Anda di halaman 1dari 369

Numerical Modelling of Time-dependent Cracking and Deformation of Reinforced Concrete Structures

Kak Tien Chong

A thesis submitted as a partial fulfilment of the requirements for the degree of Doctor of Philosophy

December 2004

UNSW
THE UNIVERSITY OF NEW SOUTH WALES SYDNEY AUSTRALIA School of Civil & Environmental Engineering

CERTIFICATE OF ORIGINALITY
I hereby declare that this submission is my own work and to the best of my knowledge 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 educational institution, except where due acknowledgement 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 projects design and conception or in style, presentation and linguistic expression is acknowledged.

Kak Tien Chong

To those who have the thirst for knowledge

ABSTRACT
For a structure to remain serviceable, crack widths must be small enough to be acceptable from an aesthetic point of view, small enough to avoid waterproofing problems and small enough to prevent the ingress of water that may lead to corrosion of the reinforcement. Crack control is therefore an important aspect of the design of reinforced concrete structures at the serviceability limit state. Despite its importance, code methods for crack control have been developed, in the main, from laboratory observations of the instantaneous behaviour of reinforced concrete members under load and fail to account adequately for the time-dependent development of cracking. In this study numerical models have been developed to investigate timedependent cracking of reinforced concrete structures. Two approaches were adopted to simulate cracking in reinforced concrete members. The first approach is the distributed cracking approach. In this approach, steel reinforcement is smeared through the concrete elements and bond-slip between steel and concrete is accounted for indirectly by including the tension stiffening effect. The second approach is the localized cracking approach, in which concrete fracture models are used in conjunction with bond-slip interface elements to model stress transfer between concrete and steel. Creep of concrete has been incorporated into the models by adopting the principal of superposition and the time-dependent development of shrinkage strain of concrete is modelled using an approximating function. Both creep and shrinkage were treated as inelastic pre-strains and applied to the discretized structure as equivalent nodal forces. Apart from material non-linearity, non-linearity arising from large deformation was also accounted for using the updated Lagrangian formulation. The numerical models were used to simulate a series of laboratory tests for verification purposes. The models were assessed critically by comparing the numerical results with the test data and the numerical results are shown to have good correlations with the test results. In addition, a comparison was undertaken among the numerical models and the pros and cons of each model were evaluated.

iv

A series of controlled parametric numerical experiments was devised and carried out using one of the numerical models. Various parameters were identified and investigated in the parametric study. The effects of the parameters were thoroughly examined and the interactions between the parameters were discussed in detail.

ACKNOWLEDGEMENTS
The work presented in this thesis was undertaken in the School of Civil and Environmental Engineering at the University of New South Wales. I wish to express my sincere gratitude to Professor R. Ian Gilbert for giving me the opportunity to participate in this research project. His patient supervision, suggestions, critical comments and continuous support throughout the course of this study are very much appreciated. I would also like to thank my co-supervisor, Associate Professor Stephen J. Foster, with whom I had many inspiring and fruitful discussions on numerical methods and his patient guidance is one of the most important factors promoting the completion of this study. This research was funded by Australian Research Council (ARC) Discovery Grant No. DP0210039 and an Australian Government International Postgraduate Research Scholarship (IPRS). The ARC and Scholarship supports are gratefully acknowledged. I would like to express my deepest gratitude to my family for their love, support and encouragement while I was thousand of miles away from home. Finally, I wish to thank my beloved girl friend, Peggy, who has walked me through all the good times and bad times throughout these years, without whose love and support the completion of this thesis would not have been possible.

vi

CONTENT
ABSTRACT ACKNOWLEDGEMENTS NOTATION CHAPTER 1 INTRODUCTION 1.1 Background and Significance 1.2 Objective and Scope 1.3 Outline of Thesis CHAPTER 2 LITERATURE REVIEW 2.1 Introduction 2.2 Instantaneous Behaviour of Concrete 2.2.1 Uniaxial Compression 2.2.2 Uniaxial Tension 2.2.3 Biaxial Loading and Failure Criteria 2.3 Time-dependent Behaviour of Concrete Creep 2.3.1.1 Factors affecting Creep 2.3.1.2 Creep Recovery 2.3.1.3 Principle of Superposition 2.3.2 Shrinkage 2.3.2.1 Chemical Shrinkage 2.3.2.2 Drying Shrinkage 2.3.2.3 Effects of Shrinkage 2.3.3 Interaction of Fracture and Creep 2.3.3.1 Influence of Loading Rate on Peak Load 2.3.3.2 Load Relaxation at Fracture Zone 2.3.3.3 Creep Rupture 2.3.3.4 Time-dependent Fracture Models 2.4 Behaviour of Reinforcement 2.5 Bond between Reinforcement and Concrete 2.5.1 Local Bond Stress-slip Relationship 2.5.2 Influence of Bond on Cracking 2.5.3 Tension Stiffening 2.6 Non-linear Modelling of Concrete Structures
vii

iv vi xii

1 3 5

7 7 7 9 11 13 15 16 17 18 22 23 24 26 26 27 28 28 29 31 32 33 36 37 38

2.3.1

2.6.1 Discrete Crack Approach 2.6.2 Smeared Crack Approach 2.6.2.1 Fixed Crack Model 2.6.2.2 Rotating Crack Model 2.6.2.3 Multiple Fixed Crack Model 2.6.3 Constitutive Models for Concrete 2.6.3.1 Elasticity-based Models 2.6.3.2 Plasticity-based Models 2.6.3.3 Continuous Damage Models 2.6.3.4 Microplane Models 2.6.4 Fracture Models for Concrete 2.6.4.1 Fracture Mechanics 2.6.4.2 Fictitious Crack Model 2.6.4.3 Crack Band Model 2.6.5 Regularization of Spurious Strain Localization 2.6.5.1 Non-local Models 2.6.5.2 Gradient Models 2.6.5.3 Crack Band Formulation as Partial Regularization 2.6.5.4 Regularization by Inclusion of Material Viscosity 2.6.6 Modelling of Steel Reinforcement 2.6.7 Modelling of Steel-Concrete Bond 2.6.7.1 Tension Stiffening 2.6.7.2 Discrete Bond Modelling 2.6.8 Computational Creep Modelling CHAPTER 3 FINITE ELEMENT MODELS FOR REINFORCED CONCRETE 3.1 Introduction 3.2 Continuum Modelling 3.3 Distributed Cracking Approach 3.3.1 Tension Chord Model 3.3.2 Cracked Membrane Model 3.4 Localized Cracking Approach 3.4.1 Crack Band Model 3.4.2 Non-local Smeared Crack Model 3.4.2.1 Issue Related to Non-local Continuum with Local Strain 3.4.2.2 Proposed Non-local Smeared Cracking Formulation 3.5 Orthotropic Membrane Formulation 3.6 Material Constitutive Models 3.6.1 Instantaneous Behaviour of Concrete

39 41 42 43 44 45 46 49 53 55 58 58 59 60 61 62 65 66 67 67 69 69 70 72

76 77 77 79 81 83 84 86 87 89 91 95 95

viii

3.6.1.1 Stress-strain Relationships for Concrete 3.6.1.2 Biaxial Compression State of Stress 3.6.1.3 Tension-Compression State of Stress 3.6.1.4 Biaxial Tension State of Stress 3.6.2 Time-dependent Behaviour of Concrete 3.6.3 Shrinkage 3.6.4 Creep 3.6.5 Solidification Theory for Concrete Creep 3.6.5.1 Rate-type Constitutive Model 3.6.5.2 Finite Element Implementation of Creep 3.6.6 Time-dependent Crack Width 3.6.6.1 Cracked Membrane Model 3.6.6.2 Crack Band Model 3.6.6.3 Non-local Model 3.6.7 Stress-strain Relationship for Reinforcing Steel 3.6.8 Local Bond-slip Model for Bond Interface Element 3.6.9 Concrete Tension Stiffening 3.7 Non-linear Finite Element Implementation 3.7.1 Spatial Discretization 3.7.2 Time Discretization 3.7.3 Principal of Virtual Work 3.7.4 Incremental Iterative Solution Procedures 3.7.5 Geometric Non-linearity 3.7.6 Convergence Criteria 3.7.7 Computational Solution Algorithm 3.8 Finite Element Formulations 3.8.1 Four-node Isoparametric Quadrilateral Element 3.8.2 Two-node Truss Element 3.8.3 Four-node Isoparametric Bond Interface Element CHAPTER 4 EVALUATION OF THE FINITE ELEMENT MODELS 4.1 Introduction 4.2 Mesh Sensitivity of the Localized Cracking Models 4.3 Creep of Plain Concrete under Variable Stress 4.4 Long-term Flexural Cracking Tests 4.4.1 Introduction 4.4.2 Analysis of Long-term Flexural Cracking Tests and Material Properties

96 98 99 100 101 102 103 103 106 108 111 111 111 112 112 113 116 118 119 120 121 123 128 129 130 133 133 136 138

141 141 147 151 151 154

ix

Analysis of Long-term Flexural Cracking Tests using the Distributed Cracking Model Cracked Membrane Model 4.4.3.1 Four-point Bending Beam Tests under Sustained Load 4.4.3.2 Uniformly Loaded One-way Slabs under Sustained Load 4.4.3.3 Discussion 4.4.4 Analysis of Long-term Flexural Cracking Tests using the Localized Cracking Model Crack Band Model 4.4.4.1 Four-point Bending Beam Tests under Sustained Load 4.4.4.2 Uniformly Loaded One-way Slabs under Sustained Load 4.4.4.3 Discussion 4.4.5 Analysis of Long-term Flexural Cracking Tests using the Localized Cracking Model Non-local Smeared Crack Model 4.4.5.1 Four-point Bending Beam Tests under Sustained Load 4.4.5.2 Uniformly Loaded One-way Slabs under Sustained Load 4.4.5.3 Discussion 4.4.6 Summary for Analysis of Long-term Flexural Cracking Tests 4.5 Long-term Restrained Deformation Cracking Tests 4.5.1 Introduction 4.5.2 Analysis of Restrained Deformation Cracking Tests and Material Properties 4.5.3 Comparisons of Numerical and Experimental Results 4.5.4 Discussion 4.6 Other Numerical Examples 4.6.1 Continuous Beams Subjected to Long-term Sustained Load 4.6.2 Time-dependent Forces Induced by Supports Settlement of Continuous Beams 4.6.3 Slender Columns Subjected to Long-term Eccentric Axial Loads CHAPTER 5 NUMERICAL EXPERIMENTS 5.1 Introduction 5.2 Description of Numerical Experiments 5.2.1 Beam Specimens 5.2.2 Slab Specimens 5.2.3 Testing Method 5.2.3.1 Test Series A to J: Material and Environmental Parameters 5.2.3.2 Test Series K: Amount of Shear Reinforcement 5.2.3.3 Test Series L: Impact of 500 MPa Steel Reinforcement 5.2.3.4 Test Series M: Load Histories 5.3 Presentation and Discussion of Results

4.4.3

156 156 160 162 169 170 179 182 187 188 193 196 201 203 203 205 208 217 220 220 225 230

237 239 241 244 246 247 250 251 252 256

Test Series A Bottom Concrete Cover Test Series B Diameter of Tensile Reinforcing Steel Test Series C Quantity of Tensile Reinforcement Test Series D Quantity of Compressive Reinforcement Test Series E Tensile Strength of Concrete Test Series F Bond Strength between Steel and Concrete Test Series G Concrete Tensile Strength Fluctuation Limit Test Series H Magnitude of Creep Test Series I Magnitude of Shrinkage Test Series J Bond Creep Test Series K Quantity of Shear Reinforcement Test Series L Impact of 500 MPa Steel Reinforcement Test Series M Load Histories 5.3.13.1 Comparisons between LH-2 and LH-1 5.3.13.2 Comparisons between LH-5 and LH-2 5.3.13.3 Comparisons between LH-3 and LH-4 5.3.14 Section Geometry and Boundary Conditions 5.4 Summary CHAPTER 6 SUMMARY AND CONCLUSIONS 6.1 Summary 6.2 Conclusions 6.3 Recommendations for Future Research APPENDIX A FE IMPLEMENTATION OF RATE OF CREEP METHOD APPENDIX B ILLUSTRATION OF TREATMENT FOR INELASTIC PRE-STRAIN
BY SIMPLE HAND CALCULATION

5.3.1 5.3.2 5.3.3 5.3.4 5.3.5 5.3.6 5.3.7 5.3.8 5.3.9 5.3.10 5.3.11 5.3.12 5.3.13

257 261 264 267 270 274 277 279 285 285 289 291 293 293 296 298 301 301

305 308 311 313

318

APPENDIX C CEB-FIP MODEL CODE 1990 CREEP AND SHRINKAGE MODELS REFERENCES 324 327

xi

NOTATION
A Ac Acp , Bcp Ac.eff Ae A fct , B f ct As Asc Ash , Bsh Ast Asv A0 B B C da D Dc D cts D c12 Db Area; empirical time-dependent parameter. Cross-sectional area of concrete. Parameters for time-dependent variation of creep coefficient. Effective area of concrete in tension. Surface area of finite element; tangential contact surface area for bond element. Parameters for time-dependent growth of concrete tensile strength. Cross-sectional area of steel. Cross-sectional area of compressive steel. Parameters for time-dependent development of shrinkage strain. Cross-sectional area of tensile steel. Cross-sectional area per stirrup. Negative infinity area of retardation spectrum. Empirical time-dependent parameter. Strain-displacement matrix. Creep compliance function. Maximum aggregate size. Material elasticity matrix. Constitutive matrix for concrete. Constitutive matrix for tension stiffening. Material constitutive matrix in principal directions. Bond constitutive matrix. Elastic stiffness matrix. Elasto-plastic stiffness matrix for plasticity-based model. Constitutive matrix for steel. Secant stiffness matrix. Secant stiffness matrix in material local axis.

De
D ep Ds

Dsec , D sec
Dsec nt

xii

sec D12

Secant stiffness matrix in principal axis. Material tangent constitutive matrix. Modulus of elasticity. Secant moduli for bond-split and bond-slip, respectively. Secant modulus of bond. Initial modulus of concrete. Elastic modulus of concrete at 14 days and 28 days, respectively. Secant modulus at peak of concrete stress-strain curve. Secant moduli for tension stiffening in x, y directions, respectively. Unloading modulus for concrete in compression. Concrete secant moduli in major and minor principal directions, respectively. Initial elastic modulus of reinforcing steel. Secant modulus of reinforcing steel. Secant modulus. Secant moduli for steel reinforcement in x, y directions, respectively. Unloading modulus for reinforcing steel. Unloading modulus for concrete in tension. Hardening moduli for reinforcing steel. Asymptotic modulus of concrete. Elastic modulus of -th Kelvin chain unit. Yield function for plasticity-based model; damage loading function for continuous damage model; local state variable. Function; time-dependent function. Flexural tensile strength of concrete. Mean compressive strength of concrete. Cracking stress under tension cut-off regime. Direct tensile strength of concrete. Direct tensile strength of concrete at age t days.
xiii

D tan E Ebn , Ebt Eb. sec Ec E c.14 , E c.28 Ecpk Ectsx , Ectsy Ecu Ec1 , Ec 2 Es
sec Es

Esec
E sx , E sy E su Etu E w , Eu E0 E f F f cf f cm f cr f ct f ct.t

f ct 0 f cu f sy f sw , f su f c' f c* f F Fe G Gc12 gf Gf gp G h hc hs i, j J J k K Ke K sec K tan K0 lch

Concrete tensile strength at zero crack opening rate. Compressive strength in uniaxial compressive stress-strain curve. Yield stress of steel reinforcement. Hardening stresses of steel reinforcement. Characteristic compressive strength of concrete. Biaxial compressive strength of concrete. Weighted average state variable for non-local model. Structural equivalent pre-strain nodal force vector. Element equivalent pre-strain nodal force vector. Shear modulus. Concrete secant shear modulus in principal directions. Fracture energy density. Fracture energy. Plastic potential function for plasticity-based model. Matrix for inclusion of Poissons effect to biaxial stress. Average width of fracture process zone; volume associated with viscous strain. Crack band width. Notational size of concrete member. Counters. Compliance function. Jacobian matrix. Decay factor. Structural stiffness matrix. Element stiffness matrix. Secant stiffness matrix. Tangent stiffness matrix. Initial tangent stiffness matrix. Material characteristic length.

xiv

l fct L l, m, n L M sw , M u N N P pb pe Pe ps q 2 , q3 , q 4 Q r R RH R s sa si st s1 , s2 , s 3 s rm s rmx , s rmy s rm0

Random fluctuation limit of concrete tensile strength. Function for retardation spectrum; length of truss element; length of bond element. Orthonormal base vectors. Differential operator matrix. Moment due to self-weight and moment capacity, respectively. Total number of Kelvin chains; total number of elements; shape function. Displacement interpolation matrix. External structural nodal force vector. Body force vector. Nodal force vector. External element nodal force vector. Surface traction vector. Empirical parameters for solidification theory of creep. Structural internal force vector. Distance between target point and source point in non-local analysis. Relaxation function; non-local interaction radius. Relative humidity. Out-of-balance force vector. Slip between concrete and reinforcing steel. Axial length of truss element. Instantaneous slip. Time-dependent slip. Slips defining CEB-FIP bond-slip model. Crack spacing. Crack spacings of an orthogonally reinforced concrete membrane element in x and y directions, respectively. Maximum crack spacing. Location vectors of neighbourhood strains. Time or age of concrete. Temperature.

s
t T

xv

te t0 t' Tb TB T u, v ua uc uti , u ni u ue u V Ve V' wcr wu

Thickness of plane stress element. Age at first loading. Variable for age at loading. Bond transformation matrix. Diagonal bond transformation matrix. Strain transformation matrix. Nodal displacements corresponding to x and y directions, respectively. Nodal axial displacements. Perimeter of concrete member. Element nodal displacements parallel to and normal to bond element, respectively. Structural nodal displacement vector. Element nodal displacement vector. Continuous field displacement vector. Volume of structure. Volume of finite element. Volume of structure in displaced configuration. Crack opening displacement. Crack opening at which stress transfer at fictitious crack vanishes. Location vectors of local strain. Damage energy release rate. Weight function; ratio of major and minor principal compressive stresses. Tension softening parameters. Normalized weight function. Shear retention factor; confinement factor; strength reduction factor. Confinement factor in the major principal direction. Confinement factor in the minor principal direction. Slip at which reinforcing steel starts to yield. Strain. Concrete strain between cracks.

x Y

1 , 2 , 3 ' 1 2 y

betw.cr

xvi

Concrete strain. Concrete elastic strain. Concrete elastic strain corresponding to c.un . Instantaneous concrete strain. Creep strain. Concrete strain corresponding to peak stress in compressive stressstrain curve. Concrete cracking strain. Concrete cracking strain corresponding to c.un . Concrete strain corresponding to c.un on stress-strain curve. Microplane strain components of microplane model. Mean strain in an element. Shrinkage strain. Concrete strain corresponding to peak stress in tensile stress-strain curve. Concrete cracking strain when cohesive stress between crack faces vanishes. Concrete strains in major and minor principal directions, respectively. Equivalent uniaxial concrete strains in major and minor principal directions, respecitively. Concrete elastic strain in major and minor principal directions, respectively. Concrete cracking strain in major and minor principal directions, respectively. Adjusted concrete strain corresponding to peak stress in biaxial compressive stress-strain curve.

ce
ce.un

ci , inst cp
cpk cr cr.un c.un L,M ,N m

sh tpk
u 1 , 2 1u , 2u

1u.e , 2u.e 1u.cr , 2u.cr


* cpk

* sh

Final shrinkage strain. Concrete non-local cracking strain. Modified equivalent uniaxial strains due to non-locality in major and minor principal directions, respectively.

cr 1u , 2u

xvii

1u.cr , 2u.cr
~

Concrete non-local cracking strain in major and minor principal directions, respectively. Equivalent strain of continuous damage model. Strain vector. Concrete instantaneous strain vector. Principal strain vector. Concrete elastic principal strain vector. Creep strain vector. Concrete elastic strain vector. Concrete cracking strain or plastic strain vector. Viscous strain vector. Strain vector in material local axis. Shrinkage strain vector. Viscoelastic strain vector. Pre-strain vector. Non-local strain vector. Concrete non-local cracking strain vector. Creep coefficient. Final creep coefficient. Microscopic creep compliance function. Viscoelastic microstrain of -th Kelvin chain unit. Viscoelastic microstrain vector. Apparent macroscopic viscosity associated with viscous component of creep. Effective viscosity of solidified matter associated with viscous component of creep. Viscosity of -th Kelvin chain unit. Hardening or softening parameter for plasticity-based model; historydependent parameter for continuous damage model. Ratio of crack spacing to maximum crack spacing. Counter for Kelvin or Maxwell chain unit.
xviii

ci 12 c12 cp

ce , e cr , p
f nt sh v 0

cr

Poissons ratio; volume associated with viscoelastic strain. Poissons ratio in major principal direction resulting from stress in minor principal direction and vice versa, respectively. Angle; angle between global x and local n axes; angle between global x and major principal axes. Ratio of cross sectional area of steel to cross sectional area of concrete. Ratio of area of tensile steel to effective area of concrete in tension. Element reinforcement ratios in x and y directions, respectively. Stress. Stress in concrete.

12 , 21 eff x , y

cn , ct , cnt Concrete stresses in local n (normal to crack), t (crack direction) and cnt (shear) directions, respectively. ctsm
Mean concrete tension stiffening stress. respectively.

ctsmx , ctsmy Mean concrete tension stiffening stresses in the x and y directions, ctsm1 c.un c1 , c 2 et ext int
Mean concrete tension stiffening stress in major principal direction. Concrete stress just before commencement of unloading. Concrete stress in major and minor principal directions, respectively. Function of cohesive stress versus cracking strain. External stress. Internal stress. Local stress in concrete. Out-of-balance stress. Stress in reinforcing steel. Mean steel stress. Minimum steel stress. Steel stress at crack or maximum steel stress. Steel stresses in global x, y directions, respectively. Concrete stresses in global x, y and xy (shear) directions, respectively. Function of cohesive stress versus crack opening displacement.

loc
out s sm s min sr sx , sy x , y , xy wt

xix

nt 12

Stress vector. Stress vector in material local axis. Principal stress vector. Bond stress between concrete and reinforcing steel. Bond stresses before yielding and after yielding of steel, respectively. Maximum bond stress in local bond stress-slip curve. Failure bond stress. Retardation time of -th Kelvin chain unit. Damage variable. Natural coordinate system. Angle of orientation of truss or bond element from the x-axis. Temperature-dependent shift function. Diameter of reinforcing bar. Diameter of tensile reinforcing bar.

b b0 , b1 max f

, s
st

xx

CHAPTER 1 INTRODUCTION

1.1 Background and Significance


Reinforced concrete is a composite material made up of steel reinforcement embedded in hardened concrete. The two materials are inter-complimentary. Concrete is ideal for withstanding compressive forces and steel reinforcement is ideal for carrying tensile forces, thereby compensating for the low tensile strength of concrete. For the steel reinforcement to effectively carry the internal tensile forces, the tensile concrete must crack. Under normal in-service conditions, cracking is inevitable in many reinforced concrete structures. In most structures, cracking occurs due to the application of external service loads and due to restrained deformation. Although inclusion of steel reinforcement does not prevent this type of cracking, it does help distributing cracks evenly over the cracked regions and therefore effectively controls the development of cracks. There are several types of cracking that cannot be controlled by reinforcement. These include cracking originating from the development of internal pressure in concrete due to corrosion of reinforcement, plastic shrinkage of concrete that occurs in the first few hours after casting and expansion of concrete associated with chemical reactions. Cracking caused by these factors can, instead, be prevented by good quality control during construction of the structure. In this study, only cracking arising from external loads and restrained deformation (for instance, shrinkage induced movements and support settlement), which can be effectively controlled by adequate inclusion of steel reinforcement, is investigated. The ability of reinforcement to distribute cracks depends greatly on the quality of bond between the reinforcing steel and the concrete. The composite interaction between the two materials is established and maintained by bond, which effectively transfers load between the steel and the concrete. The main mechanism in the development of

bond is the mechanical interaction between the ribbed or deformed surface of the reinforcing bar and the concrete. However, other mechanisms such as surface friction and chemical adhesion also play a role. The quality of bond has a prominent influence on crack formation and hence affects the spacing between cracks and the crack width. Cracking results from tension caused by external loads and from tension caused by restrained deformation due to the effects of shrinkage and temperature. In addition, the bond in the vicinity of a crack under sustained loads deteriorates due to creep of the concrete adjacent to the reinforcing steel. This results in an increase in slip between the steel and concrete with time. The deterioration of bond under long-term loads further complicates the cracking process, as does the gradual build-up of tension caused by restraint to shrinkage. As a consequence, cracking is time-dependent, with the extent and width of cracks gradually increasing with time under sustained loading. For a structure to remain serviceable, crack widths must be small enough to be acceptable from an aesthetic point of view, small enough to avoid waterproofing problems and small enough to prevent the ingress of water that may lead to corrosion of the reinforcement. Excessively wide cracks can even provoke public concern for the safety of the structure. Crack control is therefore an important aspect of the design of reinforced concrete structures at the serviceability limit state and the topic has received much research attention. The design procedures for crack control can be divided broadly into two alternative types. The first type is by calculating the crack width explicitly and comparing the calculated magnitude with the code stipulated crack width limits. In the second approach, crack control is deemed to be satisfactory as long as some specific detailing requirements are met, such as permissible maximum distance between reinforcing bars in the tension region, maximum bar diameters for a specific steel stress, minimum reinforcement area and so on. For practising structural engineers, the second approach may seem appealing due to its simplicity. Nevertheless, the simplified procedures provided in most design codes are generally less than adequate. In addition, code methods have been developed, in the main, from laboratory observations of the instantaneous behaviour of reinforced concrete members under load and fail to account for the time-dependent development of cracking and the inevitable increase in crack widths with time due primarily to

shrinkage. As a result, even the use of the first approach (explicit calculation of crack width) may lead to significant error in the estimation of the final crack width. The inability to recognize and quantify the non-linear effects of cracking, creep and shrinkage can lead to excessive deflections and crack widths and miscalculation of support reactions. Experimental data relating to the time varying distribution of cracking, in particular the final crack spacings and crack widths are scarce in the open literature. The development of a rational and reliable design procedure for control of cracking, however, requires considerable experimental data for the purpose of investigating and understanding the critical factors that affect time-dependent cracking. Two alternatives may be used to yield the required data for this purpose. The conventional experimental approach involves the fabrication of numerous test specimens and testing them in the laboratory for a specific objective. Long-term cracking tests are, however, costly not only in terms of time but also resources of the laboratory since engagement of a large area in the laboratory over a long period of time is necessary. Alternatively, numerical techniques such as the finite element method may be employed to simulate a wide range of virtual long-term cracking tests so as to facilitate a parametric investigation. This is an efficient and inexpensive method compared to the conventional experimental approach. Nevertheless, to achieve this the critical factors that affect accuracy of modelling must be identified. Realistic material models that take account of cracking, creep and shrinkage of concrete and bond-slip between steel and concrete are the keys to a reliable numerical simulation of timedependent cracking of reinforced concrete structures.

