties are calculated as:

Ae 30188mm 2 ; B e 1208341mm 3 ; I e 10.8 u 10 7 mm 4

The time-dependent increment of top fibre strain and curvature can be calculated from

equation B.5 and B.6 (notice that in the calculation below, the quantity
2
E e Ae I e  B e 2.07 u 1016 N.mm4.)

1208341 u (2.5 u 10 6 )  10.8 u 10 7 u 218 u 10 3


 o 99.2 u 10 5 ;
2.07 u 1016

30188 u  2.5 u 10 6  1208341 u 218 u 10 3


 9.07 u 10 6
2.07 u 1016

which indicates that during the sustained loading period, shrinkage warping induced time-

dependent curvature of the fully-cracked section along the specimen. Since the values of 0

and  are dependent on the magnitude of applied moment Mi, as the restraining forces are

related to the instantaneous curvature and strain, the time-dependent curvature along the

length of the specimen varies. Therefore, in this study, a number of sections at regular

intervals along the specimens were chosen, and time-dependent curvatures of these sections

are then used to calculate the fully-cracked deflections at mid-span. As for SLTN4-12A, after

70 days under sustained load can be calculated by integrating the curvature diagram and is

given as

G fc 21.7mm (Table 5-13)

Then the tension stiffening deflection after 70 days can be calculated by subtracting the
311

measured deflection at mid-span from the calculated fully-cracked deflection. Similarly, given

the concrete properties at different loading stages (such as creep and shrinkage strains), the

time-dependent fully-cracked deflections of the specimens can also be determined at these

stages. Then the long-term tension stiffening deflections at different loading stages can be

calculated accordingly.
312

REFERENCES

CEB-FIP (2004). "Eurocode 2: Design of concrete structures Part 1-1: General rules and rules

for buildings. BS EN 1992-1-1: 2004." European Committee for Standardisation.

Standards Australia (AS), (2001). "Australian Standards for concrete structures" AS 3600-

2009, Sydney, Australia.

CEB-FIP, (1993). CEB-FIP Model code (1990 design code), Comitte Euro-International du

Beton, Thomas Telford, London.

Abrishami, H. H., and Mitchell, D. (1996). "Influence of splitting cracks on tension

stiffening." ACI Structural Journal, 93(6), 703-710.

Al-Fayadh, S. (1997). "Cracking behaviour of reinforced concrete tensile members." Division

of Concrete Structure, Chalmers University of Technology, Goteborg, Sweden.

Attard, M. M., Minh, N. G., and Foster, S. J. (1996). "Finite element analysis of out-of-plane

buckling of reinforced concrete walls." Computers and Structures, 61(6), 1037-1042.

Balazs, G. L. (1993). "Cracking analysis based on slip and bond stresses." ACI Materials

Journal (American Concrete Institute), 90(4), 340-348.

Barros, M. H. F. M., Martins, R. A. F., and Ferreira, C. C. (2001). "Tension stiffening model

with increasing damage for reinforced concrete." Engineering Computations (Swansea,


313

Wales), 18(5-6), 759-785.

Bazant, Z., and Prasannan, S. (1989b). "Solidification theory for concrete creep II.

Verification and application." Journal of Engineering Mechanics, 115(8), 1704-1725.

Bazant, Z. P. (1972). "Prediction of concrete creep effects using Age-Adjusted Effective

Modulus Method." ACI Journal (American Concrete Institute), 69, 212-217.

Bazant, Z. P. (1977). "Viscoelasticity of porous solidifying material -- concrete", Journal of

Engineering Mechanics, ASCE, 103(EM6), 1049-1067.

Bazant, Z. P. (1982). "Mathematical models for creep and shrinkage of concrete." Symposium

on Fundamental Research on Creep and Shrinkage of Concrete, Swiss Federal Institue of

Technology, Lausane, 163-256.

Bazant, Z. P. (1988a). "Material models for structural creep analysis." Mathematical

Modelling of Creep and Shrinkage of Concrete, Z. P. Bazant, ed., John Wiley & Sons, Inc.,

New York, 99-215.

Bazant, Z. P. (1988b). Mathematical Modeling of Creep and Shrinkage of Concrete, John

Wiley and Sons.

Bazant, Z. P., Abrams, M. S., Chern, J. C., and Gillen, M. P. (1982). "Normal and refractory

concretes for LMFBR applications, Vol.1: Review of literature on high-temperature behaviour

of Portland Cement and refractory concretes." Report NP-2437, Electric Power Research

Institute, Palo Alto, Calif.

Bazant, Z. P., Asghari, A., and Schmidt, J. (1976). "Experimental study of creep hardened

cement past at variable water content." Materials and Structures (RILEM, Paris), 9, 279-290.
314

Bazant, Z. P., and Baweja, S. (1995a). "Justification and refinements of Model B3 for

concrete creep and shrinkage. 1. Statistics and sensitivity." Materials and

Structures/Materiaux et Constructions, 28(181), 415-430.

Bažant, Z. P., and Chern, J. C. (1985). "Concrete creep at variable humidity: constitutive law

and mechanism." Materials and Structures (RLEM, Paris), 18, 1-20.

Bažant, Z. P., Hauggaard, A. B., Baweja, S., and Ulm, F.-J. (1997). "Microprestress-

solidification theory for concrete creep: I: aging and drying effects." Journal of Engineering

Mechanics, 123(11), 1188-1194.

Bazant, Z. P., and Najjar, L. J. (1973). "Comparison of approximate linear methods for

concrete creep." 99(ST9), 1851-1874.

Bazant, Z. P., and Oh, B. H. (1983). "Crack band theory for fracture of concrete." Materiaux

et Constructions, Materials and Structures, 16(93), 155-177.

Bazant, Z. P., and Prasannan, S. (1989a). "Solidification theory for concrete creep I.

Formulation." Journal of Engineering Mechanics, 115(8), 1691-1703.

Bažant, Z. P., and Wittmann, F. H. (1983). "Creep and Shrinkage in Concrete Structure." John

Wlley and Sons Ltd.

Bažant, Z. P., and Wu, S. T. (1974). "Creep and shrinkage law for concrete at variable

humidity." J. Engrg. Mech. Div., ASCE, 100(6), 1183-1209.

Bazant, Z. P., and Xi, Y. (1995). "Continuous retardation spectrum for solidification theory of

concrete creep." Journal of Engineering Mechanics, 121(2), 281-288.


315

Bažant, Z. P., and Xi, Y. (1994). "Drying creep of concrete: Constitutive model and new

experiments separating its mechanisms." Materials and Structures/Materiaux et Constructions,

27(165), 3-14.

Beeby, A. W., and Scott, R. H. (2002). "Tension stiffening of concrete (Behaviour of tension

zones in reinforced concrete including time dependent effects)." Concrete Society.

Belarbi, A., and Hsu, T. T. C. (1991). "Constitutive laws of reinforced concretein biaxial

tension compression." Report NO.:UHCEE 91-2, Department of Civil Engineering,

University of Houston, Houston. .

Bischoff, P. H. (2001). "Effects of shrinkage on tension stiffening and cracking in reinforced

concrete." Canadian Journal of Civil Engineering, 28(3), 363-374.

Bischoff, P. H. (2005). "Reevaluation of deflection prediction for concrete beams reinforced

with steel and fiber reinforced polymer bars." Journal of Structural Engineering, 131(5), 752-

762.

Bischoff, P. H., and Paixao, R. (2004). "Tension stiffening and cracking of concrete reinforced

with glass fiber reinforced polymer (GFRP) bars." Canadian Journal of Civil Engineering,

31(4), 579-588.

Bolander, J., Jr., Satake, M., and Hikosaka, H. (1992). "Bond degradation near developing

cracks in reinforced concrete structures." 52.

Branson, D. E. (1968). "Design procedures for computing deflections." Journal of the

American Concrete Institute, 65(9), 730-742.

Bresler, B., and Selna, L. (1964). "Analysis of time-dependent behaviour of reinforced


316

concrete structures." Symposium on creep of concrete, ACI Special Publication SP-9, No.5,

115-128.

Brooks, J. J., and Neville, A. M. (1977). "A comparison of creep, elasticity and strength of

concrete in tension and compression." Magazine of Concrete Research, 29(100), 131-141.

Castel, A., Vidal, T., and Francois, R. (2006) "Effective tension active cross-section of

reinforced concrete beams after cracking." Materials and Structures, 39, 115-126.

CEB-FIP. (1997). "CEB-FIP model code (MC-97)." Comité Euro-International du Béton

(CEB), ed., Thomas Telford, London.