1.2 Objective and Scope


The main objective of the work reported in this thesis is to investigate the timedependent behaviour of reinforced concrete structures using the finite element method, with particular interest in the formation of cracks, both in terms of spacing and width of the cracks. The following tasks were undertaken to achieve the aforementioned objective:
3

(1) Formulation of a two-dimensional plane stress finite element model and incorporation of material constitutive models that can realistically characterize cracking, creep and shrinkage of concrete and bond-slip between concrete and reinforcing steel, thereby facilitating numerical modelling of reinforced concrete structures under service load conditions. (2) Calibration and evaluation of the finite element model using the test data obtained from the experimental program (Gilbert and Nejadi, 2004; Nejadi and Gilbert, 2004) conducted in parallel with this study and other test data obtained in the literature. (3) Critical assessment of the time-dependent cracking models developed in this study and comparisons of the pros and cons of each model. (4) Investigation of the effects of various parameters on time-dependent cracking of reinforced concrete structures by running a series of numerical parametric experiments on beam and slab specimens. The time-dependent finite element models developed in this study account for the time effects of concrete, namely creep and shrinkage. Although temperature is, in some respects, considered as one of the factors affecting the time-dependent behaviour of reinforced concrete structures, it is not within the scope of the work described in this thesis and therefore thermal effects are not considered. In other words, the ambient temperature is assumed to remain constant for all simulations carried out in this work. The finite element models were developed to investigate time-dependent cracking under service load conditions. Under these conditions, compressive stresses in concrete are rarely in excess of 40% of the concrete compressive strength and, for this reason, a linear creep model was implemented into the finite element models. Nevertheless, the finite element models developed here are capable of computing the instantaneous behaviour to failure of reinforced concrete structures under the full range of loading, or, subjected to certain limiting assumptions, the time-dependent behaviour of members under loads up to failure can also be established. In addition to material non-linearity, the finite element models can also account for the complication arising from geometric non-linearity. This greatly facilitates

simulation for problems related to creep buckling, which is important in stability analysis of slender columns.

1.3 Outline of Thesis


Chapter 2 gives a review of the previously published works that are relevant to this study. This chapter is divided into two parts. The first part is an overview of the literature dealing with instantaneous and time-dependent material behaviour in general. The second part deals with previously published approaches for non-linear modelling of reinforced concrete structures using the finite element method. In Chapter 3, the formulations of the time-dependent finite element models are presented. Firstly, various types of concrete cracking models are described. This is then followed by a description of the material constitutive models employed in the finite element studies. The finite element implementation of the time-dependent models and the non-linear solution procedures are also outlined. Finally, the formulations of the finite elements engaged in this study are shown. Chapter 4 evaluates the finite element models described in the previous chapter. A mesh sensitivity test is carried out to verify the objectivity of the results calculated by the proposed cracking models. The creep model is also evaluated by testing a simple uniaxially loaded specimen subjected to variable load histories. The accuracy of the finite element models is examined by comparing the numerical results with existing laboratory test data, both in terms of formation of cracks and deformation. Chapter 5 describes a parametric investigation using the proposed finite element model to study time-dependent cracking. A series of controlled parametric numerical tests is devised and analysed. The results of the numerical tests are discussed and the effects of each parameter are examined. Chapter 6 summarizes the conclusions drawn from the investigation of this work and a proposal for future research is made.

Appendix A presents a finite element model developed in the early stage of this study. The model employs a relatively simple creep model based on the rate of creep method. A simulation example is shown and the results are compared with test data. In Appendix B, a simple hand calculation is shown to demonstrate the method used in the finite element procedures to account for inelastic pre-strains such as creep and shrinkage. Finally, Appendix C presents the creep and shrinkage models of the CEB-FIP Model Code 1990 (1993).

CHAPTER 2 LITERATURE REVIEW

2.1 Introduction
The knowledge for producing concrete and the basic idea of using reinforcement for building structures has existed for thousands of years. However, due to the heterogenous nature of concrete and the composite actions of the constituent materials, research is still undertaken on reinforced concrete structures in order to gain a deeper understanding of the complex behaviour of the individual and composite materials. This chapter consists of two major parts. In the first part, a state-of-the-art review is given on the properties of concrete and steel reinforcement as well as the interaction between concrete and reinforcement. The material properties and composite action of the two materials are of vital importance in the modelling and analysis of reinforced concrete structures. The time-dependent aspect of concrete is discussed and the influencing factors are identified. The second part is a review of the non-linear modelling methods for reinforced concrete structures which includes discussions of the different cracking models, material constitutive models and time-dependent modelling techniques.

2.2 Instantaneous Behaviour of Concrete


2.2.1 Uniaxial Compression

Concrete is a highly non-linear material in uniaxial compression. The properties of concrete in uniaxial compression are obtained from cylinder tests or cube tests. A standard cylinder of 300 mm height and 150 mm diameter is used for cylinder tests
7

under standard conditions in Australia and North America, while tests on 150 mm cubes are often used in European and Asian countries. Figure 2.1 shows a typical uniaxial compression stress-strain curve. Concrete is a heterogeneous material, the shape and the peak of the stress-strain curve varies greatly and are dependent on the proportions and properties of the constituents, the size and shape of the specimen, the rate of loading and also the age of the concrete. In Figure 2.1 the response of concrete may be taken to be linear-elastic up to 30% ~ 40% of the peak stress, f cu (often termed the compressive strength of concrete). Beyond this point, concrete behaves in a non-linear manner up to the peak stress. At levels of stress that are just above the linear-elastic range, microcracking between the aggregates and mortar becomes apparent and bond cracks are formed (Mindess et al., 1981). As the stress increases further into the range 70% to 90% of the compressive strength, microcracks start to open and bridge the bond cracks. This leads to the formation of continuous crack patterns and the stress eventually reaches the maximum compressive stress. Immediately after the peak stress, the concrete undergoes strain softening, which is depicted in Figure 2.1 as the descending branch of the curve. Strain softening of concrete in compression is a complicated process, it is dependent on the size of the specimen and the strength of the concrete (Kaufmann, 1998). The softening branch of the stress-strain curve in compression is steeper for longer specimens than for shorter specimens and is due mainly to the localization of deformations while other parts of the specimen are unloading. The shape of the unloading curve depends on the concrete strength, with higher strength concrete exhibiting a more brittle response (i.e. a steeper unloading curve). This is attributed to the fact that the specific fracture energy of concrete in compression does not increase much with the concrete strength. With the area under the stress-strain curve roughly constant, higher strength concrete must have a steeper descending curve. Many uniaxial stress-strain relationships for concrete in compression have been proposed in the literature, such as the well-known expressions developed by Hognestad (1951), Desayi and Krishnan (1964), Saenz (1964) and Thorenfeldt et al. (1987).

Uniaxial stress

fcu

0.4 fcu
Ec 1 Longitudinal strain

cpk

Fig. 2.1 - Uniaxial compression curve.

2.2.2

Uniaxial Tension

The strength of concrete in tension is much lower than the strength in compression. The ratio of tensile strength to compressive strength is between 0.05 and 0.1, as obtained experimentally by Johnson (1969). The tensile strength of concrete is normally evaluated using the split cylinder test (called the Brazilian test) for the indirect tensile strength, while the modulus of rupture test is used to evaluate the flexural tensile strength f cf . The direct tensile strength f ct is about 0.9 times the indirect split cylinder strength and about 60% of the modulus of rupture and is often adopted in situations where the stress field around the locations where cracking occurs is uniformly distributed. An example is the cracking analysis in finite element modelling. For flexural cracking, the flexural tensile strength is the quantity required. Normally for general engineering purposes, the tensile strength may be expressed as a function of the compressive strength. As recommended by the Australia Standard AS 3600-2001, the lower characteristic direct tensile strength at age 28 days may be estimated by f ct = 0.4 f c' and the lower characteristic flexural tensile strength is given by f cf = 0.6 f c' (2.2) (2.1)

where f c' is the characteristic compressive strength of concrete (obtained from cylinder tests) and all strength quantities in both Eq. 2.1 and Eq. 2.2 are in MPa. A full stress-deformation response of concrete in tension cannot be determined using the aforementioned tests. A direct tension test is required in order to capture the full deformation of a concrete member in tension. This is a highly sensitive test that requires the specimen to be sufficiently short, tested using very stiff testing machines, using precise measuring devices and under displacement controlled conditions. The typical stress-elongation response of concrete in tension is shown in Figure 2.2. The pre-peak behaviour of concrete in tension is mostly linear-elastic except for the stresses near the peak stress. At about 60% to 80% of the peak stress, microcracks begins to form fairly uniformly throughout the specimen. The overall response of the concrete member becomes softer and exhibits highly non-linear behaviour. Due to the quasibrittle nature of concrete, tensile stress in concrete does not reduce abruptly to zero after the peak. On the contrary, damage in the specimen starts to localize into a fracture process zone at the weakest section while the rest of the specimen undergoes unloading. In the fracture process zone, concrete is able to transfer stress across the crack opening direction due to bridging of aggregate particles, the tensile stress then drops gradually with increasing deformation until a complete crack is formed. At this point, no further tensile stress can be taken by the concrete member and this eventually leads to a complete tensile failure. This process is known as strain localization and the concrete is said to undergo tension softening. However, a direct experimental evaluation of the full stress-strain curve of concrete in tension has not been possible to date. This is attributed to the localized nature of fracture in concrete. The elongation of the specimen is contributed to by the fracture process zone while the other parts of the specimen actually cause a reduction in length due to elastic unloading. Therefore, a direct measurement of the stress-strain curve for concrete specimens in tension, even from the same concrete mix, would give unobjective responses for specimens of different lengths. To overcome this problem, the prediction of cracking in concrete must not be based solely on a strength criterion but must also take into account the energy dissipated in cracking of concrete. Hence the use of fracture mechanics should be considered.

10

t fct
Ec

t wcr wcr wcr l


1

t fct
Ec

fct

Gf t wcr

ltpk

tpk

(a)

(b)

(c)

Fig. 2.2 - Uniaxial tensile response: (a) stress-elongation curve and the influence of specimen size to tension softening; (b) stress-strain curve for the specimen outside the strain localization zone; (c) post-peak stress-crack opening displacement curve, G f is known as the fracture energy and is equal to the area under the softening curve.

2.2.3

Biaxial Loading and Failure Criteria

Concrete subjected to biaxial loading exhibits a different response to that under uniaxial loading. Since many practical engineering problems may be simplified to a state of plane stress, much research has been carried out to investigate the behaviour of concrete in biaxial states of stress for both normal strength and high strength concretes (Kupfer et al., 1969; Liu et al. 1972; Nelissen, 1972; Tasuji et al.; 1979, van Mier, 1986; Nimura, 1991). Kupfer et al. were probably the first researchers to conduct reliable biaxial tests on concrete. They tested different concretes with compressive strengths between 19 MPa and 60 MPa and found that the biaxial strength envelopes constructed in terms of the ratio of the orthogonally applied stresses and the compressive strengths, are similar for all the concretes tested. A typical biaxial strength envelope of Kupfer et al. (1969) is shown in Figure 2.3. The biaxial state of stress may be divided into three loading combinations, namely biaxial compression, tension-compression and biaxial tension. Under biaxial compression, the compressive strength is increased due to the presence of the lateral compressive stress. The shape of the stress-strain curves is similar to that under uniaxial

11

compression but with a higher compressive strength up to about 1.25 times the uniaxial compressive strength, depending on the magnitude of the lateral compressive stress. Kupfer and Gerstle (1973) proposed an equation to model the biaxial compressive strength envelope in which the maximum compressive strength 2c may be obtained by

2c =

1 + 3.65 (1 + )
2

f c'

(2.3)

where = 1 / 2 ; and 1 and 2 are the principal major and minor stresses respectively. The in-service behaviour of a concrete structure usually involves regions that lie in the tension-compression and in the biaxial tension states of stress. Under conditions of tension-compression, which is the second or fourth quadrant in Figure 2.3, concrete exhibits a considerable reduction in compressive strength with a small increase in transverse tension. Kupfer and Gerstle suggested a straight-line strength envelope, the tensile strength 1t decreasing with an increasing compressive stress and given by

1t = 1 0.8

2
f c'

f ct

(2.4)

For biaxial tension, Kupfer and Gerstle suggested the use of a constant uniaxial tensile strength, which is essentially identical to the Rankine failure criterion and this is generally adopted in modelling cracking in concrete.

12

0.2 0.0 -0.2 -0.4

2 / f c ' -0.6
-0.8 -1.0 -1.2 -1.4 -1.4 -1.2 -1.0 -0.8 -0.6 -0.4 -0.2 0.0

1 / f c '

0.2

Fig. 2.3 - Biaxial concrete strength envelope (Kupfer et al., 1969).

2.3 Time-dependent Behaviour of Concrete


The discussion of the behaviour of concrete has so far concentrated on the instantaneous response. Concrete is indeed a type of viscoelastic cementitious material that is characterized by time-dependent behaviour. When a concrete specimen is subjected to a sustained load, it undergoes an immediate deformation. This is followed by a further deformation with increasing time. The increase in deformation with time is not negligible and may be several times larger than the instantaneous deformation. Under constant temperature condition, the additional time-dependent response is attributed to the effects of both creep and shrinkage of concrete. Considering a concrete specimen subjected to a uniaxial load at a constant ambient temperature, the total strain of the specimen at time t may be decomposed into the following three strain components

(t ) = ci (t ) + cp (t ) + sh (t )

(2.5)

in which ci (t ) is the instantaneous strain, cp (t ) is the creep strain and sh (t ) is the shrinkage strain. The decomposition of the total strain into individual components in

13

Eq. 2.5 is a simplification. In reality, creep and shrinkage are interdependent phenomena. However, this assumption is generally acceptable for engineering purposes. Figure 2.4 shows the deformation of a specimen loaded uniaxially in compression at age t 0 at a constant stress level. Shrinkage strain develops at the commencement of drying (here defined as t = 0) while creep strain begins to develop after the specimen is stressed. For loading within the elastic range of the stress-strain relationship, the instantaneous strain is equal to the elastic strain. The dashed line in Figure 2.4b denotes the true elastic strain of the specimen due to the increase in the elastic modulus with age.

sh Shrinkage strain from t0 to t

t0

(a)

Creep strain

Shrinkage strain

Nominal instantaneous strain t0 t

(b)

True instantaneous strain

Fig. 2.4 - Time-dependent strain development: (a) development of shrinkage strain after commencement of drying; (b) change in the different strain components for a specimen after being subjected to a sustained load first applied at t 0 .

14

2.3.1

Creep

Creep of concrete was first reported by Hatt (1907). It is a time-dependent deformation that develops at a decreasing rate under a sustained loading. At low stress levels, creep originates in the hardened cement paste while aggregates provide only restraint to the deformation. The cement paste consists of solid cement gel with numerous capillary pores, which is made of colloidal sheets formed by calcium silicate hydrates and contain evaporable water. The mechanism of creep in concrete is disputed and no satisfactory theory is available to describe the formation of creep. However, it is generally believed that creep is caused by changes in the solid structure, due to the disordered and unstable nature of the bonds and contacts between the colloidal sheets (Baant, 1982). At high stress levels, interfacial cracks begin to form between the cement paste and the aggregate and this finally results in a further increase in deformation. The influence of stress level on creep is illustrated in Figure 2.5. The applied constant stress is plotted against the corresponding creep strain. At service load conditions, stress in concrete seldom exceeds 50% of the compressive strength. At these stress levels, the creep response is approximately linear, and therefore, creep is often assumed to be proportional to stress. c / fc 1.0 (t - t) = 0 Failure limit

0.5

(t - t) = infinity Linear range (ci+ cp)

Creep limit

Fig. 2.5 - Influence of stress level and sustained duration on concrete mechanical strain (Gilbert, 1988).

15

Creep may be divided into two components, namely basic creep and drying creep. Under conditions of hygral equilibrium (no moisture exchange with the ambient medium and therefore no drying shrinkage), the gradual increase in strain with time in a loaded specimen is known as basic creep. In a drying specimen, creep occurs simultaneously with shrinkage, and is significantly higher than basic creep in a sealed specimen. The extra creep that occurs in excess of the basic creep is termed drying creep or the Pickett effect. Drying creep is primarily caused by stress-induced shrinkage (Baant and Chern, 1985, Baant, 1988).

2.3.1.1

Factors affecting Creep

The magnitude of creep in concrete structures is influenced by many factors. These factors can be categorized into two groups: the technological parameters and the external parameters (CEB, 1997). The former refers to the parameters associated with the particular concrete mix such as water-cement ratio, mechanical properties and quantity of aggregates and type of cement. Lorman (1940) suggested that creep is approximately proportional to the square of the water-cement ratio when all other factors are kept constant. This is due to the fact that the cement paste content varies for different water-cement ratios, and the cement paste is the main ingredient that influences creep deformation of concrete. On the other hand, creep may also be reduced by using stiffer aggregates and by increasing the aggregate content, thereby inducing a higher restraining force on the cement paste. The second group of parameters affecting creep are those associated with the external conditions, such as age of concrete at first loading, relative humidity, and shape and size of specimen. The observations of Davis et al. (1934) and Glanville (1933) show that the rate of creep during the first few weeks under load is much greater for concrete loaded at an early age than for older concrete. Figure 2.6 shows typical creep curves for specimens loaded at different ages. It can be observed in Figure 2.6b that for a specimen loaded at a very old age, the strain level tends to approach the value

/ E0 in which E0 is the asymptotic elastic modulus of concrete equivalent to Ec () .


The relative humidity has a large influence on creep. It was observed that drying

16

concrete has a higher rate of creep and a higher final creep level than concrete that is stored in an environment with a higher relative humidity. Creep also depends on the shape and size of the specimen. Investigation shows that creep decreases with an increase in size of the specimen. However, the size factor becomes insignificant if the thickness of the specimen exceeds about 900 mm (Neville et al., 1983). For the influence of shape, Chivamit (1965) found that the initial rate of creep of a cruciform section is higher than that of a circular section of the same cross-sectional area. This may be attributed to the large surface area that exists in the cruciform section, with creep occuring more rapidly on the drying surface than in the interior of a specimen. Nevertheless, the influence of shape appears to be significantly smaller than the influence of size and therefore it is often neglected for engineering purposes.

(ci+ cp)

(ci+ cp)
t = t 0 t = t1 t = t2 t = t3 /Ec(t) /E0

t0

t1

t2

t3

log (t - t)

(a)

(b)

Fig. 2.6 - Typical creep curves for a specimen subjected to a sustained stress at various ages t ' : (a) creep curves on the normal time scale; (b) creep curves on the logarithmic time scale.

2.3.1.2

Creep Recovery

In practice, concrete structures are seldom subjected to a single constant load throughout their service life. Other than the self-weight, service loads generally vary with time depending upon the function of the structure at different stages in its life.

17

Consider a specimen subjected to a sustained compressive stress, when the stress is completely removed, it immediately undergoes an elastic unloading known as the instantaneous recovery, as depicted in Figure 2.7. If the increase in elastic modulus with time is ignored, the magnitude of the instantaneous recovery is equal to the instantaneous strain when the stress is first applied to specimen. The process is followed by a gradual reduction in strain, which is referred to as creep recovery. Creep deformation is predominantly an inelastic process. A large amount of creep strain developed over the loading period is irrecoverable, with only a relatively small amount being recoverable through creep recovery (generally less than 30% of the total creep strain).

(ci+ cp)

Instantaneous recovery Creep strain Creep recovery Instantaneous strain t0 Irrecoverable residual strain t

Fig. 2.7 - Recovery of strain components upon removal of external load.

2.3.1.3

Principle of Superposition

It is well known that under service conditions, the maximum stress in a concrete structure seldom exceeds 50% of the compressive strength and creep may be assumed to be proportional to stress. For a concrete member subjected to a constant sustained uniaxial compressive stress , the sum of the instantaneous strain, creep strain and shrinkage strain may be written as

(t ) = J (t , t ' ) + sh (t )

(2.6)

18

where J (t , t ' ) is the compliance function (or creep function) which is defined as the strain at time t produced by a unit sustained stress applied at age t ' . The compliance function may further be expressed in an expanded form as follows J (t , t ' ) = 1 + C (t , t ' ) Ec (t ' ) (2.7)

where Ec (t ' ) is the elastic modulus of concrete at age t ' and C (t , t ' ) is the creep compliance. The first term in Eq. 2.7 represents the instantaneous deformation. By introducing the creep coefficient (t , t ' ) defined as (t , t ' ) = C (t , t ' ) Ec (t ' ) , Eq. 2.7 may be rewritten as J (t , t ' ) = 1 [1 + (t , t ' )] Ec (t ' ) (2.8)

In a reinforced concrete structure, the concrete stress at any point varies continuously throughout its service life even though the loads may be kept constant with time. This is due to the redistribution of stress between the concrete and the steel reinforcement caused by the gradual development with time of creep and shrinkage strains in the concrete. By exploiting the advantage of the linear relationship between creep and stress in the service range, the principle of superposition is generally adopted to compute the deformation of the structure when subjected to varying stress histories. The principle of superposition for an aging material was derived by Volterra (1913, 1959) and was first applied to concrete by McHenry (1943). The principle states that the total strain at time t is calculated by summing the strain increments produced by each stress increment d applied at any age t ' , and each strain increment is not affected by the stresses applied earlier or later. The principle of superposition is illustrated in Figure 2.8. This may be expressed as a Stieltjes integral (Baant, 1982) and the uniaxial total uniaxial strain over a period t is given by adding the shrinkage strain to the integral as
t

(t ) = J (t , t ' ) d (t ' ) + sh (t )
0

(2.9)

19

(a)

(ci+ cp)

strain curve 1

t0

t1

t0 (ci+ cp)

t1

(b)

t0 2 t1 t

t0

t1

strain curve 2

(c)

(ci+ cp)

1 1 + 2 t0 t1 t t0 strain curve 1 + strain curve 2 t1 t

Fig. 2.8 - The principle of superposition: (a) constant applied stress 1 and the corresponding strain; (b) constant applied stress 2 and the corresponding strain; (c) combined stress history of (a) and (b) and the resulting strain obtained from superposition of strain curve 1 and strain curve 2.

Bazant (1982) pointed out that the Stieltjes integral is advantageous over the commonly used Riemann integral because it is applicable to both continuous and discontinuous stress histories. For a continuous stress history, the Riemann integral is preferred, which is given by
t

(t ) = J (t , t ' )
0

(t ' ) dt ' + sh (t ) t'

(2.10)

20

Various creep models are available in the literature, for example ACI model (ACI Committee 209, 1982), CEB-FIP model (CEB-FIP Model Code, 1978) and BP model based on the double-power (Baant and Panula, 1978, 1980, 1982). These models have a common theoretical ground, that is, concrete is taken as material that undergoes aging, which means that the material properties are described as a function of the age, t ' . A different type of creep model based on the solidification theory for aging creep was proposed by Baant and Prasannan (1989a, b). Instead of treating concrete as aging material, the aging aspect of concrete creep is thought to be a consequence of the growth of volume fraction of the load-bearing solidified matter due to hydration of cement. This model was later known as the B3 model (Baant and Baweja, 1995a, b) and presented as a RILEM recommendation. One of the main arguments of this model is to remedy the shortcomings of the aging creep models, which may not necessarily satisfy the thermodynamics restrictions. From the numerical point of view, the model is much simpler to implement than the creep models with aging. Recently, a microprestress-solidification theory was introduced by Baant et al. (1997), which is an improved version of the solidification theory. The new model was formulated based on the physical processes in the microstructures of the hydration products so as to provide a physically justified explanation for long-term aging and drying creep of concrete. A different approach to implement the principle of superposition is by expressing the superposition integral in terms of the relaxation function R(t , t ' ) , which is defined as the stress caused by a unit constant strain imposed at time t ' . The typical relaxation curves of a relaxation function are shown in Figure 2.9. By this way, the total uniaxial stress may be written as an integral of the stress increments produced by the stressdependent strain increments as
t

(t ) = R(t , t ' ) [d (t ' ) d sh (t ' )]


0

(2.11)

The shrinkage strain increment must be isolated from the total strain increment in Eq. 2.11 since the development of shrinkage is considered to be independent of the concrete stress. The creep function and the relaxation function are interchangeable.

21

This may be achieved by solving Eq. 2.9. Baant and Kim (1979) proposed an approximated relaxation function from a given creep function as R(t , t ' ) 1 0 0.115 J (t , t ' ) 1 J (t , t ' ) J (t , t 1) J (t , t '+ ) (2.12)

where the unit of time is day, 0 is approximately 0.008 day and = 0.5(t t ' ) . Reasonable accuracy was shown by comparing the results of Eq. 2.12 with the relaxation curves calculated from a direct solution of Eq. 2.9.

R (t, t)
E(t)

R (t, t)

t = t3 t = t2 t = t1 t = t 0

t0

t1

t2

t3

log (t - t)

(a)

(b)

Fig. 2.9 - Typical relaxation curves for a specimen subjected to an imposed strain at various ages t ' : (a) relaxation curves on the normal time scale; (b) relaxation curves on the logarithmic time scale. (Baant, 1982).