Chong, K. T., Foster, S. J., and Gilbert, R. I. (2008). "Time-dependent modelling of RC

structures using the cracked membrane model and solidification theory." Computers and

Structures, 86(11-12), 1305-1317.

Clark, L. A., and Cranstown, W. B. (1979). The influence of bar spacing on tension stiffening

in reinforced concrete slabs, Advances in concrete slab technology, Pergammon Press,

London.

Clark, L. A., and Speirs, D. M. (1978). Tension stiffening in reinforced concrete beams and

slabs under short-term load, Cement and Concrete Association.

Collins, M. P., and Porasz, A. (1989). " Shear strength for high strength concrete." Bull. NO.

193, Design Aspects of High Strength Concrete, Comite Euro-International du Beton: 75-83.

Committee, A. C. (2005). "Building code requirements for structural concrete " ACI318-05,

ACI Committee 318, Detroit.


317

Cox, J. V. (1994). "Development of a plasticity bond model for reinforced concrete - Theory

and validation for monotonic applications." TR-2036-SHR, Naval Facilities Engineering

Service Center, Port Hueneme, CA.

Cox, J. V., and Herrmann, L. R. (1998). "Development of a plasticity bond model for steel

reinforcement." Mechanics of Cohesive-Frictional Materials, 3(2), 155-180.

Darwin, D., and Graham, E. K. (1993), "Effect of Deformation Height and Spacing on Bond

Strength of Reinforcing Bars." ACI Structural Journal, 90(6), 646-657.

Darwin, D., and Pecknold, D. A. (1977). "Nonlinear biaxial stress-strain law for concrete."

American Society of Civil Engineers, Journal of the Engineering Mechanics Division, 103(2),

229-241.

Den Uijl, J. A., and Bigaj, A. J. (1996). "A bond model for ribbed bars based on concrete

confinement." Heron, 41(3), 201-226.

Dischinger, F. (1937). "Untersuchungen ueber die Knicksicherheit, die elastische Verformung

und das Kriechen des Betons bei Bogenbruecken." Bauingenieur, 18(39/40), 595-621.

Domone, P. L. (1974). "Uniaxial tensile creep and failure of concrete." Magazine of Concrete

Research, 26(88), 144-152.

Ebead, U. A., and Marzouk, H. (2005). "Tension-stiffening model for FRP-strengthened RC

concrete two-way slabs." Materials and Structures, 38, 193-200.

Edwards, A. D., and Yannopoulos, P. J. (1979). "Local bond-stress to slip relationships for hot

rolled deformed bars and mild steel plain bars." Journal of the American Concrete Institute,

76(3), 405-420.
318

Eligehausen, R., Popov, E. P., and Berbero, V. V. (1983). "Local bond stress-slip relationships

of deformed bars under generalized excitations: experimental results and analytical model."

Earthquake Engineering Research Center, College of Engineering, University of California,

Berkeley, California.

European Committee for Standardization (CEN). (1992). "Eurocode 2: Design of concrete

structures Part 1-1: General rules for buildings." DD ENV 1992-1-1m European Prestandard,

Brussels, Belgium.

federation internationale du beton (fib). (2000). "Bond of reinforcement in concrete.”

International Federation for Structure Concrete (fib), Lausanne, Switzerland.

Fields, K., and Bischoff, P. H. (2004). "Tension stiffening and cracking of high-strength

reinforced concrete tension members." ACI Structural Journal, 101(4), 447-456.

Floegl, H., and Mang, H. A. (1982). "Tension stiffening concept based on bond slip."

American Society of Civil Engineers, Journal of the Structural Division, 108(ST12), 2681-

2701.

Foster, S. J., and Marti, P. (2003). "Cracked membrane model: Finite element

implementation." Journal of Structural Engineering, 129(9), 1155-1163.

Franke, L. (1976). "Einfluss der Belastung-Sdauer auf das Verbundverhalten von Stahl in

Betaon." Deutscher Ausschuss fur Stahlbeton, Heft 268.

Gambarova, P. G., Paolo Rosati, G., and Zasso, B. (1989a). "Steel-to-concrete bond after

concrete splitting. Test results." Materials and Structures, 22(127), 35-47.

Gamble, B. R., and Parrott, L. J. (1978). "Creep of concrete in compression during drying and
319

wetting." Magazine of Concrete Research, 30(104), 129-138.