2.3.2

Shrinkage

The porosity of concrete is determined by the porosity of the cement paste, which is made up of air voids, capillary pores and gel pores (Bisschop, 2002). Contraction occurs when the absorbed water in a porous body is removed. This process is known as shrinkage and it occurs throughout the service life of concrete structures. Shrinkage of concrete may be defined as the time-dependent change in volume in an unstressed state at constant temperature. In general, shrinkage may be divided into four different types, namely plastic shrinkage, thermal shrinkage, chemical shrinkage and drying shrinkage (Gilbert, 2002). Plastic shrinkage is a consequence of evaporation of water from the

22

surface of wet concrete while still in the plastic state, and this may result in cracking during the setting process. Thermal shrinkage is the contraction that occurs in setting concrete as a result of subsequent dissipation of heat generated during hydration of cement. Plastic shrinkage is prominent during the setting period and thermal shrinkage is especially important for mass concrete structures, such as dams, where a massive amount of internal heat is generated during the hydration process. The effects of plastic shrinkage and thermal shrinkage are not considered in this study.

2.3.2.1

Chemical Shrinkage

Chemical shrinkage refers to the contraction of concrete that results from chemical reactions within the cement paste. Two major types of chemical shrinkage may be identified in concrete: autogenous shrinkage and carbonation shrinkage. Autogenous shrinkage is defined as the bulk volume reduction caused by chemical reactions during cement hydration, which occurs under isothermal condition in the absence of hygral exchange with the ambient medium. Hydration of cement continues long after setting is completed. This continuation of hydration results in removal of water from the capillary pores and leads to a drying process within the material. This process is known as self-desiccation and is the source of autogenous shrinkage. Unlike drying shrinkage, autogenous shrinkage is independent of the size of the specimen and is treated as an intrinsic characteristic of the material (CEB, 1997). Except at extremely low water-cement ratios, autogenous shrinkage is relatively small, only about 5% of the maximum drying shrinkage (Baant, 1988). Reaction of calcium hydroxide from the cement paste with the atmospheric carbon dioxide is the cause of carbonation shrinkage. However, atmospheric carbon dioxide rarely penetrates through concrete surface for more than a few millimetres, and the effect of carbonation shrinkage is therefore insignificant compared to drying shrinkage. Thus, in some cases, autogenous shrinkage and carbonation shrinkage may well be negligible (Baant, 1988).

23

2.3.2.2

Drying Shrinkage

Drying shrinkage is the most prominent shrinkage and is defined as the time-dependent reduction of volume at constant temperature and relative humidity due to loss of water from concrete stored in unsaturated air. The process of drying shrinkage begins as soon as the absorbed water is lost to the environment. The mechanisms of drying shrinkage is so far not fully understood, however, it is believed that the bulk shrinkage of the cement paste is attributed to phenomena such as capillary stress, disjoining pressure, movement of interlayer water, and changes in surface free energy (Mindess and Young, 1981; Hansen, 1987; Bisschop, 2002). Drying shrinkage is approximately proportional to the loss of water from concrete (Carlson, 1937; Pickett, 1946). Nevertheless, the loss of water with time depends on the size of the specimen. This inevitably makes it difficult to use the data on the loss of water to predict the final shrinkage. Mensi et al. (1988) developed a generalized pattern of loss of water with distance from the drying surfaces of specimens based on the assumption that the rate of diffusion of vapour is proportional to the square root of the time elapsed. It is, however, far more complicated for real structures, as the size and shape are non-uniform throughout. Figure 2.10 shows the relations of the loss of water and the age of concrete for test prisms of various sizes.

10

Water loss after mixing (% total vol.)

700 x 700 x 280 700 x 700 x 560


0x 700 x 70 840

0 0 x 168 00 x 70 7 00 00 x 28 7 00 x 7
NB: Size of specimen in mm

3 Age (years)

Fig. 2.10 - Water loss in specimens of various sizes (LHermite, 1978).

24

The drying shrinkage that occurs in concrete that has been dried in air is not fully recoverable by rewetting, even if the wetting period is longer than the period of drying. For most concretes, the irreversible shrinkage can be as large as 30% to 60% of the ultimate first drying shrinkage (Pickett, 1956; Helmuth and Turk, 1967; LHermite, 1960). A possible reason for the irreversible shrinkage is the development of additional bonds within the gel during the drying process that subsequently reduces the gel pores. The irreversible shrinkage residual may be reduced if the cement paste in concrete is hydrated to a considerable extent before drying (Neville, 1995). Since the main factor causing shrinkage is the evaporable water in cement paste, a high water-cement ratio in concrete results in a high amount of shrinkage. For a concrete with water-cement ratio between 0.2 and 0.6, shrinkage of hydrated cement paste is found to be directly proportional to the water-cement ratio (Brook, 1989; Neville, 1995). The amount of aggregate also has an important influence on shrinkage. Aggregate provides restraining actions to the cement paste that undergoes drying shrinkage. The influence of water-cement ratio and aggregate content on shrinkage is shown in Figure 2.11. In practical application, it is not necessary to distinguish between the components of shrinkage. The concrete shrinkage strain is usually considered to be the sum of the drying, chemical and thermal components (Gilbert, 2002). No thermal effects are considered in this work, and so shrinkage is considered as a composite phenomenon of drying shrinkage and chemical shrinkage.
1600

50 %

Shrinkage (x 10 )

1200 800 400 0 0.3

% 60

Aggregate content by volume (%)

-6

70%

80%

0.4

0.5

0.6

0.7

0.8

Water / cement ratio

Fig. 2.11 - Influence of water/cement ratio and aggregate on shrinkage (dman, 1968).

25

2.3.2.3

Effects of Shrinkage

Shrinkage usually occurs in different amounts at different locations within a concrete element depending on the shape of the structure. Shrinkage tends to be largest on the surface due to rapid moisture loss and lowest in the interior of the concrete furthest from the drying surface. The high shrinkage on the surface is restrained by the lower shrinkage in the interior, which induces a differential shrinkage within the member. This gives rise to the development of tensile stress on the surface and compressive stress at the interior and may eventually lead to surface cracking. In addition, differential shrinkage due to unsymmetrical drying may even causes warping in a concrete member. Concrete structures are usually made up of plain concrete and reinforcing bars. The embedded reinforcing bars restrain the concrete from shrinking freely due to bond action. Consider a singly reinforced section, or an unsymmetrically reinforced section (amount of tension and compression reinforcements are not equal). Different restraints are exerted by the top bars, if any, and the bottom bars as shrinkage develops. A shrinkage induced curvature is produced and this may eventually result in undesirable shrinkage-induced deflection of the member. Moreover, most concrete structures consist of statically indeterminate members and the development of shrinkage provokes redistribution of internal actions that may lead to cracking. Unsightly wide cracks are commonly observed for members in which significant restraint is provided to movement caused by shrinkage. In some cases, cracks are even observed before the application of load.

2.3.3

Interaction of Fracture and Creep

The study of time-dependent fracture of quasi-brittle materials has gained increasing attention in the last decade. In classical fracture mechanics, the mechanical behaviour of materials is assumed to be time-independent. In fact, the bond breakage process at the fracture front is time-dependent, unlike most metals. The viscoelasticity of the creep outside the fracture process zone and the time-dependent effect in the fracture process

26

zone are not negligible. The influence of creep on fracture is recently evidenced in some experimental studies of time-dependent fracture under quasi-static loading conditions (Baant and Gettu, 1992; Zhou, 1992, 1993; Zhou and Hillerborg, 1992; Baant and Xiang, 1997).

2.3.3.1

Influence of Loading Rate on Peak Load

In the three-point bending fracture tests of Baant and Gettu (1992), the peak load is higher for a faster loading rate. Figure 2.12a shows Baant and Gettus test results of two specimens loaded with different crack mouth opening displacement (CMOD) rates. They also tested specimens of three different sizes in order to investigate the influence of loading rate on both peak load and size effect. The results are shown in Figure 2.12b, in which the lines depict the theoretical model of Baand and Jirsek (1993).

(a)

(b)

Fig. 2.12 - (a) Load-CMOD curves for two three-point bend concrete fracture specimens under different rates of loading (after Baant and Gettu, 1992). Dashed lines are the theoretical predictions of Wu and Baant (1993). (b) Influence of loading rate and specimen size on the peak load (after Baant and Gettu, 1992). Dashed lines are the theoretical predictions of Baant and Jirsek, 1993.

27

2.3.3.2

Load Relaxation at Fracture Zone

Another important fracture phenomenon related to the creep effect is load relaxation. Zhou and Hillerborg (1992) performed tension relaxation tests on notched cylinder specimens under displacement control at a constant rate. The displacement was increased right after the peak and held constant for 60 minutes. Load relaxation was observed at the constant displacement, which is depicted by the vertical stress drop in Figure 2.13a. After the first relaxation, the displacement was increased again and two additional relaxations were performed for durations of 30 minutes. The test results are shown in Figure 2.13.

(a)

(b)

Fig. 2.13 - Tensile relaxation tests: (a) stress versus displacement curve; (b) stress versus time. (after Zhou and Hillerborg, 1992; diagrams extracted from Baant and Planas, 1998).

2.3.3.3

Creep Rupture

Zhou and Hillergorg (1992) undertook a series of three-point bending tests on notched beams, each subjected to sustained constant loading, in order to investigate the effects of creep on fracture. It was found that the crack gradually grew with time although no additional load was added. A typical result of the tests and the prediction of their proposed theoretical model are shown in Figure 2.14. The response is characterized
28

firstly by a decreasing CMOD rate and is followed by a rather constant CMOD rate over a period of time. The specimen finally failed by creep rupture accompanied by a large CMOD.

100 80

Experimental Theoretical

CMOD (m)

60 40 20 0

100 200 300 400 500 600

time (s) Fig. 2.14 - Results of creep rupture tests of Zhou and Hillerborg, 1992 (diagrams extracted from Baant and Planas, 1998).

2.3.3.4

Time-dependent Fracture Models

Since the first publication of work on the time-dependent fracture by Zhou (1992) and Baant and Gettu (1992), many researchers have attempted to develop reliable theoretical models to simulate the observed behaviour. In general, three approaches are available in the literature. The first approach is based on the concept of rate-dependent softening. Baant (1993) suggested that the bulk creep of the material and the rate-dependence of bond rupture in the fracture process zone are the factors responsible for time-dependent fracture of concrete. Baant derived the rate of bond rupture in the fracture process zone based on the theory of activation energy (Glasstone et al., 1941) and expressed the & cohesive stress as a function of the crack opening wcr and the crack opening rate wcr at constant temperature (Wu and Baant, 1993) as follows:

29

& & & = F ( wcr , wcr ) = wt ( wcr ) + [k 0 f ct 0 + k1 wt ( wcr )] sinh 1 ( wcr / wr )

(2.13)

where the superimposed dot denotes the derivative of time, wt is a function for the stress versus crack opening curve obtained from an infinitely slow loading, k 0 and k1 & are material constants, f ct 0 is the tensile strength at zero opening rate and wr is an empirical constant which is called the reference opening velocity. de Borst et al. (1993b) proposed a simpler equation for the cohesive stress by the use of a viscosity term m: m & = et ( cr ) 1 + cr f ct

(2.14)

where cr is the cracking strain in the fracture process zone and et is a function for the stress versus cracking strain curve. The second approach is based on the use of rheological models to represent creep and relaxation in the fracture zone. In the model of Zhou and Hillerborg (1992), a modified Maxwell chain element is adopted for modelling creep fracture in the cohesive zone while treating the material outside the cohesive zone as linear elastic. The cohesive stress is expressed in an incremental form as d = d r + d w (2.15)

where d r is the stress increment due to relaxation and d w is the stress increment due to a crack opening increment dwcr and are given by d r = ( 0 )(e dt / 1) d w = 1 dw Fu (e dt / + 1) 2 (2.16)

(2.17)

In Eq. 2.16 and Eq. 2.17, is the stress at time t, 0 is the stress obtained from the static stress-crack opening curve corresponding to crack opening wcr , Fu is the unloading slope and and are material constants.
30

The third approach describes time-dependent fracture by combining a timeindependent micromechanical model for tension softening and a time-dependent rheological model. This approach is aimed at the modelling of very slow or static fracture such as cracking in mass structures like dams and no rate effects are accounted for. One of these models was introduced by Sathikumar et al. (1998) and was later enhanced by Barpi and Valente (2001, 2003) using a fractional order rate law. In the authors view, the effect of time-dependent fracture due to creep is only significant in the case of fracture of plain concrete, especially when the rate of loading is an important factor. Concrete structures are rarely made up entirely of plain concrete, they are strengthened with steel reinforcement which induces significant restraint to cracking. Moreover, creep is accompanied by shrinkage which is a major cause of crack widening with time. The influence of the rate of loading is important in seismic and impact analyses, which beyond the scope of the present study. The effects of creep on cracking in the context of this study are relatively unimportant. Therefore no further consideration is made related to this phenomenon in the subsequent sections.

2.4 Behaviour of Reinforcement


In general, two types of reinforcing steel may be identified based on the difference in the stress-strain response (Kaufmann, 1998). The first type is the hot-rolled, low-carbon or micro-alloyed steel. The tensile stress-strain diagram is shown in Figure 2.15a. Reinforcing steel of this type is characterized by an initial linear elastic stress-strain relationship up to the yield stress f sy and the stress stays constant over a yield plateau. This is followed by a strain hardening range and the steel fails completely when the stress reaches the tensile strength f su . The second type is the cold-worked or highcarbon steel. This steel type, also, exhibits a linear elastic response at initial loading, but has no distinct yield point. The stress-strain relationship shifts from linear elastic to strain hardening through a smooth transition (Figure 2.15b). Due to the absence of an observable yield point, the 0.2% proof stress is usually taken as the yield stress and, the yield strain sy is as shown in Figure 2.15b.

31

s fsu fsy
Es 1

s fsu fsy

Es 1

sy

su

0.2%

sy

su

(a)

(b)

Fig. 2.15 - Stress-strain curves for reinforcing steel: (a) hot-rolled, heat-treated, lowcarbon or micro-alloyed steel; (b) cold-worked or high-carbon steel.

In accordance with the Australian and New Zealand standards AS/NZS 4671 (2001), three classes of reinforcing steel are available, namely low ductility, normal ductility and high ductility for earthquake resistant design, which are denoted by the letters L, N, and E, respectively. The previous 400Y grade steel has recently been replaced by 500-grade steel with a higher minimum guaranteed yield strength of 500 MPa, which can be further designated with the appropriate class of ductility mentioned above. The 500-grade of steel bar often comes with deformed ribbed surface and is used as main reinforcement. Another generally adopted grade of steel in reinforced concrete design is the grade 250N plain (or round, bar surface undeformed) steel bar which has a minimum guaranteed yield strength of 250 MPa. As a result of the low bond associated with plain round bars, the 250N plain bars can only be used as column ties and beam stirrups (AS 3600, 2001).

2.5 Bond between Reinforcement and Concrete


The load-carrying capacity of reinforced concrete structures depends on many factors, including the quality of bond between the reinforcing steel and the concrete. The composite interaction between the two materials is established and maintained by the bond stress, which effectively transfers load between the steel and concrete. The main

32

mechanism in the development of bond stress is the mechanical interaction between the ribbed or deformed surface of the reinforcing bar and the concrete, although other mechanisms such as surface friction and chemical adhesion also play a role.

2.5.1

Local Bond Stress-slip Relationship

Numerous tests have been performed dedicated to the understanding of the bond mechanism between reinforcing steel and concrete. In general, bond is treated as an average effect over the contact surface area of concrete and steel and is commonly described by a local bond stress-slip relationship. The local bond stress-slip relationship is obtained by means of pull-out tests on reinforcing bars embedded in concrete and is characterized by four different stages as shown in Figure 2.16a (fib, 2000). At the initial stage, the bond stress remains as low as 20% to 80% of the tensile strength of concrete. No cracking is observed at this stage. Bond mechanism is realized primarily by chemical adhesion and partly by the micromechanical interaction due to the roughness of steel surface. As the pull-out deformation proceeds, the relative displacement between steel and concrete increases and results in a higher bond stress. The second stage commences when the chemical adhesion is entirely broken down, in which the bond stress is in excess of 1 as shown in Figure 2.16a. Large bearing stresses are exerted by the lugs of the deformed bar and causes formation of transverse microcracks in the concrete adjacent to the tip of the lugs. These microcracks are also known as internal bond cracks (Figure 2.16b) and were first observed experimentally by Goto (1971). At stage three, the bond stress increases beyond the tensile strength of concrete and longitudinal splitting cracks begin to form radially around and parallel to the reinforcing bar as shown in Figure 2.16b. The splitting action is triggered when the radial stresses induced by the wedge action of the ribs exceed the tensile strength of concrete (Tepfer, 1973, 1979). This stage ends as soon as the splitting crack reaches the surface of the concrete member. In the case of low confinement, such as concrete with light transverse reinforcement, the concrete member usually fails abruptly with a through splitting crack.
33

Bond stress, b

Confinement increases

Stage 2

Pull-out failure

1
Stage 1
Abrupt splitting failure

Splitting failure Stage 4

Slip, s

Stage 3

(a)

Internal bond crack

Splitting crack

Pull-out direction

(b) Fig. 2.16 - Pull-out mechanisms: (a) local bond stress-slip relationship and stages of debonding (modified from fib, 2000); (b) internal bond cracks and splitting cracks (Vandewalle, 1992).

At the final stage, the bond stress increases to the peak, beyond which bond stress gradually decreases with a large increase in slip. Depending on the degree of confinement provided by the transverse reinforcement, the concrete member may fail by splitting, pull-out, or both simultaneously. In the case of heavy transverse reinforcement, the bond failure is caused by pull-out of the bar, whereas for concrete member with low transverse reinforcement, bond breaks down by splitting failure with longitudinal cracks penetrating through the concrete cover. In this case, the softening
34

branch of the bond stress versus slip curve is shifted down by a considerable amount. In practice, appropriate design of transverse reinforcement and concrete cover must be enforced in order to prevent the brittle bond failures involved at this stage. The local bond stress-slip relationships have been investigated by many reearchers in order to describe the four stages of pull-out mechanisms discussed above (for example Rehm, 1961; Nilson, 1968; Martin, 1973; Mirza and Houde, 1979; Ciampi et al., 1981; Shima et al., 1987). The CEB-FIP Model Code 1990 (1993) has adopted the relation proposed by Ciampi et al. (1981) and provides specific parameters for confined concrete and unconfined concrete. Upon yielding of steel reinforcement, the bond between steel and concrete is significantly reduced due to the lateral contraction of steel bar resulting from the Poissons effect. Shima et al. (1987) included reinforcing steel strain as a parameter in their proposed relation so as to extend the application to the post-yield range. Huang et al. (1996) expressed the post-yield bond stress-slip relation with a lower softening branch for both normal strength concrete and high strength concrete (Figure 2.17) with the base curve similar to the bond model of the CEB-FIP Model Code 1990 (1993). The development of splitting crack results in a negative effect on bond transfer, similar to yielding of the steel bar. It reduces normal forces around the bar and hence decreases the bond transfer. Gambarova et al. (1982, 1989, 1996, 1997) investigated the effect of splitting cracks on the bond stress-slip relationship by using pull-out specimens with preformed splitting cracks. Bond stress, b
max y Reinforcing steel in plastic range Reinforcing steel in elastic range

f y.f sy s1 s3 sy.f s3 s5

s4

Slip, s

Fig. 2.17 - Local bond stress-slip relationship of Huang et al. (1996).

35

The present study focuses on the formation of transverse cracks in concrete structures at service loads, in which the bond stress levels are normally low. Hence, the effect of splitting cracks on bond will not be considered here.

2.5.2

Influence of Bond on Cracking

The quality of bond has a marked influence on crack formation in reinforced concrete structures, both in terms of the spacing between cracks and the crack width. Consider a reinforced concrete member subjected to an axial tensile force, as shown in Figure 2.18a. The first primary crack is formed when the concrete stress, transferred from the steel reinforcement via bond action, reaches the tensile strength of the concrete. This first crack occurs at that section along the member where the tensile strength is smallest. At the crack, the concrete stress is zero, the steel carries the entire tensile force, slip takes place between the concrete and the steel and the bond stress is negligible. Adjacent to the crack, the bond stress b increases rapidly, the tensile stress in the concrete c increases and the tensile stress in the steel s decreases. At some distance so from the first crack, the bond stress reduces to zero and the concrete and steel stresses are unaffected by the crack. If further tensile force is added, a second primary crack will form when the concrete tensile strength is exceeded on the next weakest section greater than so from the first crack. The process continues until the final crack pattern is established. No further cracking can occur if the distance between the cracks is not large enough to develop sufficient bond to allow the concrete stress on any section between the cracks to reach the concrete tensile strength. The final crack spacing is therefore somewhere between so and 2so. Figure 2.18b shows a magnified view of the concrete-steel debonding at the primary crack.

36

Primary crack

bond stress b steel stress s concrete stress c

Internal crack

(a) stress distributions; (b) stress transfer by bond.

(b)

Fig. 2.18 - Cracking in a reinforced concrete element: (a) direct tension member and

2.5.3

Tension Stiffening

One of the most important consequences of the effect of bond in cracked reinforced concrete is tension stiffening. The conventional approach completely disregards the effectiveness of the tensile concrete in a cracked reinforced concrete member and the total applied tension is thought to be resisted only by the steel reinforcement. This apparently underestimates the total stiffness of the structure and leads to an unrealistically higher prediction of deformation. As demonstrated in the previous section, concrete between the cracks is capable of carrying tensile stress due to stress transfer via the bond between concrete and steel reinforcement. The structural response of a cracked reinforced concrete tension member is stiffer than that of the naked reinforcing steel and this stiffening phenomenon is called tension stiffening. The tension stiffening effect may be further illustrated by considering the load versus average strain response of a reinforced concrete tension chord between two cracks (see Figure 2.19a). In order to compare with the stiffness of the bare reinforcing steel, the load on the tension chord is expressed as the force per unit area of steel reinforcement. The comparison is shown in Figure 2.19c. It is evident that the contribution of concrete between the cracks increases the stiffness. This can be observed by comparing the secant modulus of the tension chord E sm to the elastic

37

modulus of the bare reinforcing steel E s at the same stress level 1 shown in Figure 2.19c. The saw-tooth response in the stress-strain curve represents the formation of primary cracks in the tension chord.

= N /As
Embedded steel

(a)

cr
Esm

1 Bare steel Es

sm s

(b)

(c)

Fig. 2.19 - Concrete tension stiffening: (a) tension chord with embedded reinforcing steel; (b) bare reinforcing steel; (c) comparison of stiffness of embedded steel and bare steel.

2.6 Non-linear Modelling of Concrete Structures


Reinforced concrete is well known for its non-linear behaviour. The non-linearity in reinforced concrete originates from the non-linear stress-strain relationship of plain concrete. The structural behaviour is further complicated by cracking of concrete which causes a considerable redistribution of stresses within the intact concrete as well as the stress transfer from concrete to steel reinforcement. The finite element method is the most often adopted numerical tool to simulate the non-linear behaviour of reinforced concrete structures. Rational and reliable representations of cracking of concrete must be used in conjunction with the finite element method in order to accurately describe the structural behaviour. Two major approaches are available in the literature for the modelling of cracking in concrete

38

structures, namely the discrete crack approach and the smeared crack approach. These two approaches are discussed in the subsequent sections.

2.6.1

Discrete Crack Approach

In the discrete crack model, cracking is simulated as a propagation of discontinuities in a structure. Two approaches may be identified in the discrete crack model, the interelement crack model and the intra-element crack model. In the first approach, crack propagation is modelled by means of separation of element edges and the cohesive tractions in the fracture zone are simulated using either linkage elements or interface elements (as depicted in Figure 2.20a). This approach was employed widely in the early development of finite element models for reinforced concrete structures (Ngo and Scordelis, 1967; Nilson, 1969; Mufti et al., 1972; Al-Mahaidi, 1979). However, these discrete crack models were later replaced by the smeared crack approach which is more attractive in a variety of computational aspects. The early discrete crack approaches suffer from two major drawbacks. The first drawback is the limitation of the crack trajectory, which is constrained to follow the path along the predefined inter-element boundaries. The second drawback is the increasing computation cost and the decreasing efficiency due to the additional degreeof-freedom of the separated nodes along the new crack faces. To remedy the first drawback, a remeshing technique must be adopted. In this technique, the finite element mesh topology is modified at each crack increment according to the direction of the crack propagation. This technique was pioneered by Saouma and Ingraffea (1981) and Saouma (1981) for linear elastic fracture mechanics problems and was extended to nonlinear mixed mode fracture mechanics of concrete (Ingraffea et al, 1984; Bocca et al., 1991; and Gerstle and Xie, 1992; ervenka, 1994; Valente, 1995). The second approach is the intra-element discrete crack model in which the crack is allowed to propagate through the finite element as shown in Figure 2.20b. Two types of model are available in this approach, namely the embedded discontinuity model and the model based on the partition-of-unity concept. The embedded discontinuity model originally emerged as a tool to deal with strain localization problems such as shear band

39

in metals and has been further extended to cohesive material such as concrete (Ortiz et al., 1987; Belytschko et al., 1988; Klisinski et al., 1991, Lofti and Shing, 1994; Simo and Oliver, 1994). The discontinuity at cracks is regarded as a displacement jump in the element by incorporating additional localization modes to the standard shape function of the finite element (Sluys and Berends, 1998). The second type of intra-element crack model is formulated on the basis of the partition-of-unity concept (Duarte and Oden, 1996; Melenk and Babuka, 1996). In this model, discontinuous shape functions are used and the displacement jump across the crack is represented by extra degrees of freedom at the existing nodes (Wells and Sluys, 2001). The discrete crack approach may sound physically appealing since the crack is represented by a traction free discontinuity and it does not suffer from the spurious strain localization problems upon mesh refinement which to a degree undermines the smeared crack approach. The discrete crack approach is most useful in structural problems with prominent localized cracking. Nevertheless, concrete structures are often dominated by diffused cracking, which is found especially in reinforced concrete members with closely spaced reinforcement such as shear walls and the tension side of a reinforced concrete beam. In such cases, the smeared crack approach is often considered to be advantageous compared to the discrete crack approach.

(a)

(b)

(c)

Fig. 2.20 - Concrete cracking models: (a) discrete inter-element crack approach; (b) discrete intra-element crack approach; (c) smeared crack approach (crosses in the shaded element are the cracked integration points).