Gamble, W. L. (1982). "Creep of concrete in variable environments." American Society of

Civil Engineers, Journal of the Structural Division, 108(ST10), 2211-2222.

Gao, D., Benmokrane, B., and Masmoudi, R. (1998). "A calculating method of flexural

properties of FRP-reinforced concrete beam: Part 1: crack width and deflection.Tech. Rep."

Dept. of Civil Engineering, Univ. of Sherbrooke,, Quebec, Canada.

Ghali, A. (1993). "Deflection of reinforced concrete members: a critical review." ACI

Structural Journal, 90(4), 364-373.

Ghali, A., Hall, T., and Bobey, W. (2001). "Minimum thickness of concrete members

reinforced with fibre reinforced polymer bars." Can. J. Civ. Eng., 28(4), 583-592.

Gilbert, R. I. (1983). "Deflection calculations for reinforced concrete beams." Institution of

Engineers, Australia, Civil Engineering Transactions, 25(2), 128-134.

Gilbert, R. I. (1988). Time effects in concrete structures, Elsevier Science Publisher.

Gilbert, R. I. (1999). Deflection calculation for reinforced concrete structures - Why we

sometimes get it wrong, ACI Structural Journal, 96(6), 1027-1032.

Gilbert, R. I. (2006). "Design for flexural crack control - recent amendments to the Australian

Standard AS3600." 1st International Structural Specialty Conference, CSCE, Calgary Alberta,

Candada.

Gilbert, R. I. (2007). "Tension stiffening in lightly reinforced concrete slabs." Journal of

Structural Engineering, 133(6), 899-903.


320

Gilbert, R. I., and Warner, R. F. (1978). "Tension stiffening in reinforced concrete slabs."

Journal of the Structural Division, ASCE, , 104(12), 1885-1900.

Gilbert, R. I., and Wu, H. Q. (2009). "Time-dependent stiffness of cracked reinforced concrete

elements under sustained actions." Australian Journal of Structural Engineering, 9(2), 151-

158.

Giuriani, E., Plizzari, G., and Schumm, C. (1991). "Role of stirrups and residual tensile

strength of cracked concrete on bond." Journal of Structural Engineering, 117(1), 1-18.

Goodman, R. E., Taylor, R. L., and Brekke, T. L. (1968). "Model for mechanics of jointed

rock." American Society of Civil Engineers Proceedings, Journal of the Soil Mechanics and

Foundations Division, 94(SM3), 637-659.

Goto, Y. (1971). "Cracks Formed in Concrete Around Deformed Tension Bars." ACI Journal

(American Concrete Institute), 68(4).

Gribniak V., (2009). "Shrinkage influence on tension stiffening of concrete structures."

Doctoral Dissertation, Cilnius Gediminas Technical University.

Guptam, A. K., and Maestrini, S. R. (1990). "Tension-stiffness model for reinforced concrete

bars." Journal of the Structural Division, ASCE, 116(3), 769-791.

Hamad, B. S. (1995). "Bond Strength Improvement of Reinforcing Bars with Specially

Designed Rib Geometries." ACI Structural Journal, 92(1), 3-13.

Hillerborg, A. (1983), Analysis of one single crack. Fracture Mechanics of Concrete. Ed. F. H.

Wittmann. Elsevier, 223-249.


321

Illston, J. M. (1965). "The creep of concrete under uniaxial tension." Magazine of Concrete

Research, 17(51), 77-84.

Illston, J. M., and Stevens, R. F. (1972). "Long-term cracking in reinforced concrete beams."

Proceedings of the Institution of Civil Engineers (London). Part 1 - Design & Construction,

53(Part 2), 445-459.

Kaklauskas G, Gribniak V, Bacinskas D and Vainiunas P (2009). Shrinkage influence on


tension stiffening in concrete members. Engineering Structures 31(6): 1305-1312.

Kankam, C. K. (1997). "Relationship of bond stress, steel stress, and slip in reinforced

concrete." Journal of Structural Engineering, 123(1), 79-85.

Kaufmann, W. (1998). "Strength and deformations of structural concrete subjected to in-

plane shear and normal forces," Swiss Federal Institute of Technology, Zurich.

Kaufmann, W. and Marti, P., "Structural Concrete: Cracked Membrane Model," ASCE

Journal of Structural Engineering, 124(12), 1467-1475.