40

2.6.2

Smeared Crack Approach

The smeared crack approach is a concept developed based upon the framework of continuum mechanics which is described in the notions of stress and strain. This approach was introduced by Rashid (1968) to overcome the drawbacks of the early discrete crack models and has become the most widely adopted method for modelling cracking in structural engineering problems. The cracking in a smeared crack model is assumed to be smeared over a certain volume of material. Cracking is treated as a reduction of average material stiffness in the direction of the major principal stresses. In a finite element model, this is done at the integration point level of the element by reducing the stiffness according to a particular constitutive relationship (see Figure 2.20c). One of the most important advantages of the smeared crack approach is that the mesh topology of a structure remains unchanged during the nucleation of cracks. This offers great convenience in numerical implementation. As discussed earlier, cracking in reinforced concrete members is usually distributed due to the stabilising effect of reinforcement. Therefore, the distributed nature of cracking in the smeared crack approach best describes the cracking phenomenon in most concrete structures. In the case of concrete fracture, the smeared crack approach is justified due to the fact that the fracture process zone of concrete is densely filled with microcracks within a finite width of fracture before it finally localizes into a line of discontinuity (Baant, 1985). Notwithstanding the advantages in modelling reinforced concrete, the classical smeared crack approach suffers from a major deficiency when dealing with localized cracking. In fracture problems, the smeared crack approach tends to localize cracking into a single row of elements and produces unobjective tensile post peak results upon refinement of element size. This phenomenon is known as spurious mesh sensitivity and will be discussed in detail later. On the other hand, the classical smeared crack approach also suffers from mesh directional bias which tends to predict crack propagation in alignment with the direction of the mesh. Nevertheless, these deficiencies may be eliminated by utilising one of a number of regularization models which will be discussed later.

41

Three major variants of smeared crack approach are available in the literature depending on the method of development of the crack planes: fixed crack model, rotating crack model and multi-directional fixed crack model.

2.6.2.1

Fixed Crack Model

The first smeared crack model introduced by Rashid (1968) was based on the fixed crack concept. In a fixed crack model, the crack initiates normal to the major principal stress and has a fixed direction throughout the loading the process. Consider a twodimensional linear elastic material, the stress may be written as x E y = 2 1 xy 1 1 0 0 x y 1 (1 ) xy 2 0 0

(2.18)

The material stiffness matrix becomes orthotropic as soon as the maximum principal stress exceeds the concrete tensile strength. The orthotropic stress-strain relationship may be written in the normal (n) and tangential (t) directions of the crack, which gives nt = D sec nt nt (2.19)

where Dsec is the secant stiffness matrix in the directions of orthotropy. In the early nt versions of the fixed crack model (Rashid, 1968; ervenka, 1970), the stiffness normal to the crack and the shear stiffness are both set to zero and the Poissons effect vanishes immediately after cracking. 0 0 0 E 0 0 0 0

D sec = 0 nt

(2.20)

42

The sudden drop of stiffness, however, gives rise to numerical difficulties. To improve this situation, a shear retention factor was introduced (for instance, Suidan and Schnobrich, 1973) for a smooth transition from uncracked to completely cracked states. It may also be conceived as a way to account for the aggregate interlock. The shear component is retained in the stiffness matrix as 0 0 0 0 G

D sec = 0 nt

E 0 0

(2.21)

in which the shear retention factor is in the range 0 1 . To include the strain softening phenomenon of concrete, a refined model was proposed by Baant and Oh (1983). They introduced a normal reduction factor and incorporated the postcracking Poissons effect, which leads to the following expression E 2 1 E D sec = nt 1 2 0

E
1 E 1 2 0
2

0 0 G

(2.22)

In real structures, the direction of principal stresses is constantly changing during the loading process. Because of the fixed crack direction, the rotation of the principal stresses causes the development of shear stress on the crack surface. A large shear stress may be induced if the rotation of the principal stresses is significant. Consequently, the fixed crack model often predicts a stiffer response than experimental results.

2.6.2.2

Rotating Crack Model

The rotating crack model was proposed by Cope et al. (1980), which was motivated by the experimentally observed phenomenon of the rotating principal stress directions in concrete structures. The underlying smeared cracking formulation of the rotating crack

43

model is essentially no different to that of the fixed crack model. A crack is formed normal to the major principal direction upon violation of the tensile strength criterion. However, in the rotating crack model the crack direction rotates with the principal stress directions during the entire loading process. Consequently, the orthotropic constitutive relationship of the cracked concrete changes accordingly with the crack directions. Eq. 2.19 may now be written in the principal 1-2 axes instead of the local crack n-t axes as
sec 12 = D12 12

(2.23)

Although questions were raised for the physical justification of the rotating crack model (Baant, 1983) due to the fact that cracks in real structures do not rotate but the principal stress directions, the rotating crack model certainly is an efficient tool for accurate concrete modelling. In addition, the shear retention factor, which is a sensitive factor in the fixed crack model, is no longer needed in the rotating crack model. This simplification greatly facilitates the numerical implementation and is practically appealing from an engineering point of view.

2.6.2.3

Multiple Fixed Crack Model

The multi-directional fixed crack model is a refined version of the fixed crack model, which aims at improving the simulation of the change in direction of the principal stress. The basic concept is identical to the fixed crack model in which rotation of crack is not permitted. However, new cracks are allowed to develop at different orientations if the angle between two consecutively formed cracks exceeds a threshold angle. A systematic incremental strain-decomposition approach was proposed by de Borst and co-workers (de Borst and Nauta, 1985; de Borst, 1987; Rots, 1988). The total concrete strain increment is decomposed into a concrete elastic strain increment ce and a cracking strain increment cr as follows = ce + cr (2.24)

44

where the cracking strain increment cr is made up of the cracking strains of all cracks at different orientations, which can be written as cr = cr1 + cr 2 + cr 3 + ... (2.25)

de Borst and co-worker derived a constitutive relation for cracked concrete, in which the stress increment is related to the strain increment by a reduced elastic stiffness matrix similar to the elasto-plastic stiffness matrix of plasticity. Since crack opening and crack closing may happen simultaneously, the computational procedure for a multiple fixed crack model may get quite complicated and therefore special computational strategies must be devised to follow the possible bifurcation path (Baant and Planas, 1998).

2.6.3

Constitutive Models for Concrete

One very important aspect in smeared crack modelling of concrete is the development of constitutive models that are capable of describing the behaviour of concrete. In the broad sense, the constitutive models of concrete may be classified into two categories: the macroscopic phenomenological models and the micromechanical models (Baant and Prat, 1988). The macroscopic models may be further divided into two subgroups according to the treatment of the stress-strain evolution law. The first group is the models based on a total formulation, in which the total stress is related to the total strain by a secant constitutive relation. Examples of total formulation models are the elasticity-based cauchy elastic models and the continuous damage models. The second group refers to the models adopting an incremental stress-strain concept via the use of tangent moduli, which is referred to as the incremental formulation. Plasticity-based models, elasticity-based hypoelastic models, plastic-fracturing models and endochronic models are the examples of this subgroup of macroscopic models. Whilst the macroscopic models describe the material behaviour via a direct stressstrain constitutive relationship, the micromechanical model is characterized by a constitutive relationship based upon the mechanical interaction of the cement matrix

45

and aggregates loaded on the microscopic and macroscopic levels. A well-known example of the micromechanical model is the microplane model. In the following sections, a brief review and discussion is given of the frequently adopted constitutive models in concrete modelling, namely the elasticity-based models, the plasticity-based models, the continuous damage models and the microplane model.

2.6.3.1

Elasticity-based Models

In general, the elasticity-based models can be formulated based on a total formulation or an incremental formulation. In the elasticity-based models based on the total formulation, or called the cauchy elastic models (ASCE Task Committee, 1982, Chap. 2), the total stress is a function of the total strain and is given by

= f ( )
It may be written in a secant constitutive relationship in tensorial notation as
sec ij = Dijkl kl

(2.26)

(2.27)

sec where ij and kl denote the stress and strain tensors, respectively, and Dijkl is a

fourth-order secant stiffness tensor, which is a function of the stress state. It is evident that the total stress is uniquely related to a single total strain via the material response function as shown in Eq. 2.26. Due to the deformation pathindependence introduced by the total stress-strain constitutive relationship, these models are not able to provide a physically sound base for representing the behaviour of concrete. In the case of loading and unloading stress histories, the total formulation elasticity-based model requires a different stress-strain relationship in order to handle the unloading behaviour of concrete. This further gives rise to difficulties in generalizing the total formulation models to represent behaviour of concrete subjected to cyclic loading.

46

The elasticity-based models based on the total formulation concept have been extensively used in simulating the behaviour of reinforced concrete structures under inservice loading in which compressive loads are relative small and remain elastic. Concrete prior to cracking can be modelled sufficiently accurate as an isotropic linear elastic material (for example Ngo and Scordelis, 1967; Rashid, 1968; Baant and Gambarova, 1980; Baant and Oh, 1983). The majority of these models concentrate on the behaviour of cracking in concrete, including strain-softening and localization in particular. A secant or an initial stiffness is often utilized to model unloading in concrete after cracking. The cauchy elastic model has also been extended to represent non-linearity of concrete. One of the earliest biaxial models based on the cauchy elastic formulation was proposed by Kupfer and Gerstle (1973). They formulated an isotropic total stressstrain model based on biaxial monotonic tests of concrete and established a series of expressions for the secant shear and bulk moduli. However, good results were only obtained for comparisons with experimental data at low stress levels. Despite the criticisms related to the physical interpretation, the total formulation has been widely adopted in recent years to simulate the non-linear behaviour of concrete. This is due primarily to the simplicity of implementation of the models, especially for extending an isotropic linear elastic model for non-linear modelling. The models have been shown to work well in concrete modelling in many studies (Stevens et al., 1991; Ramaswamy et al.; 1995, Vecchio, 1989, 2001; Foster and Marti, 2003). One of the important models in modelling concrete membrane structures is the modified compression field theory (Vecchio and Collins, 1986), which is formulated based on the total formulation concept. A concrete compression softening model was proposed along with the theory, which states that the peak compressive stress for cracked concrete under compression-tension state of stress decreases with increasing principal tensile strain. Vecchio and Collins (1989) introduced a strength reduction factor , which is defined as the ratio of the reduced peak compressive stress and the concrete compressive strength, and is given by

47

0.85 0.27 1 2

1.0

(2.28)

where 1 and 2 are the major principal and minor principal strains, respectively. Vecchio (1989) implemented the modified compression field theory into a secant stiffness non-linear finite element program and demonstrated good agreement with experimental results for reinforced concrete membranes. The second type of elasticity-based models is formulated based on an incremental stress-strain constitutive relationship, which is also called the hypoelastic models (Truesdel, 1955). Stress and strain are related by the material tangent stiffness that is dependent on the current state of stress and may be expressed in tensorial notation as
tan d ij = Dijkl d kl

(2.29)

where d ij and d kl are the incremental stress and strain tensors respectively and
tan Dijkl is a fourth-order tangent stiffness tensor. From a thermodynamics viewpoint,

these models provide a more physically convincing base than the models based on a total formulation due to the path-dependent nature of the deformation history. A typical biaxial model based on the isotropy concept was proposed by Gerstle (1981). In his model, the response of concrete is divisible into hydrostatic and deviatoric components and the constitutive relationship is expressed in terms of tangential bulk and shear moduli. Good correlation was observed with experimental data. A different approach was developed based on the experimentally observed stressinduced anisotropy in concrete. As a special case of anisotropy, concrete subjected to a biaxial state of stress is treated as an orthotropic material and the constitutive relation is constructed along the axes of orthotropy in the principal stress directions. An important model of this type was presented by Darwin and Pecknold (1977). In their incremental orthotropic model, the concept of equivalent uniaxial strain was proposed to subtract the Poissons effects in the principal directions so as to represent the degradation in concrete with an equivalent uniaxial stress-strain relationship. The total equivalent

48

uniaxial strain iu is calculated through the accumulation of each incremental change in strain caused by the incremental change in principal stress and is given by
N inc

iu =

i Ei i =1

(2.30)

where N inc is the total number of load increments. The incremental constitutive relations are written in the principal directions of orthotropy as d12 = D12 d12 in which D12 is the tangent stiffness matrix given by E1 E1 1 E 2 E 2 D12 = 1 2 0 0 1 ( E1 + E 2 2 E1 E 2 4 0 0 (2.31)

(2.32)

where is an equivalent Poissons ratio equal to 1 2 . E1 and E 2 are the tangent moduli obtained from uniaxial stress-strain curves. The model was demonstrated to be capable of modelling concrete under monotonic and cyclic loading.

2.6.3.2

Plasticity-based Models

To capture the more general behaviour of concrete over the full loading range, the plasticity-based models are advantageous over the elasticity-based models. The parameters required in the plasticity-based models are greatly reduced compared to the elasticity-based models. The plasticity theory was originally developed to represent the behaviour of ductile materials such as metals. Extending the theory to represent the behaviour of concrete, for which the non-linear behaviour is characterized by dense microcracking in the material, requires significant modifications to the classical plasticity models.

49

The establishment of a standard plasticity model involves three essential conditions, a yield surface, a hardening rule and a flow rule. Plastic deformation is deemed to initiate when the stress of a material reaches the yield surface in stress space. The evolution of loading surfaces after yield is subsequently governed by the hardening rule. During plastic deformations, the plastic strain evolution rate is controlled by the flow rule using a plastic potential function (Chen and Han, 1988; Chen et al., 1992). In the plasticity theory, inelastic deformation of materials is measured by the amount of plastic strain developed during the course of loading. Expressing in a differential form, the total strain rate is divided into the elastic and plastic components & & & = e + p (2.33)

& & where the superimposed dot denotes the first derivative of time, e and p are the elastic and plastic strain rate vectors, respectively. The stress rate is related to the elastic strain rate by a symmetrical linear elastic constitutive matrix, given by & & & = D e ( p ) (2.34)

where the bracket term is the elastic strain rate vector obtained from Eq. 2.33. For the case of isotropic hardening or softening plasticity, the yield surface is described as a function in stress space given by f (, ) = 0 (2.35)

where is a scalar value called the hardening or softening parameter that is typically dependent on the strain history. A stress point is not permissible to lie outside the yield surface and must remain on the yield surface during the course of plastic flow. Hence, a condition is introduced, the Pragers consistency condition, which is given by the equation: & f (, ) = 0 (2.36)

For a given flow rule, the plastic strain rate vector is expressed as the product of a & positive proportionality scalar and a vector m given by

50

& & p = m

(2.37)

& in which the scalar and the vector m indicate the magnitude and the direction of the plastic flow, respectively. The vector m is defined as the gradient of the plastic potential function g p of the flow rule, expressed as gp

m=

(2.38)

By defining a vector n as the gradient of the yield function n= f (2.39)

the consistency equation of Eq. 2.36 can be written as & nT + f & =0 (2.40)

where the superscript T denotes the transpose of the vector n. By defining a hardening or softening modulus h as follows h= 1 f & & (2.41)

& & and replacing and ( f / ) in Eq. 2.40 using the relations given by Eq. 2.34, Eq. & 2.37 and Eq. 2.41 and yield the expression for the proportionality rate constant & n T De h + n T Dem

& =

(2.42)

Given a yield function and a plastic potential function, the growth of plastic strain can be determined. Now the stress rate may be related to total strain rate as & & = D ep (2.43)

51

where D ep is the elasto-plastic stiffness matrix, which can be perceived as a reduced elastic stiffness matrix given by Dem n T De h + n T Dem

D ep = D e

(2.44)

The elasto-plastic stiffness matrix presented above is non-symmetrical and was derived based on a non-associated flow rule. For an associated flow rule, the yield function coincides with the plastic potential function ( f = g p ) and hence gives a symmetrical elasto-plastic stiffness matrix. Numerous efforts have been dedicated to develop the yield functions, hardening rules and flow rules in order to accommodate the use of plasticity theory in modelling of concrete (Chen and Chen, 1975; Buyukozturk, 1977; Murray et al. 1979). Chen and Chens (1975) work is one of the earliest attempts to establish a plasticity-based model for concrete. They developed an isotropic hardening plasticity model adopting two similar functions to define the surface in pure compression and tension-compression stress states. However, the model has been subjected to criticisms on its applicability to concrete, since the model postulates concrete to be linear elastic even at high stress levels. Han and Chen (1985) introduced a non-uniform hardening plasticity model assuming the associated flow rule. In their model, a failure surface is defined as a bounding surface to enclose all the loading surfaces. The failure surface remains unchanged throughout the loading process. During hardening, the loading surface expands from the shape of the initial yield surface and reaches the final shape overlapping the failure surface. The model was further improved based on the nonassociated flow rule and demonstrated a good correlation with experimental data. In a more recent development, Feenstra and de Borst (1996) presented an energybased composite plasticity model to describe the behaviour of plain and reinforced concrete in biaxial stress under monotonic loading conditions. Based on the framework of incremental plasticity, they employed a Rankine yield criterion and a Drucker-Prager yield criterion to monitor the respective stresses in tension and compression. The
52

plasticity model is amalgamated with the energy approach based on the crack band theory (Baant and Oh, 1983) for loading in both tension and compression. The model was developed particularly to model tension-compression biaxial stress states in concrete structures, such as shear wall panels.

2.6.3.3

Continuous Damage Models

Continuum damage mechanics was originally proposed as a simple model to study creep failure of metal alloys. It provides a means for describing the degradation of elasticity in material that is caused by microstructural damage (Kachanov, 1958; Rabotnov, 1969; Lemaitre and Chaboche, 1990). In the 1980s, continuum damage mechanics began to attract increasing attention from researchers in the concrete field and the pioneering application to quasi-brittle material was made by Mazars (1984) and Mazars and Pijaudier-Cabot (1989). A set of damage variables were introduced to represent the local loss of material integrity. The formation of a crack is indicated by the severity of damage in that part of the material domain and propagates in the direction of the growth of damage in the fracture process zone. In damage mechanics, the plastic response of concrete is ignored and strains are assumed to be fully recoverable upon removal of stresses. Based on the concept of total formulation, the basic structure of the constitutive relationship (de Borst, 2002) is given by
sec ij = Dijkl (q m , mn , mnop , ...) kl

(2.45)

sec where ij is the stress tensor, kl is the strain tensor and Dijkl is a fourth-order secant

stiffness tensor which may be dependent on several internal variables such as scalar valued variable q m and tensor-valued variables mn and mnop . The constitutive relation in Eq. 2.45 is, at first glance, similar to that in Eq. 2.27 for the elasticity-based models based on a total formulation. In fact, there exists a major difference between the two models, that is, the internal variables engaged in the continuous damage model are

53

history-dependent while the elasticity-based models based on a total formulation are load path-independent as discussed earlier. To illustrate the theory of damage mechanics, an isotropic scalar damage model is considered. The total stress-strain relationship is written as = (1 )D e (2.46)

where D e is the elastic stiffness matrix and the scalar is a damage variable. The damage level of a material is measured via the damage variable for which the value varies from zero to one, indicating the respective state of a fully intact material to a complete loss of material integrity. As in plasticity, a damage model is accompanied by a loading and unloading function f given by ~ ~ f ( , ) = (2.47)

~ where is a history-dependent parameter and is the equivalent strain. The damage loading function f and the rate of the history parameter must satisfy the Kuhn-Tucker conditions at all instances, which are given by f 0, & 0 and & f = 0 (2.48)

In a multiaxial generalization, damage growth is deemed to occur when the damage & loading function f and the its first derivative of time f are equal to zero, that is f =0 and & f =0 (2.49)

The damage variable is obtained through a damage evolution law expressed as a function of the history parameter such that

= F ( )

(2.50)

For modelling the propagation of a crack in concrete, Mazars (1984) expressed the equivalent strain as a function of the principal strains as follows

54

~ =

i =1

( i ) 2

(2.51)

where i is the principal strains, and i = i for i > 0 and i = 0 for i 0 . de Vree et al. (1995) proposed a different function for the equivalent strain and called it the modified von Mises definition. This function is given by ~ = k 1 1 I1 + 2k (1 ) 2k (k 1) 2 (1 2 ) 6k (1 + ) 2

I2 + 2 1

J2

(2.52)

where is the Poissons ratio, I1 and J 2 are the first invariant of the strain tensor and the second invariant of the deviatoric strain tensor, respectively. The parameter k is the ratio of the uniaxial compressive strength to the uniaxial tensile strength ( f c' / f ct ). The isotropic damage models are sufficient for describing progressive crack propagation in concrete, however, a more complicated damage model is generally required for the modelling of reinforced concrete members when damage induced anisotropy is not negligible. For two-dimensional planar problems, concrete is treated as an orthotropic material and two damage variables, 1 and 2 , are introduced to represent the loss of material integrity in the cracking opening direction and the shear direction, respectively (de Borst et al., 1998). Ragueneau et al. (2000) extended the damage-based model to characterize a residual hysteretic response for seismic analysis of concrete structures. The main feature of the model is the coupling of damage mechanics and a sliding effect which produces the hysteretic behaviour of concrete under cyclic loading.

2.6.3.4

Microplane Models

The constitutive models that have been discussed so far are the macroscopic models. In contrast to the macroscopic models are the micromechanical models, in which the constitutive laws are established based upon the microscopic behaviour of the cement

55

and aggregate under loading. The microplane model is a constitutive model of the microscopic type. The model was developed based on the idea of Taylor (1938), which was developed in detail for plasticity polycrystalline metals by Batdorf and Budianski (1949) under the name slip theory of plasticity. The concept was extended by Baant and Gambarova (1984). They replaced the slip planes by damaged planes and named it the microplane model. The macroscopic stress and strain describe the constitutive relationship of a macroscopic model. In the microplane model, the constitutive relationship is defined by the relationship of the stress and strain acting on the planes at various arbitrary orientations in the microstructure of the material, which is called the microplane (see Figure 2.21a). These planes can be imagined as damage planes in the microstructure of the material. The macroscopic behaviour of the material is essentially a composite effect imposed by each microplane oriented in various directions. In the microplane model, the strain on a microplane is obtained as a projection of the macroscopic strain under a so-called kinematic constraint. A set of unit orthonormal base vectors l, m and n with components li , mi and ni , respectively, are introduced on the micro level, in which li and mi lying on the microplane and ni normal to the microplane (see Figure 2.21a). By applying the kinematic constraint, the macroscopic strain tensor ij is resolved into a normal component N and two shear components M and L (Baant et al., 2000) and are given by

N = N ij ij ,
where N ij = ni n j

M = M ij ij

and

L = Lij ij

(2.53)

(2.54a) (2.54b) (2.54c)

M ij = 0.5(mi n j + m j ni ) Lij = 0.5(li n j + l j ni )

56

The lower case subscripts refer to the components of the tensors in Cartesian coordinates xi , where i = 1, 2, 3. Repetition of lower case subscript, referring to Cartesian coordinates xi , implies summation over i = 1, 2, 3. Knowing the strain components on the microplane, the microplane stresses ( N ,

M and L ) may be determined for a given constitutive relationship on the


microplane level. The macroscopic stress is related to the microplane stresses by applying the principle of virtual work and integrating over the surface of a unit hemisphere. In a simplified form, Bazant (1984) expressed the macroscopic stress tensor as

ij =

3 2

sij d

(2.55)

where sij = N N ij + L Lij + M M ij . For finite element implementation, Baant and Oh (1986) proposed an efficient formula to evaluate the integral by using a system of 21 microplanes per hemisphere (as shown in Figure 2.21b) for each integration point of the finite elements. A fourth generation of the microplane model, labelled M4, was presented by Baant et al. (2000). From the first generation of the microplane model M1 focussing on the modelling of tensile fracture to the latest version M4, the microplane model was extended to successfully handle various load types and has shown good agreements with the basic experimental data for uniaxial, biaxial and triaxial loadings.

57

(a)

(b)

Fig. 2.21 - Microplane model: (a) strain components on a microplane; (b) 21-point optimal Gaussian integration for each hemisphere (circled points represent the directions of the microplane normals; a total of 42 circled points on the whole sphere). (after Baant et al., 2000).

2.6.4

Fracture Models for Concrete

2.6.4.1

Fracture Mechanics

There are two major approaches in fracture mechanics, namely the linear elastic fracture mechanics (LEFM) and the non-linear fracture mechanics (NLFM). Griffith (1921) was the first to develop LEFM and to adopt the energy conservation principle to describe the formation of a crack. The theory is based on linear-elasticity and states that the entire body of a material is elastic except in the vanishingly small region at the crack tip. In reality, an inelastic region does exist in the neighbourhood of a crack tip for all types of material, in order to apply the theory the inelastic region must be relatively small in comparison to the size of the body. LEFM applies perfectly to brittle materials such as glass and brittle ceramic as they possess a very concentrated fracture process zone at the crack tip. However, for quasi-brittle materials like concrete, rock and ceramic, the crack tips are surrounded by a relatively large fracture process zone, which is the main contributing factor to the tension softening phenomenon. LEFM is therefore strictly applicable to massive

58

structures such as concrete dams. The comparisons of the characteristics of the fracture process zones for different types of material are shown in Figure 2.22. Although early attempts had been made to model the non-linear behaviour of fracture process zone by researchers such as Irwin (1960), Dugdale (1960) and Barenblatt (1962), fracture mechanics for quasi-brittle material had not been extensively developed until the 1980s. Amongst the NLFM theories, the most significant and most widely adopted models are the fictitious crack model (Hillerborg et al., 1976) and the crack band model (Baant and Oh, 1983). The former is an extension based on the concept of the cohesive crack that was introduced by Dugdale (1960) and Barenblatt (1962) to model various non-linearities at the crack front. The fictitious crack model is therefore also called the cohesive crack model or DugdaleBarenblatt model. The crack band model for strain softening is based on a similar concept to the fictitious crack model, in which the fracture process zone is modelled by a crack band with a finite width instead of a cohesive crack.

(a)
Linear-elastic zone

(b)
Softening zone

(c)
Non-linear hardening zone

Fig. 2.22 - Different types of fracture process zone: (a) brittle materials; (b) ductile materials; (c) quasi-brittle materials. (Baant and Planas, 1998).

2.6.4.2

Fictitious Crack Model

In the fictitious crack model, the response of a concrete member in tension is the same as the process described in Section 2.2.2. The simplification made by Hillerborg was that, after the peak load, microcracking in the fracture process zone is lumped into a line crack (fictitious crack) with a finite opening that is able to transfer stress while the

59

remaining of the concrete member unloads (also see Figure 2.2 in Section 2.2.2). Hence, the post-peak total elongation of a concrete specimen in an idealized direct tension test may be calculated as l = unld L + wcr (2.56)

where unld is the strain in the unloading intact material, L is the total length of the specimen and wcr is the crack opening displacement. Based on the simplification of the model, Hillerborg assumed that non-linear fracture in concrete is described in full by a single stress-crack opening displacement relationship which may be written as = f ( wcr ) . The fracture energy G f and the tensile strength f ct are the two key parameters defining the constitutive relationship (see Figure 2.2c). The fracture energy is defined as the energy consumed in a unit area of concrete fracture, it is equivalent to the area under the stress versus crack opening curve. This may be expressed as
wu 0

Gf =

dwcr

(2.57)

where is the cohesive stress in the fictitious crack and wu is the crack opening at which the stress transfer at the fictitious crack vanishes. Thus, given that the fracture energy and tensile strength are known and with an assumed stress-crack opening displacement softening function, the deformation of a concrete member in tension may be predicted using Eq. 2.56.

2.6.4.3

Crack Band Model

When the fictitious crack model was first introduced, Baant (1976) also presented a non-linear fracture model for concrete, which forms the basis of the well-known crack band model that is broadly used in modelling concrete structures in present engineering

60

design. The concept was then refined and introduced in full by Baant and Oh (1983) to describe the strain softening phenomenon of concrete. The crack band model is not dissimilar to the fictitious crack model. The fictitious crack model involves a distinct line crack, while the crack band model assumes microcracks to be distributed densely over a certain width over the fracture zone. Thus, Baant introduced a characteristic length in addition to the two parameters G f and f ct discussed previously for the fictitious crack model. In the context of the crack band model, the characteristic length is the crack band width hc . Instead of using a stresscracking opening displacement relation, the crack band model describes the fracture behaviour of concrete with a tensile stress-strain relationship in which the strain softening tail of the curve must be modified for different widths of the crack band in order to guarantee a constant fracture energy dissipation rate in the fracture zone.

2.6.5

Regularization of Spurious Strain Localization

The earliest continuum modelling of cracking in concrete was based on the strength criterion. In these models, the concrete tensile response is characterized by a linear elastic pre-peak stress-strain relation and is followed by a sudden drop in stress to zero upon initiation of cracking. However, these models suffer from the severe deficiency of mesh size dependence. Numerical results for the same structure vary notably for finite element discretizations of different mesh sizes especially in the case of localized cracking with little or no reinforcement. Baant and Cedolin (1979) pointed out that objective results could only be obtained based on an energy criterion by considering the energy release rate G f and, hence, promoted the use of tension softening model for cracked concrete, that is, by the inclusion of a descending branch of the tensile stressstrain curve. Nevertheless, proper regularization tools must be employed in conjunction with the tension softening model in order to conserve the energy dissipation rate in crack formation. The direct use of a softening stress-strain curve, like the strength criterion, can also lead to pathological mesh sensitivity (Baant, 1976; Crisfield, 1982). This is

61

due to the fact that, when cracking takes place in a smeared crack model, the damage tends to localize into a single band of element. In addition, energy dissipation per unit volume in a smeared crack model is governed by the local strain softening constitutive law. Consequently, the energy dissipation decreases to zero as the mesh size reduces to an infinitely small volume, which inevitably causes spurious mesh sensitivity. With the increasing applications of fracture mechanics to continuum modelling of quasi-brittle materials like concrete, many regularization approaches have been developed in the last decade. A detailed review of a number of regularization models for quasi-brittle materials was recently presented by Baant and Jirsek (2002). Regularization of strain localization serves a mathematical purpose to prevent the loss of well-posedness of the incremental boundary value problem, which generates an infinite number of solutions when the material stability (Druckers stability) is lost (Benallal et al., 1988; de Borst et al., 1993c, 1998). The earliest models of its kind are the models based on the Cosserat theory (Cosserat and Cosserat, 1909), which was originally introduced to capture the mechanical behaviour of material due to microstructural heterogeneity on the mezzo scale. The theory suggests that deformation in material originates not only from the conventional translational deformation but also the rotational movement of the material particles. Attempts have been made to regularize strain localization problems in strain softening materials by using the Cosserat theory (Mhlhaus and Vardoulakis, 1987; Mhlhaus, 1991; Mhlhaus et al., 1991). However, it has been replaced by the non-local approach that has gained much popularity over the past 10 years, or so.

2.6.5.1

Non-local Models

Most engineering problems have been successfully solved by the use of local continuum models, for which the stress at a point in the continuum is dependent solely on the strain at the same point. The local continuum models work perfectly well in cases where the characteristic size of the material heterogeneity is relatively small compared to the size of the structure. Although strain localization of concrete is a macroscopic phenomenon, the process of fracture takes place at the scale of the heterogeneity, which involves dense microcracking in the fracture zone. Consequently,
62

apart from the mathematical shortcoming, the local continuum models, which are characterized by the local stress and strain of a point, likewise, are not able to accurately describe the physical process of fracture in concrete. While a local continuum model describes the local stress and strain, the stress in a non-local continuum depends not only on the strain at that point but also on the weighted averages of a state variable in the neighbourhood within a distance from that point. The spatial interaction of the non-local approach essentially introduces an internal length scale (de Borst and Mhlhaus, 1992; de Borst et al., 1993a) or a characteristic length (Baant and Pijaudier-Cabot, 1988; Baant and Obolt, 1990) into the constitutive model so as to ensure a constant energy release rate in the fracture process zone. The notion of non-locality emerged in the late nineteenth century and was later studied extensively in elastic materials (for example, Eringen, 1965; Krner, 1967; Eringen and Edelen, 1972). The non-local concept was first applied to strain softening damage materials by Baant et al. (1984). In general, the non-local approach involves the replacement of a local state variable f by a weighted average state variable f . The condition for effective nonlocal modelling is that the selected state variable must be able to reflect the damage evolved in the strain softening process. The averaged state variable is obtained by the use of a non-local averaging operator given by (Jirsek, 1998)

f (x) = ' (x, s) f (x) dV


V

(2.58)

where x and s are the vectors representing the locations of the local variable and the neighbourhood variables, respectively (Figure 2.23a), V is the volume of the structure and ' is a normalized weight function defined as

' (x, s) =

(| x s |)

( | x s | ) dV

(2.59)

where is a scalar weight function shown in Figure 2.23b, which is a function of r = | x s | , the distance between point x and point s. The weight function usually takes the form of the Gaussian distribution function as
63

r2 (r ) = exp 2l 2 ch

(2.60)

where lch is the characteristic length of the non-local continuum. Otherwise, a bellshape polynomial function may also be used (Baant and Obolt, 1990). That is

(r ) = 1

r2 R2

(2.61)

where the pointer bracket denotes a Macauley bracket, in which x = x for x > 0 and x = 0 for x 0 , R is the non-local interaction radius which may be related to the characteristic length lch . y lch s x x (a) (b) s

Fig. 2.23 - Non-local continuum: (a) representative volume for non-local averaging; (b) weight function and the relation to the characteristic length. (Baant and Planas, 1998).

The first non-local model for strain softening is the imbricate continuum model proposed by Baant et al. (1984). In this model, the strain was taken to be the state variable for non-local averaging. This model was found to guarantee mesh insensitivity, however, the differential equations for the equilibrium and boundary conditions involved were way too complicated for straightforward numerical implementation (Baant, 1990). The consequence of this is the introduction of the concept of the nonlocal continuum with local strain in which the strain is kept local but the constitutive
64

relation for strain softening is dependent on the non-local state variable that causes strain softening. The non-local damage model (Pijaudier-Cabot and Baant, 1987; Baant and Pijaudier-Cabot, 1988) was developed based on this concept and has proved to be successful in regularizing strain localization problems. The model was built on the framework of continuum damage mechanics and non-local averaging was performed either on the damage energy release rate or the damage variable. Several variants of the non-local damage models were attempted by using different non-local variables. For example, Baant and Lins (1988a) non-local smeared cracking model based on the non-local total strain formulation and used the model to investigate structural size effect. They showed that the model is capable of rectifying mesh directional bias that exists in a local smeared crack model. Jirsek and Zimmermann (1998b) formulated a non-local rotating crack model and evaluated the stress by multiplying the stiffness matrix with the total strain. The stiffness matrix was calculated based on the non-local total strain. Jirsek (1998) summarized the possible non-local formulations with various non-local state variables and compared the characteristics of each model. The non-local concept has also been extended to the microplane model (Baant and Obolt, 1990; Obolt and Baant, 1996) and a number of plasticity-based models (Baant and Lin 1988b; Nilsson, 1997; Borino et al., 1999).

2.6.5.2

Gradient Models

The mesh sensitivity of strain softening may also be prevented by the use of a gradient model. The basic idea of a gradient model is that, the stress at a point is not only a function of the strain but also the first or second spatial derivatives of the strain (Baant, 1990). The term gradient originates from the notion of the spatial derivatives which can be perceived as the gradients of strains with respect to the distance away from the point under consideration. The gradient model may be derived from the nonlocal model by expanding the spatial integral given by Eq. 2.58 into a truncated Taylors series and retaining the even-order derivative (Baant, 1984; Lasry and Belytschko, 1988; de Borst, 1997). That is

f = f + c 2 f
65

(2.62)

where c is a gradient coefficient depending on the type of the weight function used and is a function of the internal length scale, 2 is the Laplacian operator defined as
2 =

i 2 / xi ,

in which xi (i = 1, 2, 3) is the Cartesian coordinate. Due to the

approximated non-local nature of Eq. 2.62, the gradient model is also known as the weakly non-local model (Baant and Jirsek, 2002). The state variable f of the gradient models is usually a specific form of strain component depending on the type of constitutive model where the gradient formulation is built on. For example, in the ~ framework of continuum damage mechanics, the equivalent strain (see Eq. 2.47) is often taken as the gradient dependent state variable (Peerlings et al., 1996). While in a plasticity model the plastic strain is used for the evaluation of the gradient term (de Borst and Mhlhaus, 1992; Pamin and de Borst, 1995). One major disadvantage of the original gradient model is the need for the use of higher order elements to ensure continuity in both the displacement and strain fields within the element. In the work of Peerlings et al. (1996), an implicit gradient formulation was developed. The gradient dependent state variable is evaluated by solving a partial differential equation rather than a straightforward expression of Eq. 2.62. The solution of the partial differential equation was shown mathematically to be a special case of the fully non-local model, thus alleviating the complication arising from the use of higher-order elements.

2.6.5.3

Crack Band Formulation as Partial Regularization

The crack band model (Baant and Oh, 1983) is the simplest regularization method that is capable of preventing spurious strain localization on mesh refinement. The formulation of the model is discussed in Section 2.6.4.3. The crack band model is probably the first fracture model that has successfully promoted the use of an energy criterion in smeared crack modelling for a strain softening material. As mentioned earlier, the inclusion of a characteristic length into constitutive models for strain softening is of crucial importance to produce objective post-peak results. In the crack band model, the characteristic length is the so-called crack band

66

width, which is the width of the element undergoing fracturing. The basic idea of the model is to ensure a constant energy dissipation rate by adjusting the softening stressstrain relationship according to the size of the element and, hence, guaranteeing an objective tensile response. However, the fact that fracture process takes place numerically within an element is physically unjustified since the fracture process zone indeed holds a finite width at the fracture front. Moreover, the crack band model is also known to suffer from mesh directional bias. This aspect has been studied extensively, for example, by Rots (1988), Li and Zimmermann (1998) and Jirsek and Zimmermann (1998a).

2.6.5.4

Regularization by Inclusion of Material Viscosity

The mesh sensitivity arising from strain softening can also be prevented by the inclusion of material viscosity into the constitutive equation for dynamics problems. This approach was proposed by Needleman et al. (1988) for damage in metals and has been applied to strain softening material by Sluys (1992). The inclusion of viscosity indirectly introduces an internal length scale into the constitutive model and regularizes the mesh sensitivity problem. However, Baant and Jirsek (2002) pointed out that the viscosity-induced regularization is highly dependent on the load duration. The modelling results may not be objective if the load duration is much larger than the relaxation or retardation time associated with the type of viscosity employed.

2.6.6

Modelling of Steel Reinforcement

Three distinct approaches are available to represent steel reinforcement in reinforced concrete modelling, namely distributed steel formulation, embedded steel formulation and discrete steel approach (ASCE Task Committee, 1982, Chap. 3) (see Figure 2.24). In the distributed steel formulation, reinforcing steel is assumed to be smeared over concrete elements at a particular angle of orientation and is often described by the reinforcement ratio. The total stiffness of a distributed reinforced concrete finite element consists of the stiffness of the concrete and the stiffness contributed by the
67

smeared steel reinforcement. When the reinforcement is assumed to be smeared, a full compatibility between steel and concrete is naturally enforced. This type of formulation is useful for the analysis of reinforced concrete structures with densely distributed reinforcement, since the exact definition of every single reinforcing bar can be avoided. In the embedded steel formulation, each reinforcing bar is considered as an axial member incorporated into the concrete element by the principle of virtual work. The displacements of the embedded steel are consistent with the displacements of the concrete element. The major advantage of the embedded steel formulation is that the reinforcing steel can be defined arbitrarily regardless of the mesh shape and size of the base concrete element. The discrete steel approach is based on the use of separate elements to represent the reinforcing steel. One-dimensional truss elements are commonly adopted since reinforcing steel is usually assumed to carry axial load only. Steel truss elements are overlayed onto the boundary of the concrete elements by connecting the nodal points. This approach greatly facilitates the inclusion of bond-slip effects between steel and concrete, which may be achieved by inserting bond-slip elements between the concrete elements and the steel truss elements. A major disadvantage of this approach is that the mesh boundary of the concrete element must overlap the direction and location of the steel reinforcement.

(a)

(b)

(c)

Fig. 2.24 - Modelling of reinforcement: (a) distributed steel approach; (b) embedded steel; (c) discrete steel approach.

68

2.6.7

Modelling of Steel-Concrete Bond

The bond transfer between steel and concrete is closely related to the formation of distributed cracking in reinforced concrete structures. This is demonstrated in Section 2.5.2. Consequently, the correct prediction of cracking depends on a realistic modelling of the steel-concrete bond action. In numerical models, two different approaches are commonly employed to model bond between steel and concrete. The first includes a wide range of smeared crack models. The inclusion of bond-slip is made possible in an indirect manner by accounting for the tensile force carried by the intact concrete between the cracks, namely tension stiffening. The second approach describes the local bond-slip between steel and concrete using discrete bond elements, which actually takes into account the relative displacement of the reinforcing steel and the adjacent concrete.

2.6.7.1

Tension Stiffening

The tension stiffening effect has been incorporated in several different ways. One approach, developed by Scanlon and Murray (1974), is to adopt a descending branch in the concrete tensile stress-strain curve. An alternative approach proposed by Gilbert and Warner (1978) is to modify the stress-strain relationship of the tensile steel to indirectly model the tension in the concrete and to assume that the concrete possesses zero stiffness after cracking. These approaches can account for the effects of cracking on the macroscopic deformation of the member or structure but not at the mezzo or micro levels. In the modified compression field theory of Vechhio and Collins (1986), tension stiffening is modelled using a descending concrete stress-strain curve calibrated using their test data. In the disturbed cracked field model of Vecchio (2000), an improved tension stiffening descending function was used, which is made dependent on factors such as reinforcement ratio, diameter of reinforcing bar, direction of crack and orientation of reinforcements.

69

Hsu and Zhang (1996) proposed the use of both the average descending curve in tension of concrete and the average stress-strain curve of reinforcing bars embedded in concrete to model the effect of tension stiffening. They advocated that the average yield stress of an embedded reinforcing bar should be lower than the yield stress of a bare reinforcing bar. Balakrishnan and Murray (1986, 1988) employed an embedded reinforcement formulation including the effect of bond-slip by using the concept of virtual work. In addition, they adopted an approximated triangular concrete stress distribution between cracks in which the maximum stress was taken to be the concrete tensile strength and, hence, an average tension stiffening stress of 0.5 f ct . In this model, it was assumed that the tension stiffening stress vanishes completely as soon as the reinforcing bars begin to yield. The tension chord model (TCM) proposed by Marti et al. (1998) utilizes a stepped, rigid-perfectly plastic bond-slip model for the calculation for both crack spacing and crack widths, together with the average concrete stress between the cracks and, hence, the tension stiffening effect. Foster and Marti (2002, 2003) implemented the two-dimensional version of the TCM, the Cracked Membrane Model, into their finite element program (CMM-FE) and demonstrated promising results for the modelling of reinforced concrete structures in plane stress.

2.6.7.2

Discrete Bond Modelling

In this approach, the modelling of bond involves the use of discrete bond elements to simulate the relative displacement between steel and concrete. These bond elements can be subdivided into two general types, the bond link element and the interface element. The former was developed by Ngo and Scordelis (1967). Bond is modelled by placing a bond link element between each individual node along the concrete-steel interface. The element consists of two orthogonal springs that transmit shear and normal tractions between the bond-slip surfaces as shown in Figure 2.25a.

70

The second type, the interface element (Figure 2.25b), was introduced by Goodman et al. (1968) for modelling rock joints. The interface elements are now commonly used for modelling discrete crack propagation (ervenka, 1994) and bondslip phenomena (Mehlhorn and Keuser, 1985; Rots, 1985; Rots, 1988) in reinforced concrete structures. For modelling of bond, the interface elements are placed continuously along the concrete-steel interface. In addition, since the element is formulated based on the continuous relative displacement field between steel and concrete, it can describe the continuous slip between the two materials more effectively than the bond link element, which can only model bond-slip at individual nodes. Recently, an elasto-plastic cyclic bond model was developed (Lundgren, 1999) and implemented using the interface elements of the finite element program DIANA. A damaged and undamaged zones concept was introduced to describe the reduced bond friction due to the damage of concrete caused by the displacement of the rib of the deformed bar during a reverse slip process. The model was used to simulate monotonic and cyclic pull-out tests and frame corners subjected to closing moments, and showed a good correlation with test results. This demonstrates that the use of interface elements with a reliable bond model can in fact produce promising results for phenomenological modelling of reinforced concrete structures.

u4 2 u3 u8 4 u1 u7 u2 u1

u2

u6 3 u5 u4 u 3 2

s 1

(a)

(b)

Fig. 2.25 - Discrete bond models: (a) bond link element; (b) bond interface element.

71

2.6.8

Computational Creep Modelling

One of the biggest problems in implementing the superposition integral of Eq. 2.9 or Eq. 2.11 is the necessity for storing the history of stresses or strains. Stresses in a concrete structure are constantly changing with time due to cracking and redistribution of stresses between the concrete and the reinforcement. Consequently, a large amount of computer memory is needed to store every single stress increment throughout the analysis. The analysis of a structure with a large number of elements will be extremely expensive computationally. In order to avoid the explicit storage of stress or strain histories, the superposition integral may be converted into a series of differential equations which is often known as the rate-type constitutive equations (Baant, 1982; Baant, 1988). The stress or strain histories are stored internally in the rate-type constitutive equations. To facilitate such a conversion, the kernel of the superposition integral, the compliance function J (t , t ' ) or the relaxation function R(t , t ' ) , must be written in the form of the degenerate kernels as 1 (t t ' ) / 1 e =1 (t ' )

J (t , t ' ) =

(2.63)

and R (t , t ' ) =

=1

L (t ' ) e

(t t ' ) /

(2.64)

where C and L are functions of concrete age t ' , is called the retardation time for a compliance function or, the relaxation time for a relaxation function. Expanding the summations of Eq. 2.63 and Eq. 2.64 gives a series of real exponentials, which is called the Dirichlet series. Although the approximation of the Dirichlet series is not exact, its accuracy is regarded as sufficient for the purpose of creep analysis. The ratetype constitutive equations may be formed by combining the degenerate kernel with the superposition integral. This is done by substituting Eq. 2.63 into Eq. 2.9 or, Eq. 2.64 into Eq. 2.11. After some mathematical manipulations, it can be shown that the rate-

72

type constitutive equations may be represented by some sort of rheological models (Baant and Wu, 1973, 1974). For the integral-type creep law for aging material using the compliance function, the rate-type constitutive equation may be described as a series of Kelvin chain units as shown in Figure 2.26a. The total strain for a uniaxially loaded member may be written as the sum of the strains of the Kelvin chain units and the shrinkage strain and is given by

(t ) =

=1

(t ) + sh (t )

(2.65)

Each Kelvin chain unit is characterized by a spring and a dashpot connected in parallel, in which the properties are described by the spring elastic modulus E and the dashpot viscosity , respectively. Baant and Wu (1973) showed that E and can be related to the parameters of the Dirichlet series shown in Eq. 2.63. In the case of superposition of stress with relaxation functions for an aging material, Baant and Wu (1974) deduced that the Maxwell chain model could best describe the resulting rate-type constitutive equation from the superposition integral in Eq. 2.11. The Maxwell chain unit is made up of a spring and a dashpot connected in series as shown in Figure 2.26b. Considering a uniaxial case, it is evident in Figure 2.26b that the total stress is equal to the sum of the stresses carried in each Maxwell chain unit. This can be written as

(t ) =

=1

(t )

(2.66)

Similar to the Kelvin chain unit, the spring and the dashpot in the Maxwell chain unit are characterized respectively by an elastic modulus E and a viscosity corresponding to the -th unit of Maxwell chain.

73


=1

Maxwell chain unit N


=N

1 2

1
=1

2
=2

3
=3

=2

cp

Kelvin chain unit

=N

N sh sh

(a)

(b)

Fig. 2.26 - Rheological models: (a) integral-type creep model with Kelvin chain models in series; (b) integral-type relaxation model with parallel Maxwell chain models.

The remaining task left for the computation of the total stress or strain at time t is to solve the resulting differential equations (or rate-type equations) for or . To make possible an efficient algorithm for solving the differential equations, Baant and Wu (1973, 1974) introduced an exponential algorithm based on the quasi-elastic incremental stress-strain approach (Baant, 1971, 1972). The algorithm permits the use of relatively large time steps in a time analysis and shows good numerical stability. A different form of exponential algorithm was developed by Kabir and Scordelis (1979). They utilized the Dirichlet series expansion and expressed the temperaturedependent creep compliance in the following form C (t , t ' , T ) = i 1 e i s (T ) (t t ')
i =1 N

(2.67)

where i is a scale factor depending on the age at loading, i is a constant determining the shape of the creep curve and s is a shift function depending on the temperature T.

74

Kabir and Scordelis followed the algebraic manipulation of Zienkiewicz and Watson (1966) and derived a set recursive equations for calculating the creep strain increment cp . The uniaxial version of the equations is given by cpn = Ain 1 e i (Tn ) tn
i =1 N

(2.68)

where A i n = A i n 1 [e i (Tn 1 ) t n 1 ] + ai (t n ) n A i n = ai (t n ) n for n > 1 for n = 1 (2.69) (2.70)

The subscript n denotes the number of iteration. The storage of the total stress history is not needed and only the stress history of the last time step is required for the computation of the current creep strain increment. This approach has also been used by Van Zyl and Scordelis (1979), Van Greunen (1979) and Kang and Scordelis (1980) in time-dependent finite element analysis of concrete structures.

75

CHAPTER 3 FINITE ELEMENT MODELS FOR REINFORCED CONCRETE

3.1 Introduction
The investigation of behaviour of reinforced concrete structures is often undertaken by gathering test data from experiments and then following up with a detailed theoretical study. This method of investigation is often viable for the study of instantaneous structural behaviour. It is, however, expensive and time consuming for the investigation of long-term behaviour of structures. The finite element method provides an extremely powerful numerical tool that is able to accurately simulate structural behaviour if the appropriate modelling approach and material laws are effectively utilized. The major aim of this study is to develop a finite element model which can provide reliable predictions of long-term behaviour of cracked reinforced concrete structures so as to facilitate a parametric study of cracking and crack development based upon the data produced by the finite element model. In this way, the parameters affecting the development of a crack with time may be identified and quantified. In this chapter, the detailed formulation of the proposed numerical models of reinforced concrete structures subjected to sustained service loads are presented, including the finite element formulation and the iterative numerical solution procedures.

76

3.2 Continuum Modelling


In this study the smeared crack approach is adopted. The pros and cons of the smeared crack approach were discussed in Section 2.6.2. Two distinct approaches are employed to simulate the development of cracks in concrete, namely the distributed cracking approach and the localized cracking approach. In the first approach, cracks are stabilized by the reinforcement and are smeared over a region of cracked concrete (or reinforced concrete) elements. No discrete crack is observed in the distributed cracking approach and cracking is represented on the average level by a region of element with reduced stiffness. Hence, this approach may be regarded as an average model which takes account of the average effects of each component contributing to the global structural stiffness. There are no distinct cracks, so the estimation of crack spacing and crack width is problematic. In contrast, the localized cracking approach aims to capture the phenomenon of strain localization of concrete. Cracks are modelled as discrete localizations. Fracture of concrete is one of the main ingredients in this approach. For reinforced concrete, the transfer of bond between reinforcing steel and concrete are of significant importance to the formation of a crack as discussed in Chapter 2. The localized cracking approach fuses non-linear fracture mechanics of concrete and discrete bond action between steel and concrete in order to capture the formation of cracks at discrete spacings.

3.3 Distributed Cracking Approach


To accurately model crack development in reinforced concrete structures using the distributed cracking approach, a reliable model that can predict crack spacing and tension stiffening rationally is the key determinant. The cracked membrane model (Kaufmann, 1998; Kaufmann and Marti, 1998) is used to model two-dimensional plane stress reinforced concrete elements and is implemented in the finite element model. This model is formulated based on the assumptions of the tension chord model (Sigrist, 1995), which was originally developed to analyse problems of cracking and minimum reinforcement in reinforced concrete members subjected to uniaxial stress.