Kishek, M. A. (1983). "Tension stiffening and crack widths in reinforced concrete beam and

slab elements."

Koch, R., and Balazs, G. L. (1993). "Slip increase under cyclic and long term loads." Otto-

Graf-Journal, 4, 160-191.

Kupfer, H., Hilsdorf, H. K., and Rusch, H. (1969). "Behaviour of concrete under biaxial

stresses." ACI Journal (American Concrete Institute), 66(8), 656-66.

Lin, C.-S., and Scordelis, A. C. (1975). "Nonlinear analysis of RC shells of general form."
322

101(3), 523-538.

Lorman, W. R. (1940). "The theory of concrete creep." Preceedings of ASTM, 40, 1082-1102.

Lowes, L. N. (1999). "Finite element modeling of reinforced concrete beam-column bridge

connections." Thesis (PhD), University of California, Berkeley.

Lowes, L. N., Moehle, J. P., and Govindjee, S. (2004). "Concrete-steel bond model for use in

finite element modeling of reinforced concrete structures." ACI Structural Journal, 101(4),

501-511.

Lutz, L. A., and Gergely, P. (1967). "Mechanics of bond and slip of deformed bars in

concrete." American Concrete Institute -- Journal, 64(11), 711-721.

Lutz, L. A., Gergely, P., and Winter, G. (1966). "Mechanics of bond and slip of deformed

reinforcing bars in concrete." Cornell University, Ithaca, NY, United States.

Maekawa, K., Pimanmas, A., and Okamura, H. (2003). Nonlinear Mechanics of Reinforced

Concrete, New York, Spon Press.

Malvar, L. J. (1992). "Bond of reinforcement under controlled confinement." ACI Material

Journal (American Concrete Institute), 89(6), 593-601.

Marechal, J. C. (1972). "Creep of concrete as a function of temperature." In Concrete for

Nuclear Reactors, ACI SP-34(1), 547-564.

Marti, P., Alvarez, M., Kaufmann, W., and and Sigrist, V. (1998). "Tension chord model for

structural concrete." Structural Engineering International, 4(98), 287-298.

McHenry, D. (1943). "New aspect of creep in concrete and its application to design."
323

American Society for Testing Materials -- Proceedings, 43, 1069-1084.

McMillan, F. R. (1916). "Method of designing reinforced concrete slabs, discussion of A. C,

Janni's paper." Trans. ASCE, 80, 1738.

Miyakawa, T., and Kawakami, T. (1987). "Nonlinear behaviour of. cracked reinforced

concrete plate element under uniaxial compression." JSCE Proceedings, 378, 249-258.

Navaratnarajah, V., and Speare, P. R. S. (1987). "Theory of transfer bond resistance of

deformed reinforcing bars in concrete under lateral pressure." Magazine of Concrete Research,

39(140), 161-168.

Nayal, R. and Rasheed, H. A. (2006). "Tension stiffening model for concrete beams reinforced

with steel and FRP bars." ASCE Journal of Materials in Civil Engineering, 18(6), 831-841.

Nejadi, S., and Gilbert, I. (2004a). "Shrinkage cracking and crack control in restrained

reinforced concrete members." ACI Structural Journal, 101(6), 840-845.

Nejadi, S., and Gilbert, I. (2004b). "An experimental study of flexural cracking in reinforced

concrete members under sustained loads." UNICIV Report. University of New South Wales.

No. R-435.

Neville, A. M., Dilger, W. H., and Brooks, J. J. (1983). Creep of plain and structural concrete,

Construction Press (Longman Group Ltd).

Ouyang, C., Wollrab, E., Kulkarni, S. M., and Shah, S. P. (1997). "Prediction of cracking

response of reinforced concrete tensile members." Journal of Structural Engineering, 123(1),

70-78.
324

Petersson, P. E. (1981). "Crack Growth and Development of Fracture Zone in Plain Concrete

and Similar Material. Report No.:TVBM-1006." Lund Institute of Technology,Lund,Sweden.

Pickett, G. (1942). "Effect of change in moisture-content on creep of concrete under sustained

load." American Concrete Institute -- Journal, 13(4), 333-355.

Prakhya, G. K. V., and Morley, C. T. (1990). "Tension-stiffening and moment-curvature

relations of reinforced concrete elements." ACI Structural Journal (American Concrete

Institute), 87(5), 597-605.

Ragueneau, F., Dominguez, N., and Ibrahimbegovic, A. (2006). "Thermodynamic-based

interface model for cohesive brittle materials: Application to bond slip in RC structures."

Computer Methods in Applied Mechanics and Engineering, 195(52), 7249-7263.

Robins, P. J., and Standish, I. G. (1984). "Influence of lateral pressure upon anchorage bond."

Magazine of Concrete Research, 36(129), 195-202.

Roelfstra, P. E., Wittmann, F. H. (1986), "Numerical method to link strain softening with

failure of concrete". Fracture Toughness and Fracture Energy of Concrete. Ed. F. H.

Wittmann, Elsevier, 163-175.

Rostasy, F. S., and Kepp, B. "Time-dependence of bond." Paisley, Scotl, 183-192.

Salem, H., Hauke, B., Maekawa, K. (1999). "Fracture of concrete cover. Its effect on tension

stiffening and modeling." Proceedings of JSCE, 613, 295-307.

Shima, H., Chou, L.-L., and Okamura, H. (1987). "Micro and macro models for bond in

reinforced concrete." Journal of the Faculty of Engineering, University of Tokyo, Series B,

39(2), 133-194.
325

Soh, C. K., Liu, Y., Dong, Y. X., and Lu, X. Z. (2003). "Damage model based reinforced-

concrete element." Journal of Materials in Civil Engineering, 15(4), 371-380.

Somayaji, S., and Shah, S. P. (1981). "Bond stress versus slip relationship and cracking

response of tension members." Journal of the American Concrete Institute, 78(3), 217-225.

Sooriyaarachchi, H., Pilakoutas, K., Byars, E. (2005). "Tension stiffening behaviour of GFRP-

reinforced concrete" ACI Special Publications, NUMB 230, 2, 975-990.

Stevens, R. F. (1970). "The time-dependent deflections of reinforced concrete beams under

sustained loading and the development of cracking."

Svensvik, B. (1981). "Zum Verformungsverhalten gerissener Stahlbetonbalken unter

Einschlub der Mitwirkung des Betons auf Zug in Abhangigkeit von Last und Zeit." Diss. TU

Braunschweig.

Tepfers, R. (1979). "Cracking of concrete cover along anchored deformed reinforcing bars."

Magazine of Concrete Research, 31(106), 3-12.

Thorenfeldt, E., Tomasewicz, A., and al, e. (1987). "Mechanical properties of high strength

concrete and application in design." International Symposium on Utilization of High Strength

Concrete, Stavanger, Norway.

Toutanji, H. A., and Saafi, M. (2000). "Flexural behaviour of concrete beams reinforced with

glass fiber-reinforced polymer (GFRP) bars." ACI Structural Journal, 97(5), 712-719.

Trost, H. (1967). "Auswirkungen des Superpositionsprinzips auf Kriech- und Relaxations-

probleme bei Beton Und Spannbeton", Beton- und Stahlbetonbau, 62(10), 230-238.
326

Untrauer, R. E., and Henry, R. L. (1965). "Influence of normal pressure on bond strength."

American Concrete Institute -- Journal, 62(5), 577-586.

Vecchio, F. J., and Collins, M. P. (1982). "The response of reinforced concrete to in-plane

shear and normal stresses." Department of Civil Engineering, University of Toronto,Toronto.

Vecchio, F. J., and Collins, M. P. (1986). "Modified compression-field theory for reinforced

concrete elements subjected to shear." Journal of the American Concrete Institute, 83(2), 219-

231.

Volterra, V. (1913). "Lecons sur les Fonctions de Ligne." Gauthier-Villars, Paris.

Whitney, C. S. (1932). "Plain and reinforced concrete arches." American Concrete Institute --

Journal, 3(7), 479-519.

Wittmann, F. H., Roelfstra, P. E. (1980). "Total deformation of loaded drying concrete."

Cement and Concrete Research, 10(5), 601-610.

Wu, H. Q., and Gilbert, R. I. (2009a) "Modeling short-term tension stiffening in reinforced

concrete prisms using a continuum-based finite element model." Engineering structures,

31(10), 2380-2391.

Wu, H. Q., and Gilbert, R. I. (2009b) "Effect of shrinkage on time-dependent deflection of

reinforced concrete slabs." Proceedings of the 24th Biennial Conference of the Concrete

Institute of Australia, Sydney, Paper No. 6b-1.

You might also like