77

The fundamental concept can be illustrated by taking an orthogonally reinforced concrete panel subjected to in-plane stresses x , y and xy with parallel cracks spaced uniformly as shown in Figures 3.1a and 3.1b. The applied forces on the panel are equal to the sum of the force components contributed by the concrete and the steel in the global directions. Considering equilibrium of the infinitesimal elements at the crack shown in Figures 3.1c and 3.1d gives

x = cn cos 2 + ct sin 2 + cnt sin(2 ) + x sx y = cn sin 2 + ct cos 2 cnt sin(2 ) + y sy xy = 0.5( cn ct ) sin(2 ) cnt cos(2 )

(3.1a) (3.1b) (3.1c)

where the stresses with subscript c represent the concrete stresses and the subscript s denotes the reinforcing steel stresses. The element reinforcement ratios in the x and y directions are given by x and y , respectively. The t-n axis system denotes the crack direction (t) and the direction normal to the crack (n) and is the angle between x and n axes (-/2 /2) (see Figure 3.1b). (a)
(a)
y

(c) xy

(c)

cncos cnt cos cnt sin ct sin

x sx

x y

x xy

(b)
(b)

(d)
(d)

y sy

ct cos cnt cos xy

cn sin cnt sin

y 1 Fig. 3.1 - Orthogonally reinforced concrete panel subjected to in-plane stresses: (a) applied stresses; (b) axis notation; (c) and (d) stresses at crack. (after Foster and Marti, 2003).
78

3.3.1

Tension Chord Model

Consider a uniaxial reinforced concrete tension chord element with a single reinforcing bar of diameter , as shown in Figure 3.2a. Sigrist (1995) proposed that the bond between the steel and the concrete could be modelled using a stepped, rigid perfectly plastic bond shear stress-slip relation (Figure 3.2c). The bond shear stress drops from

b0 to b1 after yielding of the reinforcing steel. From this Sigrist (1995) and Marti et
al. (1998) developed an idealised model known as the tension chord model. The following outlines the derivation of Sigrists tension chord model. For a differential element of the tension chord element (Figure 3.2b) of length dx, the variation of steel and concrete stresses along the element length are described by d s 4 = b ; dx d c 4 b = dx (1 ) (3.2)

respectively, where is the ratio of the cross sectional area of steel As to the cross sectional area of concrete Ac of the tension chord and b is the bond stress. For a cracked tension chord, the average stress midway between cracks is obtained by integrating Eq. 3.2 over one half of the crack spacing ( s rm 2) and noting that the tensile stress cannot exceed the tensile strength, f ct . Therefore

4 c = (1 )

srm 2 0

b dx

f ct

(3.3)

When c = f ct , the crack spacing is at its maximum and solving Eq. 3.2 gives f (1 ) s rm0 = ct 2 b0 (3.4)

where s rm0 is the maximum crack spacing for a tension chord with a fully developed crack pattern. Marti et al. (1998) showed that the minimum crack spacing is one half of

79

the maximum crack spacing and concluded that the range of possible crack spacings for a fully developed crack pattern is s rm = s rm0 with 0.5 1.0 . The mean steel stress and maximum steel stress (at the cracks) can be expressed as a function of the mean (or average) strain m given that the distribution of the bond shear stress is known. Hence the stresses in the steel and the concrete between the cracks can be determined for any known stresses in the steel at the cracks by considering the equilibrium across the tension chord at the crack and anywhere between the cracks. (3.5)

srm

Ac

dx x (a)
N N

b b0

dx c+ dc b c s+ ds b s

b1

(b)

y (c)

Fig. 3.2 - Tension chord model: (a) reinforced concrete tension chord element; (b) differential element; (c) bond shear stress-slip relationship. (after Marti et al., 1998).

80

Following a similar derivation to that followed for Eq. 3.3, the concrete stress at a distance x from the crack can be shown to be

c ( x) =

4 b0 x (1 )

(3.6)

Knowing the concrete stress and the steel stress at crack sr , the steel stress at distance x from the crack can be obtained by considering the equilibrium at the crack and at the section at distance x from the crack as

s ( x) = sr c ( x)

(1 )

= sr

4 b0 x

(3.7)

The formulation given above was derived for a reinforced concrete member subjected to direct tension. To apply the tension chord model to reinforced concrete flexural members, the reinforcement ratio in the crack spacing equation of Eq. 3.4 is replaced by the effective reinforcement ratio eff , which is defined as the ratio of the area of tension steel Ast to the effective area of concrete in tension Ac.eff . In this study the Ac.eff is taken as that suggested in CEB-FIP Model Code 1990 (1993) and is defined as the product of the section width and a depth equal to 2.5 times the distance from the tensile face of the section to the centre of the steel, but not greater than one third of the depth of the tensile zone of the cracked section.

3.3.2

Cracked Membrane Model

Kaufmann and Marti (1998) developed the cracked membrane model (CMM) for the analysis of reinforced concrete membranes subject to in-plane stresses. The CMM combines the essence of the modified compression theory (Vecchio and Collins, 1986) and the tension chord model of Marti et al. (1998). The assumptions of the original CMM are: the crack faces are stress free and able to rotate normal to the direction of the principal major strains, giving cn = cnt = 0 . In this study concrete tension

81

softening is accounted for, therefore the crack faces are only free of shear stresses but the residual tensile cohesive stresses normal to the crack are not negligible. For equilibrium across the continuum, stresses in the x and y directions can be written in terms of the principal stresses in the 1-2 coordinate system as

x = c1 cos 2 + c 2 sin 2 + x sx + ctsmx y = c1 sin 2 + c 2 cos 2 + y sy + ctsmy xy = 0.5( c1 c 2 )sin (2 )

(3.8a) (3.8b) (3.8c)

where ctsmx and ctsmy are the mean concrete tension stiffening stresses in the x and y directions, respectively. By adopting the tension chord model, the crack spacing for a uniaxial element in the x and y directions, s rmx and s rmy , are calculated using Eq. 3.4 and Eq. 3.5 (Figure 3.3). As a simplification of the CMM, it can be shown that the Vecchio and Collins (1986) crack spacing equation is justified (Kaufmann and Marti, 1998; Foster and Marti, 2003) and the mean spacing between cracks may be taken as cos sin s rm = + s rmx s rmy
1

(3.9)

With the crack spacing known from Eq. 3.9, the instantaneous crack width is calculated considering the elasticity across the continuum and is given by

wcr = s rm [ 1 + 12 2 ctsm1 ] Ec

(3.10)

where 1 and 2 are the concrete strains in principal directions, 12 is the Poissons ratio in the 1-direction resulting from the stress applied in the 2-direction and ctsm1 is the mean concrete tension stiffening stress in the major principal direction.

82

sr
m

sr

sr

x ctsy srmx srmx ctsx

Fig. 3.3 - Crack spacing and tension stiffening stresses of an orthogonally reinforced concrete panel.

3.4 Localized Cracking Approach


Cracking in plain concrete is essentially a process of coalescing microcracks into a line of discontinuity which is not able to transfer stresses. This process is known as fracture of concrete and can be described theoretically by non-linear fracture mechanics, in which the formation of a unit area crack is treated as the consequence of dissipation of material fracture energy caused by tensile stresses. The details of non-linear fracture mechanics were discussed in Section 2.6.4. In the localized cracking approach, concrete and reinforcing steel are modelled as individual components. Localized cracking is simulated by the use of an appropriate concrete fracture model and the steel-concrete interaction is achieved by employing the bond interface element. This aims to provide a realistic description of the stress transfer between steel and concrete in the vicinity of a crack, which is of paramount importance in the prediction of the development of cracks at discrete locations (Section 2.5.2). The use of bond elements, either bond link elements (Ngo and Scordelis, 1967) or bond

83

srmy

srmy

cts

srmy

sr

interface elements (Mehlhorn and Keuser, 1985), is not new in the finite element modelling of reinforced concrete structures. However, to the authors knowledge, finite element studies focussing on the development of crack width and crack spacing, by coupling of bond elements with a reliable localized fracture model, are scarce. Furthermore, the inclusion of the effects of creep and shrinkage to study the timedependent development of crack width and crack spacing is even scarcer. In the context of a smeared crack modelling, fracture is described as a process of localization of inelastic strain into a band of finite width. The most formidable task in formulating a reliable fracture model with the smeared crack approach is to preclude the infamous mesh sensitivity issues. Appropriate models must be adopted in order to ensure that results do not vary significantly for different mesh sizes and configurations. In this study the crack band model (Baant and Oh, 1983) and the non-local model (Baant and Pijaudier-Cabot, 1988; Jirsek and Zimmermann, 1998) are employed to model fracture of concrete and are presented in the following sections.

3.4.1

Crack Band Model

An introduction to the crack band model (Baant and Oh, 1983) was given in Section 2.6.4.3. The basic idea of the model is identical to the fictitious crack model (Hillerborg et al., 1976) in which the fracture front (fracture process zone) is able to transfer stresses before a crack is completely formed. Three parameters are required to define the crack band model, namely the fracture energy G f , the material tensile strength f ct and a characteristic length known as the crack band width hc (the effective width of the cracking element normal to the crack). It is assumed that cracking is distributed uniformly over hc . Given that the average strain within the crack band, which is termed the fracturing strain or the cracking strain, is known, the crack opening displacement, wcr , is given by wcr = cr hc (3.11)

84

where cr is the cracking strain as depicted in Figure 3.4. Baant also introduced a parameter, the fracture energy density g f , which is defined as the amount of energy required to fully damage a unit volume of material. This parameter can be acquired from the area under the softening stress-strain curve (see Figure 3.4c), which gives an integral that is similar to Eq. 2.57 for the fictitious crack model. That is
u
0

g f = d cr

(3.12)

where u is the cracking strain when the crack is fully opened and the cohesive stresses between the crack faces completely vanish. By combining Eq. 2.57, Eq. 3.11 and Eq. 3.12, Baant showed that for a particular G f , g f needs to be adjusted according to hc in order to keep constant the energy consumed per unit extension of the crack band. This is given by g f = G f hc (3.13)

In the crack band model, the crack band width is taken as a constitutive material parameter. Baant and Oh (1983) examined empirically the effects of the crack band width with fracture tests for specimens of varying geometries and concluded that the

y
microcrack

fct hc gf cr

real crack

(a)

fracture process zone

u cr = wcr /hc

(b)

(c)

Fig. 3.4 - Crack band model: (a) crack path and the fracture process zone; (b) idealization of constant distribution of cracking strain; (c) tension softening stress-strain curve.

85

optimum crack band width for concrete is approximately three times the maximum aggregate size. However, it was found that within the range d a hc 10d a (where d a is the maximum aggregate size), the effect of varying the crack band width is not significant (Baant 1985).

3.4.2

Non-local Smeared Crack Model

The concept of a non-local continuum with local strain (Baant, 1990) was introduced to overcome the deficiencies in the imbricate continuum model (Baant et al., 1984) as discussed in Section 2.6.5.1. The main feature of the non-local continuum with local strain is that the strain at a point in a continuum is kept local but the constitutive relation for strain softening is dependent on the non-local state variable that causes strain softening. A typical non-local formulation of this type is the non-local damage model proposed by Baant and Pijaudier-Cabot (1988) and the constitutive relationship is given by

= [1 (Y ( ))]De

(3.14)

where the overbar denotes the non-local variable, De is the elastic stiffness matrix, is the damage variable and Y is the damage energy release rate which is a function of the non-local strain , which is calculated using the non-local averaging operator given by Eq. 2.58. Jirsek (1998) described this formulation in the context of the smeared crack approach. Eq. 3.14 can be written as

= Dsec ( )

(3.15)

in which Dsec is the secant stiffness matrix. No damage variable is involved in a smeared crack model, the secant stiffness directly accounts for the degradation of material integrity and therefore is taken as a function of the non-local strain. Jirsek and Zimmermann (1998) adopted this formulation and applied it to their rotating crack model with transition to scalar damage.
86

3.4.2.1

Issue Related to Non-local Continuum with Local Strain

The formulation given by Eq. 3.15 is investigated more closely herein. For simplicity, consider a one-dimensional case. The material constitutive law is assumed to be linear elastic and is followed by a linear softening stress-strain relationship, as shown in Figure 3.5a. According to Eq. 3.15, the stress tensor in a continuum is obtained by multiplying the secant stiffness matrix D sec evaluated from the non-local strain by the local strain tensor . Recalling the non-local averaging operator given by

(x) = ' (x, s) (x) dV


V

(3.16)

where x and s are the location vectors of the local strain and the neighbourhood strains, respectively, V is the volume of the structure and ' is a bell-shaped normalized weight function. The value of the weight function is largest at the location of the local strain (or called the target point) and it decreases for increasing distance away from the target point (see Figure 2.23b). These points in the neighbourhood are called the source points in the text that follows. At a target point in the continuum with tensile strain higher than the tensile strains in the surrounding neighbourhood source points, spatial averaging will inevitably scale down the strain at the target point since the normalized weight ' is smaller at the source points than at the target point. This decrease of the non-local strain is depicted in Figure 3.5b by the arrow. In the one-dimensional case, the secant stiffness matrix D sec is equal to secant modulus Esec and is obtained by connecting a line through the origin to the intersecting point of the non-local strain on the stress-strain curve. Knowing the secant modulus, the stress is computed as the product of Esec and , as shown in Figure 3.5c. It is evident in Figure 3.5c that, for the case of the non-local strain being lower than the local strain, the formulation given by Eq. 3.15 can potentially lead a stress level that is higher than the tensile strength of the material due to the high secant

87

modulus evaluated from the non-local strain. In some circumstances the over stress can be substantial. Though the formulation is not posing problems in structural fracture analysis, however, in the authors view, the violation of tensile strength from a structural perspective is undesirable.

fct Ec
1

fct Ec
1 1

(a)

Esec (b)

> fct fct Ec


1 1

Esec (c)

Fig. 3.5 - Non-local model with local strain: (a) bilinear tensile stress-strain curve; (b) computation of secant modulus from non-local strain; (c) stress as the product of secant modulus and local strain which is shown to be greater than the material tensile strength.

88

3.4.2.2

Proposed Non-local Smeared Cracking Formulation

In the light of the aforementioned issue, this section presents a variation of the nonlocal formulation that remedies the issue of Eq. 3.15. It is of crucial importance to retain a principal aspect of the non-local damage model proposed by Baant and Pijaudier-Cabot (1988), that is, only the causes of strain softening should be made nonlocal, while the elastic related variables are treated locally as in the conventional smeared crack model. This can be achieved by applying spatial averaging only to the cracking strain that exists in the fracture zone, which is a state variable that reflects the degradation of material integrity. The elastic part of the strain is kept local. With this approach, avoidance of the instability modes encountered in the imbricate continuum model (Baant et al., 1984) is guaranteed. The non-local formulation is then described by the following relationship

= D sec [ ce + cr ]

(3.17)

where ce and cr are the local elastic strain and the non-local cracking (or plastic) strain, respectively. In this formulation, the stress evaluation procedure is not dissimilar to the traditional method, the only difference is the replacement of the local strain by the sum of the local elastic part and non-local plastic part of the strain. To illustrate the methodology, again, the stress-strain relationship shown in Figure 3.5a, is considered. For the computation of stress, the strain is decomposed into an elastic component ce and a plastic component cr as shown in Figure 3.6a. To facilitate a comparison with the formulation given by Eq. 3.15, the same strain condition is applied, that is, the tensile strain at the target point is assumed to be higher than those of the source points in the surrounding neighbourhood. A non-local cracking strain cr with a lower value than the local cracking strain is obtained by applying the spatial averaging (indicated by an arrow in Figure 3.6b). The secant modulus Esec is evaluated from the new total strain by summing ce and cr as depicted in Figure 3.6b. The stress is computed by multiplying the total strain ( ce + cr ) by Esec .

89

With this approach, the issue of computing stresses higher than the concrete tensile strength is prevented. This is because the evolution of stress always obeys the constitutive stress-strain law in which the stress cannot exceed the tensile strength. In contrast, for the case of Eq. 3.15, the stress-strain law only provides a framework for the computation of the secant stiffness resulting from the effect of non-locality. Therefore, it does not guarantee the compliance of tensile strength criterion.

fct ce
1

cr Ec (a)

fct < fct ce

Ec cr
1

Esec (b)

Fig. 3.6 - Proposed non-local model: (a) decomposition of strain into elastic and plastic parts; (b) secant modulus is computed from the sum of elastic strain and non-local cracking strain and stress is calculated as the product of secant modulus and new total strain.

90

3.5 Orthotropic Membrane Formulation


The cracking models discussed above are implemented into an iterative secant based rotating crack model. The base element used in this study for the modelling of plain concrete is the four-node isoparametric element of Foster and Marti (2002, 2003) with the concrete taken as orthotropic. The material constitutive matrix D c12 in the principal directions is given by

12 Ec1 0 Ec1 E D c12 = Ec 2 0 21 c 2 1 12 21 0 0 (1 12 21 )Gc12


1

(3.18)

where is the Poissons ratio of concrete; Ec1 and Ec 2 are the concrete secant moduli in the principal directions and; Gc12 is the concrete secant shear modulus. To model the non-linear behaviour of concrete, Ec1 , Ec 2 and Gc12 are updated according to the stress states of the concrete. For plane stress problems in a global coordinate system, stress is related to strain by = D where = [ x (3.19)

y xy ]T and = [ x y xy ]T .

For known displacements and, hence, strains, the secant moduli are determined using a modified version of the equivalent uniaxial strain concept of Darwin and Pecknold (1977). This is done by removing the lateral deformation caused by the Poissons effect and is given by Foster and Marti (2003) as 1u 1 12 1 1 = 2u 1 12 21 21 1 2

(3.20)

where 1u and 2u are the equivalent uniaxial strains and 1 and 2 are the strains in the principal 1-2 directions. The secant moduli are obtained from the appropriate uniaxial stress-strain curve. The biaxial stresses are calculated from

91

c1 E c1 = c 2 0

0 1u E c 2 2u

(3.21)

The shear modulus is taken as that derived by Attard et al. (1996) and is given by Gc12 = 1 4(1 12 21 )

[Ec1 (1 12 ) + Ec 2 (1 21 )]

(3.22)

For cracked concrete, the Poissons ratios are taken as zero and Eq. 3.18 reduces to Ec1 D c12 = 0 0 0 Ec 2 0 0 0 Gc12

(3.23)

Prior to constructing the element stiffness matrix, the material constitutive matrix (Eq. 3.18 or Eq. 3.23) is transformed to the global coordinate system. That is D c = T T D c12 T (3.24)

where D c is the concrete constitutive matrix in the global coordinate system and T is the strain transformation matrix. The detailed formulation of a two-dimensional planar element will be discussed in Section 3.8.1. For the distributed cracking model based on the CMM, the stiffness matrices of steel reinforcement and concrete tension stiffening are added to the concrete constitutive matrix as D = D c + D s + D cts (3.25)

where D s and D cts are the material constitutive matrices of steel reinforcement and concrete tension stiffening, respectively, and are given by x E sx Ds = 0 0 0 y E sy 0 0 0 0

(3.26)

92

and Ectsx D cts = 0 0 0 Ectxy 0 0 0 0

(3.27)

where E sx and E sy are the secant moduli for steel reinforcement and Ectsx and Ectsy are the secant moduli of concrete tension stiffening. All moduli are expressed in the global x-y coordinate system. For the non-local model, the column vector of the equivalent uniaxial strain in Eq. 3.21 is replaced by the modified counterpart resulting from the effect of non-locality. It should be noted that non-locality only takes effect for concrete in tension. Amalgamating non-locality, Eq. 3.21 is written as 1 Ec1 = 2 0 0 1u Ec 2 2u

(3.28)

where 1u and 2u are the modified equivalent uniaxial strains due to the effect of nonlocality, which are calculated as the sum of the elastic component and non-local plastic component, that is

iu = iu iu = iu.e + iu.cr

for iu tpk for iu > tpk

(3.29a) (3.29b)

where iu.e (i = 1, 2) is the local elastic strain, iu.cr is the non-local cracking strain and tpk is the strain corresponding to the peak stress in the tensile stress-strain curve. Prior to decomposition of strain into elastic and plastic components, the local stress

loc is calculated from the local equivalent uniaxial strain iu via the material
constitutive law. The local elastic strain and the local cracking strain are calculated as

iu.e = loc Ec

(3.30a)

93

iu.cr = iu iu.e

(3.30b)

where Ec is the initial elastic modulus of concrete. The non-local cracking strain is computed as (Figure 3.7a)

iu.cr = ' (r ) iu.cr (r , ) dA


A

(3.31)

where the normalized weight function ' is given by

' (r ) =

(r ) (r ) dA

(3.32)

and is a polynomial bell-shaped weight function (Figure 3.7b) given by Baant and Obolt (1990) as

(r ) = 1 (r R )2

(3.33)

The pointer bracket denotes a Macauley bracket and R = 0.9086 l ch where lch is the characteristic length. R is also known as the interaction radius (Jirsek, 1998) as it represents the maximum distance of influence in the neighbourhood of the target point (Figure 3.7b).

Region A R r

Eq. 3.33 1

2R (a) shaped weight function. x 2R (b)

Fig. 3.7 - Non-local neighbourhood: (a) spatial averaging on region A; (b) bell-

94

3.6 Material Constitutive Models


A powerful numerical tool such as the finite element method requires reliable material models to perform an accurate numerical analysis. In this section, the material models adopted in this study are presented. The material models are used in conjunction with the orthotropic membrane formulation presented in Section 3.5 and implemented with an iterative non-linear solution procedure discussed in the section that follows.

3.6.1

Instantaneous Behaviour of Concrete

Though this work focuses mainly on the behaviour of reinforced concrete structures at service loads where the compressive stress in concrete rarely exceeds 40% of the compressive strength, the full compression behaviour to failure is included for the sake of completeness of the finite element model. For a two-dimensional concrete model, a realistic constitutive relationship for the biaxial state of stress is the key to the successful modelling of concrete. Three biaxial states of stress can be identified, namely the biaxial compression state, the combined compression and tension state and the biaxial tension state. It is important to apply reliable models for all three stress states so as to adequately describe the realistic characteristic of a reinforced concrete member. The instantaneous concrete model implemented in the finite element model is based on the concrete model used by Foster and Marti (2002, 2003). Instead of using a curved concrete strength envelope similar to that proposed by Kupfer et al. (1969), the strength envelope is approximated by a number of linear segments (Foster and Marti, 2003). Figure 3.8 shows the comparison of the strength envelope adopted herein and that of Kupfer et al. (1969). In the authors view, the accuracy of a linear-segmented strength envelope is not significantly different from that of a curved envelope nor does it cause discrepancies in analyses since the experimental data on which all models of the biaxial strength envelope are based are somewhat scattered in nature.

95

-2c / fc

0.4 8

1.0

(0.6, 1.25)

=1 .0

(1.15, 1.15)
Kupfer et al. (1969)

0.6

(1.25, 0.6 )

This study (after Foster and Marti, 2003)

-1c / fc 0.6 1.0

Fig. 3.8 - Strength of concrete under biaxial state of stress.

3.6.1.1

Stress-strain Relationships for Concrete

The strength of concrete in compression under biaxial states of stress differs from that in uniaxial compression. The compressive strengths in the principal stress directions are influenced by the orthogonal interaction of the principal stresses. The biaxial compressive strength of concrete f c* can be written as the uniaxial compressive strength f c' multiplied by a strength factor as
* f c = f c'

(3.34)

The factor can be thought as a scaling factor for the concrete compressive stress-strain curve in accordance with the state of stress. Thus, the concrete strain corresponds to the peak stress in a compressive stress-strain curve, cpk , is also multiplied by the same factor and is given by
* cpk = cpk

(3.35)

96

In this work, the uniaxial compressive stress-strain curve proposed by Thorenfeldt et al. (1987) is used as the base curve for the strength scaling purposes. The concrete compressive stress is expressed as

c = f c'

n n 1 + nk

(3.36)

in which

c cpk

and

n=

Ec Ec Ecpk

(3.37)

where c is the concrete strain, cpk is the concrete strain corresponding to the peak stress on the stress-strain curve, Ec is the initial elastic modulus of concrete and Ecpk is the secant modulus at the peak of the stress-strain curve that is Ecpk = f c' cpk . The negative sign in Eq. 3.36 indicates the stress in compression. The parameter k is a decay factor that controls the post-peak response and is given by Collins and Porasz (1989) as

k = 1.0 f c' k = 0.67 + 1.0 62

for

c cpk

(3.38a)

for

c > cpk

(3.38b)

The uniaxial compressive base curve and the scaled biaxial compressive stress-strain curves are shown in Figure 3.9a. The unloading modulus for concrete in compression is taken as that given by Filippou et al. (1983), that is

E cu =

c.un c.un (0.15 c.un + 0.1) cpk

c.un

(3.39)

where c.un is the concrete stress just before the commencement of unloading and

c.un is the strain corresponding to c.un on the stress-strain curve.

97

For tensile response of concrete, concrete is taken as a linear elastic material prior to cracking. After cracking, concrete undergoes tension softening before the bridging stresses at the crack completely disappear. The bilinear tensile stress versus crack opening displacement model of Petersson (1981) is adopted herein. The softening branch is converted to a stress-strain softening curve based on an energy-based scaling method similar to that of the crack band approach (Baant and Oh, 1983), as described in Section 3.4.1. The tensile stress-strain curve is shown in Figure 3.9b. Three softening parameters, 1 , 2 and 3 , are used to define the softening curve, where 1 1 = ; 3 2 2 = 3 + 1 ; 9

3 =

18 Ec G f 2 5 f ct h

(3.40)

where h is an average width over which the fracture energy is dissipated. For reinforced concrete distributed cracking problems analysed by the CMM, h is the crack spacing, while for concrete fracture problems, h is the width of the fracture zone. In other words, h is the crack band width for the crack band model and is the average width of the strain localization zone for the non-local model. The unloading modulus for concrete in tension is taken as

Etu =

c.un ce.un + 0.5 cr.un

(3.41)

where c.un is the same as that defined for Eq. 3.39 but in tension and, ce.un and

cr.un are the elastic and plastic parts of strain corresponding to c.un .

3.6.1.2

Biaxial Compression State of Stress

The biaxial compression state of stress is indicated in the first quadrant of Figure 3.8. It is evident in the biaxial strength envelope that the compressive strength of a concrete undergoing biaxial compression is higher than its uniaxial compressive strength. This is due to the confinement induced by the orthogonal compression struts in the principal stress directions. In the biaxial compression state of stress, is a confinement factor which can be determined from the biaxial strength envelope depending on the ratio of
98

the major and minor principal compressive stresses, = 1c 2c (see Figure 3.8). By combining and the equations for the linearized strength envelope within the biaxial compression state of stress, after some simple algebraic manipulation, the confining strength factor in the minor principal stress direction 2 can be expressed as

2 =

1.0 1 2.4 1.15 (1 + 5.5 1 ) 1 + 5.5

for

0 0.48

(3.42a)

2 =

for

0.48 < 1.0

(3.42b)

The confining strength factor in the major principal stress direction is then given by

1 = 2 .

3.6.1.3

Tension-Compression State of Stress

In the tension-compression biaxial state of stress, concrete undergoes tension in the major principal direction and compression in the minor principal direction. This is depicted by the second and fourth quadrants of the biaxial strength envelope shown in Figure 3.8. It is well known that the strength of concrete in compression is substantially reduced if a large tensile strain is present in the orthogonal direction (for example Vecchio and Collins, 1986; Miyakawa et al., 1987; Belarbi and Hsu, 1991). This phenomenon is known as compression softening. Therefore, is a strength reduction factor and is obtained from the modified compression field model of Vecchio and Collins (1986), which is given by

1 0.8 + 0.34

cpk

1.0

(3.43)

where 1 is the principal tensile strain. Figure 3.9c shows the influence of 1 to the scaling of the compressive stress-strain curve.

99

A tension cut-off regime is used in conjunction with the compression softening model. The tension cut-off is activated as soon as the compressive stress in the normal direction exceeds 60% of the uniaxial concrete compressive strength as shown in the biaxial strength envelope (Figure 3.8). The cracking stress f cr under the tension cutoff regime are given by

f cr = f ct
f cr = f ct 0.4 f c' ( f c' 2c )

for

0 2c f c' 0.6 0.6 < 2c f c' 1.0

(3.44a)

for

(3.44b)

The tension cut-off cracking stress is shown in Figure 3.9b in dashed line. Under service load conditions, the influence of the tension cut-off is not prominent as the concrete compressive stress in the normal direction is well below 0.6 f c' , for which the cracking stress is equal the concrete tensile strength.

3.6.1.4

Biaxial Tension State of Stress

The Rankine failure criterion is adopted for concrete under the biaxial tension state of stress, that is, the cracking is deemed to occur as soon as the tensile strength of concrete is violated regardless of the biaxial interaction of stresses. Therefore the full trilinear stress-strain relationship denoted by the solid lines in Figure 3.9b is used for this biaxial state of stress.

100

-c fc fc fc
> 1.0 Ec 1 1 Ecu < 1.0

c fct fcr 1 fct


1 Ec 1

cpk cpk cpk

-c

Etu

tpk

2tpk

3tpk

(a)
-c

(b)

-c

(c)

Fig. 3.9 - Stress-strain relationships for concrete: (a) scaling of biaxial compressive stress-strain curves; (b) tensile stress-strain curve with bilinear softening; (c) compressive stress-strain surface under biaxial tension-compression state of stress.

3.6.2

Time-dependent Behaviour of Concrete

The major factors affecting the time-dependent behaviour of reinforced concrete structures are creep and shrinkage of the concrete. At any time t after first loading, the vector of total strain is taken to be the sum of the vectors of instantaneous, creep and shrinkage strains. That is

101

(t ) = ci (t ) + cp (t ) + sh (t )

(3.45)

where ci , cp and sh are the instantaneous, creep and shrinkage strain vectors, respectively, and = [ x y xy ]T . Before cracking the instantaneous strain is

equal to the concrete elastic strain whereas the post-cracking instantaneous strain consists of an elastic component and a plastic component. For finite element implementation, creep and shrinkage strains are treated as inelastic pre-strains updated with time and applied to the structure as equivalent nodal forces. The details of the computational procedures will be presented in Section 3.7. The time-dependent development of shrinkage and concrete tensile strength are calculated using a function given by F (t ) = At B+t (3.46)

where A and B are empirically fitted parameters obtained from test control data and F (t ) is the shrinkage strain or concrete tensile strength at time t.

3.6.3

Shrinkage

Shrinkage is defined as the time-dependent and load independent strain resulting from the reduction in volume of concrete at constant temperature (due mainly to loss of water resulting from drying and hydration). Shrinkage is taken to be direction independent and shrinkage shear strain is taken as zero. Thus, the shrinkage strain vector is written as sh (t i ) = [ sh (t i ) sh (t i ) 0]T (3.47)

where sh (t i ) is negative and with magnitude calculated from Eq. 3.46 using appropriate factors for A and B.

102

3.6.4

Creep

Two creep models were developed in this study to simulate the time-dependent behaviour of concrete. In the early phase of the research, a simple creep model based on the rate of creep method (Glanville, 1930; Whitney, 1932; Dischinger, 1937; Chong et al., 2004) was employed. This method gives reasonably good creep prediction if the stress does not vary too much with time. A more refined model based on the principal of superposition was developed at a later stage using the theory of solidification for concrete creep (Baant and Prasannan, 1989a, b). This model is more versatile than the rate of creep method and can be used to handle creep problems with varying stress histories. For reinforced concrete structures under service load conditions, the concrete stress rarely exceeds 0.4 times the strength of the concrete. Accordingly, two assumptions are made in this study: (i) creep is linear with respect to stress; and (ii) the time-dependent response in tension is identical to that in compression. In the next section, the creep model based on the solidification theory will be described and this model is used to analyse reinforced concrete structures throughout this thesis. The model based on rate of creep model will only be presented in Appendix A.

3.6.5

Solidification Theory for Concrete Creep

The creep model used in this study is the solidification creep aging model of Baant and Prasannan (1989a) using Kelvin chains to describe the viscoelastic component (Figure 3.10a). In this model, the aging aspect of concrete creep is due to growth, on the microscale, of the volume fraction of the load-bearing solidified matter and is a consequence of hydration of the cement particles. An advantage of the solidification theory is that the elastic properties of concrete are taken as non-aging with the modulus of elasticity of concrete taken to be an age-independent asymptotic modulus, E0 . Baant and Baweja (1995a, b) proposed E0 = 1.6 Ec.28 , where Ec.28 is the elastic modulus at 28 days. A detailed discussion for the justification of the use of E0 is
103

presented in the works of Bazant and Prasannan (1989a) and Bazant and Baweja (1995b). The growth in volume of the solidified matter at time t is divided into the volume fractions associated with: (i) the viscoelastic strain v(t ) and; (ii) the viscous strain h(t ) as shown in Figure 3.10b. Using this approach, the creep strain is decomposed as cp (t ) = v (t ) + f (t ) (3.48)

where v and f are the viscoelastic and viscous strains, respectively. The viscoelastic and viscous strain rates are given by (Baant and Prasannan, 1989a) & v (t ) = & (t ) v(t ) (3.49a)

t & & (t ) = (t t ' ) d (t ' ) 0

(3.49b)

& f (t ) =

(t ) (t ) = 0 h(t ) (t )

(3.49c)

where t ' is the variable for concrete age at application of load, (t ) is the viscoelastic microstrain, (t t ' ) is the microscopic creep compliance function of the solidified matter associated with the viscoelastic component and 0 and (t ) are the effective viscosity of the solidified matter and the apparent macroscopic viscosity, respectively, associated with the viscous component. The viscous component is a linear function of stress calculated directly from Eq. 3.49c, whereas the viscoelastic component is evaluated by solving the superposition integral given by Eq. 3.49b. From these relationships, Baant and Prasannan (1989a) derived the analytical expression for the creep compliance of concrete t C (t , t ' ) = q 2 Q(t , t ' ) + q3 ln 1 + (t t ' ) 0.1 + q 4 ln t'

(3.50)

104

where q 2 , q3 and q 4 are empirical parameters determined from control tests and Q(t , t ' ) is a binomial integral. However, as no closed formed solution exists for the integral Q(t , t ' ) the approximation Q (t ' ) r (t ' ) f Q(t , t ' ) = Q f (t ' ) 1 + Z (t , t ' )
1 r (t ' )

(3.51)

is substituted (Baant and Prasannan, 1989a) where log Q f (t ' ) = 0.112 + 0.4308 log t ' + 0.0019(log t ')2 Z (t , t ' ) = (t ') 0.5 ln 1 + (t t ' ) 0.1 r = 1.7t ' 0.12 + 8

(3.52a)

(3.52b) (3.52c)

E1 E2
=1

1 2 1 2
= (t-t) d(t)

/E0 v
h(t) dh(t)

=2

v(t)

dv(t)

(t-t)

EN

=N

sh

(a)

(b)

Fig 3.10 - Solidification theory for concrete creep: (a) Kelvin chain description for viscoelastic component; (b) schematic representation of the solidification creep model. (Baant and Prasannan, 1989a).

105

The analytical expression for the creep compliance given by Eq. 3.50 is used in this study for the determination of the empirical parameters q 2 , q3 and q 4 for a given set of experimental creep data. In numerical implementation the explicit calculation of Q(t , t ' ) is not required.

3.6.5.1

Rate-type Constitutive Model

To facilitate the numerical creep analysis, the integral-type equation based on the principle of superposition, given by Eq. 2.9, is converted into a rate-type constitutive equation that allows the stress history to be stored implicitly. A discussion for the ratetype constitutive equations is given in Section 2.6.8. For the creep model adopted, the rate-type equation is described using the Kelvin chain model (Figure 3.10a). For an element in uniaxial state of stress, the relationship between the viscoelastic microstrain and the applied stress is E + & = (3.53a)

=1

(3.53b)

where , E and are the viscoelastic microstrain, the elastic modulus and the viscosity of the -th Kelvin chain unit, respectively, and N is the total number of Kelvin chains. For a constant stress applied at time t ' , the biaxial viscoelastic microstrain vector (t ) is obtained by solving Eq. 3.53a for and subsequently substituted into Eq. 3.53b, that is (t ) = 1 (t t ' ) (1 e ) =1E
N

(3.54)

where = E is the retardation time of the -th Kelvin chain unit.

106

By comparing Eq. 3.54 and Eq. 3.49b for a constant stress applied at time t ' , the microscopic creep compliance function may be expressed in the form of a Dirichlet series: (t t ' ) =

=1

(1 e

(t t ')

) + A0

(3.55)

where the A0 term is added to include the negative infinity area of the retardation spectrum (Baant et al., 1997) in the discretization of the spectrum that follows. Figure 3.11 shows the numerical integration of the retardation spectrum using the trapezoidal rule with intervals (ln ) . The relationship for the discretization of the Kelvin chains is given by A = L( ) (ln ) = L( ) ln 10 (log ) (3.56)

where A = 1 E and L( ) describes the retardation spectrum and is given by Baant and Xi (1995) as 0.02 (3 ) 2.8 [0.9 (3 ) 0.1 ] (3 ) 3 L( ) = q2 2 [1 + (3 ) 0.1 ]3 0.19 (3 ) 2.8 [0.9 (3 ) 0.1 ] 0.01 (3 ) 2.8 (3 ) 3 + q2 [1 + (3 ) 0.1 ] 2 2

(3.57)

Baant and Xi (1995) recommended that for a sufficiently smooth creep curve the retardation time discretization interval be taken as (log ) = 1 for each adjacent Kelvin chain. Using the log-power law proposed by Baant and Prasannan (1989a), the microscopic creep compliance function can be approximated as (t t ' ) = q 2 ln 1 + (t t ' ) 0.1

]
107

(3.58)

and substituting Eq. 3.58 into Eq. 3.55, the negative infinity area is A0 = q 2 ln 1 + (t t ' ) 0.1

=1

(1 e

(t t ' ) /

(3.59)

Retardation spectrum

L() A2
ln2

A0

A1
ln1

A3
ln3

A4
ln4

A5

A6

ln5

ln6

ln

Fig. 3.11 - Discretization of a continuous retardation spectrum.

3.6.5.2

Finite Element Implementation of Creep

The exponential algorithm proposed by Baant (1982, 1988) is adopted for finite element implementation. The use of an exponential algorithm permits the time step interval to be greater than the shortest retardation time and ensures numerical stability. Baant and Prasannan (1989b) introduced an incremental quasi-elastic stress-strain relationship for numerical analysis using the solidification creep theory. This algorithm, however, cannot be incorporated directly into the finite element formulation of this study. Therefore, modifications were made to facilitate the implementation of the creep model. The finite element model adopted in this study is a smeared crack model based on the total strain formulation. In a total strain formulation, the total stress is related to the total strain through a single constitutive relationship which is path-independent and the material response at any instant is a function only of the current state of stress or strain.

108

Consequently, the components that are required to be determined for the incorporation of the solidification creep model are the total viscoelastic and viscous strains at a specific time instance. The formulation is generalized for two-dimensional plane stress problems and is presented in the next section. By Eq. 3.49a, the change in recoverable component of creep, that is the viscoelastic component, can be expressed in the form v (t i +1 ) = i +1 vi + (3.60)

in which the subscripts i and i + indicate the reference to time t i and the time in the middle of a logarithmic time step ti + , respectively, where t i + = t 0 + [(t i +1 t 0 )(t i t 0 )] 0.5 (3.61)

and t 0 is the age at first loading. By Eq. 3.49b, Eq. 3.53b and Eq. 3.55, the change in viscoelastic microstrain is given by

i +1 =

=1

( i +1 i ) + G A0

(3.62)

The viscoelastic microstrain at time t i +1 for -th Kelvin chain used in this study is a modified form of that derived by Baant and Prasannan (1989b) and is i +1 = i e where t 1 e y
y y

1 G i 1 y (1 e ) + G E E

(3.63)

y =

= i i 1

(3.64)

For biaxial stress the Poisson effect is included via the matrix G in Eq. 3.62 and Eq. 6.63 where
109

1 G = 0

1 0

0 0 2(1 + )

(3.65)

The volume of the solidified matter at mid-time of a logarithmic time step, vi + , is then given by 1 q vi + = + 3 ti + q 2
1

(3.66)

The change in non-recoverable, viscous, component of creep is evaluated from Eq. 3.49c. By considering the change over a finite time step we write this as f (t i +1 ) G i = i + t

(3.67)

where i = i 1 + 2 . Substituting the apparent macroscopic viscosity, defined as i + = q4 1ti + , into Eq. 3.67, the change in viscous strain can be written as f (t i +1 ) = G i q 4 t t i + (3.68)

Lastly, the changes in viscoelastic and viscous strain components are added to the creep strain components obtained from the previous converged time step, that is v (t i +1 ) = v (t i ) + v (t i +1 ) f (t i +1 ) = f (t i ) + f (t i +1 ) (3.69a) (3.69b)

The sum of the creep strain components from Eq. 3.69 are then added to the shrinkage strains to give the total inelastic pre-strains, 0 . The inelastic pre-strains are converted to equivalent nodal forces and applied to the nodes of the discretized structure.

110

3.6.6

Time-dependent Crack Width

3.6.6.1

Cracked Membrane Model

Crack widths at time t are obtained in a similar way as described by Eq. 3.10. By taking the average strains between two cracks, the crack width equation has a general form that is given by wcr = s rm [ 1 betw.cr ] where betw.cr is the total concrete strain between the cracks. In a time-dependent analysis, the concrete tensile stress associated with tension stiffening induces creep deformation and drying shrinkage causes shortening in concrete between the cracks. Eq. 3.70 may be elaborated as (t ) wcr (t ) = s rm 1 (t ) ctsm1 + cp (t ) + sh (t ) 12 2 (t ) E 0 where ctsm1 is the mean tension stiffening stress in the crack opening direction. The concrete stress associated with tension stiffening and creep contribute to an expansion in concrete between the cracks. These effects obviously reduce the crack opening. Meanwhile, drying shrinkage causes a volume reduction in concrete. Since the distributed crack surfaces act as boundaries that separate concrete into individual blocks, volume reduction of each concrete block results in a gradual opening of the cracks. Due to the fact that the influence of shrinkage in widening the crack is far more dominant than the influence of concrete tensile stress and creep in closing the crack, the width of a crack in a concrete structure generally increases with time. (3.70)

(3.71)

3.6.6.2

Crack Band Model

Cracking of concrete is treated as strain localization in the crack band model. The effects of creep and shrinkage are only prominent in the intact concrete. Consequently,
111

the computation of time-dependent crack width for the crack band model is not dissimilar to that for instantaneous loading. The only difference is that the variables involved in the calculation of crack width are now functions of time. The timedependent crack width is given by wcr (t ) = hc cr (t ) = hc [ 1 (t ) ce (t )] (3.72)

where cr is the cracking strain of concrete at time t and is calculated by subtracting the elastic component of strain ce from the total principal strain 1 .

3.6.6.3

Non-local Model

One of the characteristics of the non-local model is that strain localization takes place over a width of a number of elements. The distribution of cracking strain across the width of the localization is not uniform but is concentrated in the centre of the localized band and gradually decreases to the edge of the localization. The crack band model has a well-defined localization boundary and cracking occurs uniformly over a single band of elements and therefore crack width can be calculated directly using Eq. 3.72. However, the non-local model is not as straightforward due to the non-uniform distribution of cracking strain. In this study a simple method is adopted to determine the crack width in the nonlocal model, that is, by measuring the relative displacement of any two nodes that are located adjacent to and outside of the localization zone. This method is crude and can only give a rough approximation of the crack width.

3.6.7

Stress-strain Relationship for Reinforcing Steel

The different stress-strain relationships of the two types of steel commonly used in reinforced concrete were introduced in Section 2.4. In this research, a general trilinear stress-strain model is adopted to model the reinforcing steel as shown in Figure 3.12a.

112

The reinforcing steel is assumed to be linear elastic before yielding of steel and is followed by a bilinear strain hardening. At reversal of stress, the unloading modulus E su is taken as equal to the initial elastic modulus of the steel E s . The trilinear model provides flexibility in describing a wide range of stress-strain relationships for reinforcing steel. For reinforced concrete structures under service load conditions, the stress in the reinforcing steel is generally below the yield stress therefore, for simplicity, a linear-perfectly plastic stress-strain law is sufficient in a finite element analysis. This is achieved by setting E w = Eu = 0 .

3.6.8

Local Bond-slip Model for Bond Interface Element

The CEB-FIP Model Code 1990 (1993) bond model is used in this study. The bond stress b between concrete and reinforcing steel under monotonic loading is given by the following functions at different slips, as s b = max 1 s 2
0.4

for 0 s s1

(3.73a)

b = max
s s2 b = max ( max f ) s s 3 2

for s1 < s s 2

(3.73b)

for s 2 < s s3

(3.73c)

b = f

for s3 < s

(3.73d)

The parameters used to define the bond model are given in Table 3.1 and the bond stress-slip curve is shown in Figure 3.12b.

113

s fu fw fsy
1 Eu

b max

Ew Esu 1

Es 1

Eb.sec Ebu

f u s s1 s2 (b)
1

sy w (a)

s3

Fig. 3.12 - (a) Trilinear stress-strain relationship for reinforcing steel; (b) bond stressslip relationship (CEB-FIP, 1993).

Table 3.1 - Parameters for bond stress-slip model (CEB-FIP, 1993). Unconfined concrete1 Parameters s1 s2 s3 Good bond conditions 0.6 mm 0.6 mm 1.0 mm 2.0 f c' 0.15 max All other bond conditions 0.6 mm 0.6 mm 2.5 mm 1.0 fc' 0.15 max Confined concrete2 Good bond conditions 1.0 mm 3.0 mm Clear rib spacing 2.5 f c' 0.40 max All other bond conditions 1.0 mm 3.0 mm Clear rib spacing 1.25 f c' 0.40 max

max f
1

splitting bond failure; 2 pullout bond failure.

The bond stress-slip model presented above is a time-independent relationship that does not account for the influence of the creep in concrete. Creep causes a gradual increase in slip with time and this phenomenon is known as bond creep. The consequence of bond creep is the reduction of slope of the time-independent bond stress-slip relationship. The isochrone curves recommended by the CEB-FIP Model Code 1990 (1993) is adopted to model the increase of slip with time (see Figure 3.13a).

114

The time-dependent slip st between steel and concrete under a sustained load is given by st (t ) = si (1 + 240 t ) 0.08 where t is the load duration in days and si is the instantaneous slip. For finite element implementation, the slip at time t, st (t ) , is calculated from the relative displacement of the bond interface element nodes connecting the concrete element and steel element. The time-independent slip si is back calculated from st via a relationship obtained from Eq. 3.74 by expressing si in terms of st (t ) as si = st (t ) (3.75) (3.74)

(1 + 240 t ) 0.08

The bond stress corresponding to si is then calculated using the time-independent bond stress-slip relationships given by Eq. 3.73. In a secant stiffness solution procedure, the secant modulus of the bond element is computed as Eb. sec = b st (t ) . An issue arose when the CEB-FIP bond model was implemented into the finite element code. The bond model has an infinitely large initial gradient (at zero slip), which inevitably causes computational difficulty. To prevent the numerical instability due to the extremely high initial bond stiffness, the initial modulus of the bond model is replaced by a linear modulus for any slip that is smaller than a threshold slip. The threshold slip should be a reasonably small value which does not introduce a very large gradient that may exceed the precision of the compiler. In this study a threshold slip of 10-7 mm is used with a Compaq FORTRAN 90 compiler and is found to work well for the purpose of this research (see Figure 3.13b).

115

Time-independent bond stress-slip curve

b
CEB-FIP bond model

Bond stress-slip curve at time t

Linear initial modulus

s (a)

s thres

s (b)

Fig. 3.13 - (a) Influence of creep on bond stress-slip curve; (b) replacing the infinitely large initial modulus by a linear initial modulus.

3.6.9

Concrete Tension Stiffening

For the distributed cracking model (see Section 3.3), the tension chord model discussed in Section 3.3.1 is utilized to model concrete tension stiffening in cracked reinforced concrete elements. Consider equilibrium across a reinforced concrete element in tension, the mean concrete tension stiffening stress ctsm can be calculated for any known mean and maximum steel stresses (Foster and Marti, 2003) using

ctsm = ( sr sm )

(1 )

(3.76)

where sr is the steel stress at the cracks (maximum steel stress) and sm is the mean steel stress in the reinforced concrete tension chord. For the case of a two-dimensional reinforced concrete membrane element with orthogonal reinforcement, the concrete tension stiffening stresses are calculated as per the cracked membrane model (see Figure 3.3). The maximum variation of steel stress is obtained by integrating the change of steel stress between the cracks (as given by Eq. 3.2) over one half of the crack spacing. Therefore sr can be determined by adding one half of the maximum variation of steel

116

stress to the mean steel stress, which leads to the following equations for various loading stages:

sr = E s m +

b0 s rm

for sr f sy
2 b0 E s E s s rm E + E b0 b1 w w b1 E 0.5 b0 s b1 E w

(3.77a)

sr = f sy +

b0 s rm s ( f sy E s m ) b0 rm

for s min f sy < sr f sy b1 s rm + sr = f sy + E w m Es for f sy f w f sy sr = f w + Eu m Es Ew for f sy < s min and f sy sm < f w b1 s rm + f sy < s min and f w sm f u

(3.77b)

(3.77c)

(3.77d)

where m is the average strain of the reinforced concrete element and s min is the minimum steel stress between two cracks of the tension chord. Figure 3.14 shows the distributions of bond, concrete and steel stresses between two cracks of a cracked reinforced concrete tension chord at the loading stages corresponding to Eq. 3.77a to Eq. 3.77d.

117

(a)
b0 b1

b
b0 b1

b
b0 b1

b
b0 b1

c
ctsm

c
ctsm

c
ctsm

c
ctsm

fu fw fsy

s
sr

sm s min B

fu fw fsy

s
sm s min sr

fu fw fsy

s
sm s min sr

fu fw fsy

s
sm s min sr

(b)

(c)

(d)

(e)

Fig. 3.14 - Distribution of bond stress ( b ) , concrete stress ( c ) and steel stress ( s ) between cracks: (a) uniaxial reinforced concrete member in tension; (b) Eq. 3.77a; (c) Eq. 3.77b; (d) Eq. 3.77c; (e) Eq. 3.77d.

3.7 Non-linear Finite Element Implementation


The numerical analysis of a reinforced concrete structure is complicated by the nonlinear material stress-strain relationships. This is further aggravated by the nucleation of cracks when the concrete tensile strength is violated, which results in a massive stress redistribution within the structure. The reinforced concrete modelling approaches and the material models presented in the previous sections are implemented using the displacement-based finite element method. In this section, the fundamentals of the finite element method are briefly described and the iterative non-linear solution procedures are presented.

118

3.7.1

Spatial Discretization

In a finite element analysis, a continuum is discretized into a discrete number of elements and the physical behaviour of the continuum is described by means of material models for the elements. The elements are interconnected at a discrete number of nodal points on the element boundaries. For a displacement-based finite element method, the continuous field displacement vector u at any point within an element is obtained by interpolating the nodal displacement vector u e of that element which is expressed as u = N u e (3.78)

where N is the displacement interpolation matrix containing the shape functions relating the continuous field to the nodal values. Knowing the displacements at any point within the element the strain at that point can be determined by using an appropriate linear operator, where the strain field can be written as = L u (3.79)

where is a vector of strains and the linear operator matrix L is also known as the differential operator matrix since is a differential function of u . In a displacement-based finite element analysis, the displacements are usually more readily available at the nodal level. Therefore it is more conveniently to express the strain field in terms of the nodal displacements. This is done by combining Eq. 3.78 and Eq. 3.79, which gives = B ue where B = L N and is known as the strain-displacement matrix of the element. With the strain known at a point, the stress can be calculated for a given material property. In a time-dependent analysis, it is important to correctly account for the development of time-dependent inelastic strains in a structure. In this study the time119

(3.80)

dependent inelastic strains are creep strain and shrinkage strain of concrete. Therefore, the stress in the element can be written as = D ( 0 ) (3.81)

where D is the material elasticity matrix and 0 is a pre-strain vector. For the case of concrete, 0 is the sum of creep and shrinkage strains.

3.7.2

Time Discretization

To trace the time-dependent behaviour of reinforced concrete structures due to the effects of creep and shrinkage in concrete, the time domain needs to be discretized into a number of finite time steps and the time-dependent analysis is performed based on a step-by-step integration through the time domain. Due to the aging nature of concrete, the rate of deformation of a concrete structure under a sustained load decreases with time. For an efficient and accurate time analysis, it is desirable to discretize the time domain such that smaller time intervals are used for deformations at higher rates and larger time intervals for deformations at lower rates. An effective time discretization should produce a nearly constant change in displacement over each time interval as depicted in Figure 3.15a. Gilbert (1979) suggested to discretize the time on a creep-time curve such that the same amount of specific creep occurs in each time interval (see Figure 3.15b). On the other hand, Baant (1979, 1988) recommended the use of about two to three constant time step intervals per unit logarithmic time scale is sufficient for an effective time analysis since creep curves are approximately linear on a logarithmic time scale. In this study the time discretization approach suggested by Baant (1979, 1988) is adopted. The rationale of this approach is essentially identical to that given by Gilbert (1979), as the objective is to keep the change in displacement over the time step as constant as possible so as to minimize the errors arise during the early stages after loading when the time-dependent development of strains are rapid.

120

Displacement, u
Equal displacement increments Equal creep strain increments

cp

t0 t1t2 t3

t4

t5

t6

t0 t1 t2 t3 t4

t5

t6

t7

(a)

(b)

Fig. 3.15 - Time discretization: (a) time steps with equal increments of displacement; (b) time discretization based on equally divided creep strains (Gilbert, 1979).

3.7.3

Principal of Virtual Work

According to the principal of virtual work, for any quasi-static and admissible virtual displacement u e which takes place relative to an equilibrium configuration, the internal work done by the stresses of a body must be equal to the external work done by the external forces acting on that body (Cook et al., 2001). Consider an element subjected to external forces such as body forces p b , surface tractions p s and concentrated nodal forces p e , the application of the principal of virtual work leads to

T dV =
e

u T p b dV +

A u
e

p s dA + u e T p e

(3.82)

where Ve is the volume of the element. The term on the left of the equal sign of Eq. 3.82 is the internal work done due to the element stresses. The external work done is given by the right side, which is due to the body forces, the surface tractions and the nodal forces.

121

By substituting Eq. 3.78, Eq. 3.80 and Eq. 3.81 into Eq. 3.82, we obtain

u e T B T D B dV u e u e T B T D 0 dV V Ve e
= u e T N T p b dV + V e

N T p s dA + p e e

(3.83)

Since Eq. 3.83 is valid for any virtual displacement, it can be rewritten as K e u e = Fe + Pe where K e = B T D B dV
Ve

(3.84)

and

Fe = B T D 0 dV
Ve

(3.85a, b)

and Pe = N T p b dV +
Ve

N T p s dA + p e

(3.85c)

in which K e is the element stiffness matrix, Fe is the nodal forces originate from the element pre-strains (for example creep and shrinkage strains) and Pe is the total external nodal forces applied to the element. It should be noted in Eq. 3.85b and Eq. 3.85c that the element pre-strains, the distributed body forces and surface tractions acting on the element continuous field are converted into sets of equivalent nodal forces via appropriate interpolations and integrations. In the present study, only the concentrated nodal forces and element pre-strain are considered. Hence, the timedependent deformations resulting from creep and shrinkage of concrete can be treated in the same way as applying a set of nodal forces to the element. The concept of applying the principal of virtual work to a single element can be extended to the global structural level by the full assembly of finite elements into a structural stiffness matrix. The equivalent equation to Eq. 3.84 at the structural level is Ku = F + P (3.86)

122

The displacement vector u contains the displacements corresponding to the degrees of freedom (dof) at each node in the structure. K, F and P are the structural stiffness matrix, structural equivalent pre-strain nodal force vector and the external structural nodal force vector, respectively, and are given by K = Ke ;
N

F = Fe
N

and

P = Pe
N

(3.87)

where N is the total number of elements used to model the structure. For a loadcontrolled structural analysis, F and P are known in advance and K can be assembled readily for a given set of material properties. The structural nodal displacements can be determined by solving the N dof number of equations (Eq. 3.86) given that the boundary conditions are properly specified so that the stiffness matrix K is nonsingular, meaning that rigid-body movements of the structure are prevented.

3.7.4

Incremental Iterative Solution Procedures

The equilibrium derivations for finite elements discussed previously is limited to a linear structural system for which the unknowns in the structural system can be determined directly by solving the equations. The equilibrium governing equation given by Eq. 3.86 can be written as P Q = 0 (3.88)

where Q is a vector of internal forces of the structure with a volume V which is given by Q = K u F = B T dV
V

(3.89)

Non-linear structural problems are not possible to be solved directly by the linear equilibrium conditions given by Eq. 3.88. The solution of non-linear systems require some iterative solution techniques that can solve the linear equilibrium equations iteratively until a desirable convergence tolerance is achieved. The most frequently used iterative solution procedure is the well-known Newton-Raphson (NR) method. In
123

this study the time-dependency of the non-linear structural system is treated as quasistatic, no time integration scheme is, therefore, involved in the solution procedures. The time-dependent, non-linear equations are solved in space and time to trace the evolution of stresses and strains. Consider a time-dependent non-linear structural system, the equilibrium state given by Eq. 3.88 cannot be attained. That is, the internal and external forces do not subtract to zero. In an iterative solution procedure at time t + t and at iteration j, the residual forces or the out-of-balance forces R can be written as
t + t

R (u j ) = t + t P t + t Q(u j )

(3.90)

Assuming a converged solution has been achieved for u j , a truncated Taylor series expansion and impose the equilibrium condition, that is, the out-of-balance forces must be zero. This gives
t + t

d R (u j ) R (u j +1 ) = t + t R (u j ) + u j = 0 du

(3.91)

where u j = u j +1 u j . By differentiating Eq. 3.90 with respect to u, the external load vector t + t P , which is independent of u, is eliminated and therefore gives d R (u j ) du = d Q(u j ) du = B T
V

d (u j ) du

dV

(3.92)

By taking an infinitesimal stress d , the relationship with an infinitesimal displacement d u can be written as d = D tan B d u where D tan is the material tangent constitutive matrix. (3.93)

124

The relationship given by Eq. 3.93 is used in Eq. 3.92 and is subsequently, together with Eq. 3.90, substituted into Eq. 3.91. This leads to the recurrence relationship of the NR method. That is K tan u j = t + t P t + t Q(u j ) j where K tan is the tangent stiffness matrix given by j K tan = B T D tan B dV j j
V

(3.94)

(3.95)

The

displacement

approximation

at

iteration

j +1

is

calculated

as

u j +1 = u j + u j . The incremental iterative process carries on until a convergence criterion is achieved. Figure 3.16a shows a schematic illustration of the NR method. In the NR method, the tangent stiffness matrix of the discretized structure is computed for each iteration. This inevitably incurs a high computational cost which is especially significant for large finite element analyses (with a large number of dof). Due to the fact that NR is an iterative approximation-correction approach, various variations of NR have emerged to reduce the computational efforts of assembling the updated tangent stiffness matrix for each iterative cycle. The frequently utilized variants of NR are the modified Newton-Raphson (mNR) method and the initial stiffness Newton-Raphson (iNR) method. In general, the mNR method involves the update of the tangent stiffness matrix occasionally during the iterative process. A more effective mNR algorithm is to update the tangent stiffness matrix upon each convergence of an equilibrium state. Modifying Eq. 3.94 gives K tan0 u j = t + t P t + t Q(u j ) j= (3.96)

where K tan0 is the tangent stiffness matrix obtained from the previously converged j= time step and is used throughout the current time step.

125

A secant stiffness approach is adopted in this study. Therefore, a secant stiffness matrix K sec is computed instead of the tangent counterpart. Therefore the secant NR can be expressed as K sec u j = t + t P t + t Q(u j ) j and the secant mNR as K sec0 u j = t + t P t + t Q(u j ) j= (3.98) (3.97)

where K sec is the secant stiffness matrix computed for each iterative cycle and K sec0 j j= is the secant stiffness matrix determined from the previously converged time step. The secant stiffness approach offers higher numerical stability than the tangent stiffness approach. This is particularly evident when a softening structure is analysed. A negative tangent stiffness may be encountered which can subsequently lead to numerical difficulties. As for the secant stiffness approach, the change of the secant stiffness is gradual and a positive stiffness is guaranteed. Nevertheless, the stability of a secant stiffness approach is compromised by a slower rate of convergence. Whilst an mNR method requires the update of tangent or secant stiffness matrix at any one of the equilibrium states, the iNR method only requires the formation of the initial tangent stiffness matrix, that is, the linear elastic stiffness matrix K 0 . This leads to a recurrence relationship as K 0 u j = t + t P t + t Q(u j ) (3.99)

Both mNR and iNR reduce the computational efforts per iterative cycle, however, the convenience is traded off by a lower rate of convergence compared to the regular NR method. In the authors view, the stability aspect of a non-linear analysis is far more important than the speed of the solution techniques, as the latter can always be overcome by the advancing computer technology while the former is purely numerical,

126

which depends on the iterative numerical algorithm. The iterative procedures of the tangent mNR, secant NR, secant mNR and iNR are shown schematically in Figure 3.16.

Load
t+t

Load
t+t tan Kj

P P

1 uj
t+t

P
tan K j=0

1 uj
t+t

t-t

Q(uj)

t-t

Q(uj)

uj Load
t+t

uj +1

(a)

Displacement Load
t+t

uj

uj +1

(b)

Displacement

P
sec Kj

1 uj
t+t

P K j=0
sec

1 uj
t+ t

t-t

Q(uj)

t-t

Q(uj)

uj uj +1

(c)
t+t

Displacement Load P P 1 K0 uj
t+ t

uj uj +1

(d)

Displacement

t-t

P K0 1 uj uj +1

Q(uj)

(e)

Displacement

Fig 3.16 - Non-linear iterative solution procedures: (a) tangent NR method; (b) tangent mNR method; (c) secant NR method; (d) secant mNR method; (e) iNR method.

127

3.7.5

Geometric Non-linearity

The non-linear solution techniques presented so far are limited to structures undergoing sufficiently small deformation so that the infinitesimal linear strain approximation remains valid. To account for non-linearity arising from large deformation, the second Piola-Kirchhoff stress and Green-Lagrange strain shall be considered instead of the frequently used engineering stress and strain. The Green-Lagrange strain is the exact strain measure for any size of deformation for which a quadratic term is present in its mathematical definition. If the quadratic term in the Green-Lagrange strain is omitted, the result is the engineering strain, which has a linear relationship with the change in length of the material and is only applicable when deformation is small. In addition, the second Piola-Kirchhoff stress is the stress measure for large deformation that is conjugate to the Green-Lagrange strain. However, the use of a complete different set of stress and strain measures inevitably requires a restructuring of the numerical procedures for the elements. To avoid this cumbersome process, an updated Lagrangian formulation (Bathe, 1996; Zienkiewicz and Taylor, 2000), in which the stress and strain are computed based on the last computed displaced configuration, is employed. By this way, the deformation with respect to the deformed shape of the last iterative step becomes arbitrarily small and hence the use of engineering stress and strain can be retained. Consequently, relatively few modifications are needed to extend the material non-linear iterative solution procedures to account for non-linearity arising from large deformation. The secant NR recurrence relationship given by Eq. 3.97 can now be written with reference to the last computed displaced configuration which has a volume V. That is K sec u j = t + t P j where K sec = B T D sec B dV j j
V'

V ' B

T t + t

j dV

(3.100)

(3.101)

128

where D sec is the material secant constitutive matrix evaluated at iteration j based on j the last computed configuration. The iterative process can be modified accordingly to adopt the mNR and iNR methods as discussed in Section 3.7.4. In the finite element implementation, the iterative solution procedures are identical to those for material non-linearity, but the nodal coordinates of the discritized structure are updated constantly in each iterative cycle. This can be done easily by adding the total nodal displacements calculated in the last iterative cycle to the undeformed nodal coordinates of the structure. Stress and strain are evaluated based on the updated nodal coordinates and the iterative process given by Eq. 3.100 and Eq. 3.101 continues until a set convergence tolerance is reached. It should be noted that, the geometric stiffness matrix (Zienkiewicz and Taylor, 2000), which is often found adding to the material stiffness matrix, is deliberately excluded from Eq. 3.101. Since equilibrium between internal and external forces of a structure is achieved iteratively in the solution procedure, the exact assembly of the stiffness matrix is therefore not required. Van Greunen (1979) employed the same treatment in his numerical study on reinforced concrete slabs and panels and proved success of the implementation.

3.7.6

Convergence Criteria

The incremental iterative solution techniques discussed above provide only approximate solutions to non-linear structural analyses. This postulates that the equilibrium conditions given by Eq. 3.88 can only be met in an approximate manner. Therefore convergence tests are required to ensure a sufficiently close equilibrium state is achieved. The test for convergence of the iterative cycles can be performed either on the norm of the out-of-balance forces or the norm of the displacement increments for each iteration. That is ri < r j max ( j = 1, 2, 3, ..., i ) (3.102)

129

where is the convergence tolerance, convergence criterion and criterion.

r j = R (u j ) T R (u j )

for a force

r j = u j T u j

for a displacement convergence

3.7.7

Computational Solution Algorithm

The mathematical formulations for non-linear finite element analyses presented above are implemented into a finite element code. The general algorithm for all element types based on the regular NR method is shown in the flowchart depicted in Figure 3.17. The finite element program is divided into three main modules, an input data control, an elements module and an equations solver. The input data control module collects the geometric data, material properties and configuration of the external forces. These data are stored temporarily in the computer memory and are used throughout the analysis. The elements module is responsible for the computation of structural stiffness matrix, element stresses and the out-of-balance forces. The gauss quadrature integration scheme is adopted for the integration of the element stiffness matrix (Eq. 3.85a). The structural stiffness matrix is obtained by summing up the element stiffness matrix over all elements as given by Eq. 3.87. For the first iterative cycle of the first time step, the nodal displacements of the discretized structure are zero and the time step is small, therefore the structural stiffness matrix is determined based on the linear elastic (initial) material constitutive matrix. At this iterative cycle, the internal forces vector of the discretized structure is a null vector and the out-of-balance forces (step 14 in flowchart) are directly taken as the total external forces of the current time step. After the out-of-balance forces vector is computed, it is transferred to the equations solver and the displacement increments for current iterative cycle (denoted as j in step 15 of the flowchart) is calculated by solving the structural equilibrium equations. The new displacements of the discretized structure are obtained by adding the displacement increments of the current iterative cycle to the displacements from the previous iterative cycle or from the previously converged time step.

130

In the step that follows, the convergence is checked if the required tolerance is reached, which can be based either on the displacement or force convergence criterion as discussed in Section 3.7.6. If the set tolerance is reached, the results of the current time step will be printed to a file. Depending on the analysis requirement, the analysis may proceed to the next time step with or without a further load increment, or the analysis may terminate. On the other hand, if the convergence criterion is not satisfied, a new set of stresses and strains of the elements are evaluated from the displacements calculated in step 15 of the flowchart and hence facilitates the computations of the structural stiffness matrix and the out-of-balance forces. The iterative process continues until the set tolerance is attained. The equivalent nodal forces due to pre-strains are not computed explicitly in the computational algorithm. Instead, the pre-strains are accounted for in step 9 of the flowchart before the stresses of the elements are evaluated. The internal forces computed in step 13 have indirectly included the equivalent pre-strain forces. In Appendix B, the treatment for pre-strains is illustrated with a simple hand calculation.

131

Read elements mesh data and boundary conditions

Read external nodal forces P0

Read material properties

Input data control

Increase time and external forces t = t + t ; P = k P0

Output stresses and strains then proceed to next time step or STOP program

17

Geometrically non-linear analysis?


Yes No

Determine material constitutive matrix D


10

Update nodal coordinates xi = xi + uxi yi = yi + uyi 6

Assemble structural stiffness matrix K = V BT D B dV

11

Determine strains of elements = Bu

Determine element stresses = D i


7

12

Determine pre-strains 0 of elements (e.g. creep and shrinkage strains)

Calculate structure internal forces Q = V BT dV

13

Determine instantaneous strains of elements i = - 0

Calculate out-of-balance forces R=PQ

14

Elements module

15

Tolerance not reached

Displacement or force convergence test

16

Equations solver

Fig. 3.17 - Finite element implementation flowchart.

132

Tolerance reached

Calculate structure nodal displacements u j+1 = u j + u j where u j = K-1 R

3.8 Finite Element Formulations


Two types of models are proposed in Sections 3.3 and 3.4 to simulate cracking of reinforced concrete structures. For the distributed crack model, reinforcement is modelled as an additional smeared stiffness over the concrete plane stress elements. While for the localized cracking model, concrete and reinforcing steel are modelled as individual components and the stresses between concrete and steel are transferred via bond action. Three types of element are used in the finite element analyses in this study. Plane stress elements are employed for representing reinforced concrete for the distributed cracking model and plain concrete for the localized cracking model. The discrete steel representation for the localized cracking model is modelled by means of truss elements and the bond action is realized by the use of interface elements.

3.8.1

Four-node Isoparametric Quadrilateral Element

A formulation that enables the definition of quadrilateral shapes that can have angles other than 90 at the corners of the elements is more versatile in generating meshes for structures of different shapes. The isoparametric elements formulation that is often used in structural analyses provides this particular facility. The main features of the isoparametric elements are the use of the natural coordinate system (, ) for numerical integration and the utilization of shape functions for the interpolations of coordinates and displacements. The general four-node isoparametric quadrilateral plane element adopted in this study is shown in Figure 3.18a. The nodal coordinates indicated in Figure 3.18a are based on the natural coordinate system where the origin is located at the centre of the element. In an isoparametric formulation, the term isoparametric suggests that the same set of shape functions is used for the interpolations of coordinates (x, y) and displacements (u, v) for any point within an element. The coordinates at a point are given by x = N i xi
i =1 4

and

y = N i yi
i =1

(3.103)

133

and the displacements are given by u = N i ui


i =1 4

and

v = N i vi
i =1

(3.104)

where i is the numbering of the nodal points shown in Figure 3.18a and N i is the shape function for node i and is given by N1 = 1 (1 )(1 ) ; 4 1 (1 + )(1 + ) 4 and N2 = 1 (1 + )(1 ) ; 4 1 (1 )(1 + ) 4 (3.105a, b)

N3 =

N4 =

(3.105c, d)

The four-node isoparametric quadrilateral element is also known as the bilinear quadrilateral element as the interpolated field quantities vary linearly with the twodimensional Cartesian coordinate system. The strains within an element can be calculated using the relationship given by Eq. 3.80. That is = B ue where = [ x (3.106)

y xy ]T

and

u e = [u1 v1 ... u 4

v 4 ]T .

The

strain-

displacement matrix B is given by B = [B1 B 2 where N i, x Bi = 0 N i, x 0 N i, y N i, y B3 B4 ] (3.107)

(i = 1, 2, 3, 4)

(3.108)

in which the notation N i , x denotes a partial derivative

N i . x

134

Before matrix B can be determined, the shape functions given by Eq. 3.105 are differentiated with respect to and . By applying the chain rule, we obtain N i , N i, x N = J N i , i, y where J is the Jacobian matrix given by x, J= x, y, y,

(3.109)

(3.110)

The derivatives of the shape functions with respect to the x and y can be obtained by inverting J. That is N i, x 1 y, N = i , y J x, y , N i , x, N i,

(3.111)

Having known N i, x and N i, y , matrix B can be computed. For the construction of the element stiffness matrix in global coordinates, the material constitutive stiffness matrix determined from the material coordinate system (D c12 ) as discussed in Section 3.5 is transformed such that D c = T T D c12 T (3.112)

where D c is the material matrix in the global coordinate system and T is the strain transformation matrix given by c2 s2 cs 2 c2 T = s cs 2cs 2cs c 2 s 2

(3.113)

135

with c = cos and s = sin and is the angle between the material coordinate system (directions of the principal strains) and the global coordinate system. The element stiffness matrix is obtained by applying the principle of virtual work, that is K e = t e B T D c B dA Ae (3.114)

where Ae and t e are the area and the thickness of the plane element, respectively. The integration of the element stiffness matrix is undertaken adopting a 2 2 gauss quadrature rule.
v4 4 v2 u3 v2
(1, -1)

u4

(-1, 1)

3 (1, 1)

v3

2
=1

u2

v1 1

u2 y

v1 1
=-1

ua

sa

(-1, -1)

u1

u1 x

(a)

(b)

Fig. 3.18 - (a) Four-node isoparametric quadratic element; (b) two-node truss element.

3.8.2

Two-node Truss Element

In this study reinforcing steel is assumed to carry axial forces only, therefore simple truss elements are sufficient to serve this purpose. The truss elements are overlaid onto the concrete plane element discussed in the preceding section. To ensure displacement continuity of the finite elements, a two-node linear truss element as shown in Figure 3.18b is adopted. By writing the nodal displacements as a nodal axial displacement (u a ) , we obtain u a.i = ui cos + vi sin (3.115)

136

where i (i = 1, 2) denotes the node number of the element and is the angle of orientation of the truss element from the x-axis. The linear shape functions for the element are given by N1 = 1 (1 ) 2 and N2 = 1 (1 + ) 2 (3.116a, b)

The axial displacement at any point of the element can be interpolated as u a = N i u a.i = [cos
i =1 2

0.5 [(1 )u + (1 + )u ] 1 2 sin ] 0.5 [(1 )v1 + (1 + )v 2 ]

(3.117)

Denoting the axial length as s a , the strain in the element is obtained by invoking the chain rule as

u a s a

(3.118)

The derivative of u a is obtained by differentiating Eq. 3.117 with respect to and gives u a 1 = [(u1 + u 2 ) cos + (v1 + v 2 ) sin ] 2 Also, given that the total length of the truss element is L, we obtain 2 = sa L Substitute Eq. 3.119 and Eq. 3.120 into Eq. 3.118 gives (3.120) (3.119)

= B ue
where the strain-displacement matrix B is given by B= 1 [ cos L sin cos sin ]

(3.121)

(3.122)

137

The material constitutive matrix for the case of a one-dimensional steel truss
sec element is a mere 1 1 matrix equal to the secant modulus E s of the trilinear stress-

strain model shown in Figure 3.12a. The element stiffness matrix is calculated as
sec K e = B T E s B dV Ve

(3.123)

Given that the total cross-sectional area of the reinforcing steel A is known and hence a relationship dV = 0.5 AL d , the element stiffness matrix can be obtained explicitly by integrating over the limit 1 1 as c 2 cs sec AE s s2 Ke = L sym. c2 cs c2 cs s2 cs s2

(3.124)

in which c = cos and s = sin .

3.8.3

Four-node Isoparametric Bond Interface Element

A four-node interface element (shown in Figure 3.19a) is used to model bond-slip between the reinforcing steel and the concrete. The relative displacement between node set 1 (consisting of nodes 1 and 4) and node set 2 (nodes 2 and 3) represents the slip between the concrete and the steel and is given by
+ uti = uti u ti

and

+ u ni = u ni u ni

(i = 1, 2)

(3.125)

where i is the node set number, the superscripts + 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 displacement field u tn by u tn.i are linked to the continuous

138

u tn = N i u tn.i = Tb B u e
i =1

(3.126)

where N i are the linear shape functions given by Eq. 3.116, u e is the nodal displacement in the global coordinate system and Tb is the bond transformation matrix given by cos Tb = sin sin cos (3.127)

where is the angle of the interface element to the global x-axis. The matrix B in Eq. 3.126 relates the continuous field relative displacements to the nodal displacements along the interface and is given by N B= 1 0 0 N1 N2 0 0 N2 N2 0 0 N2 N1 0 0 N1

(3.128)

The element stiffness matrix is computed by applying the principle of virtual work and gives Ke = B T D b B dA (3.129)

Ae

where Ae is the tangential contact surface area between the interface element and the adjacent materials and D b is the bond constitutive matrix. For the one dimensional bond formulation used in this study (see Figure 3.19b), Ae is the surface area of the reinforcing bar encased in the concrete and D b is given by E D b = bt 0 0 E bn

(3.130)

where Ebt is the bond stress-slip modulus for the current stress state and Ebn is the normal bond stress-split modulus. To maintain compatibility between the reinforcing

139

steel and the concrete in the normal direction, a stiff value for Ebn is used. In the tangential direction the CEB-FIP (1993) model is used to define the bond-stress versus slip relationship as discussed in Section 3.6.8 (also see Figure 3.12b). For stability of the solution process the bond stiffness Ebt is taken as the secant stiffness. For the four-node linear bond element, integration of Eq. 3.129 is undertaken explicitly giving 2 Ebt cL Ke = 6 0 2 Ebn Ebt 0 2 Ebt 0 Ebn 0 2 Ebn Ebt 0 2 Ebt 0 2 Ebt 0 Ebn 0 2 Ebn 0 2 Ebn 2 Ebt 0 Ebt 0 Ebt 0 2 Ebt 0 2 Ebn 0 Ebn (3.131) 0 Ebn 0 2 Ebn

sym.

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 K b in the global coordinate system is obtained by K b = TB T K e TB where TB = Tb Tb Tb Tb , which is a diagonal transformation matrix. (3.132)

+ + un2 ut2 3 u

n2

+ + un1 ut1

=1

ut2 2

Concrete element Interface element Steel element L

4 un1 ut1

1 =-1

zero width

(a)

(b)

Concrete element

Fig. 3.19 - Bond interface element: (a) four-node isoparametric interface element; (b) connectivity of interface element to concrete and steel elements.

140

CHAPTER 4 EVALUATION OF THE FINITE ELEMENT MODELS

4.1 Introduction
The numerical models described in Chapter 3 were incorporated into a finite element program RECAP (Foster and Gilbert, 1990; Foster, 1992) which was developed to model the behaviour of reinforced concrete structures. Prior to applying the finite element models to investigate the time-dependent behaviour of reinforced concrete structures, with particular interest in the formation of cracks at service load conditions, the reliability of the models must be assessed and evaluated. To facilitate this task, the numerical models are examined and calibrated using the experimental long-term tests undertaken by Gilbert and Nejadi (2004) and Nejadi and Gilbert (2004), which was a complementary long-term experimental program carried out in parallel with this numerical study. In this chapter, mesh dependency of the fracture models is firstly investigated with a plain concrete fracture test. A number of creep tests under varying stress histories are then modelled in order to evaluate the creep model based on the solidification theory (Baant and Prasannan, 1989a, b). The finite element models are also used to simulate the behaviour of beam and slab specimens and other timedependent tests on reinforced concrete structures. Comparisons are made between the numerical and experimental results.

4.2 Mesh Sensitivity of the Localized Cracking Models


An important aspect in finite element modelling for concrete structures is to ensure that the results of a numerical model are insensitive to different mesh configurations for a

141

specific numerical analysis. As discussed in Section 2.6.5, the numerical results for a lightly reinforced or unreinforced concrete structure can be highly dependent on the mesh size if no numerical treatment is employed to regularize spurious strain localization in the concrete (Baant and Pijaudier-Cabot, 1988; Baant and Jirsek, 2002; de Borst et al., 1993a). This is the major cause of mesh sensitivity problems associated with the modelling of reinforced concrete structures. The localized cracking models namely the crack band model and the non-local smeared crack model, are examined herein by analysing a four-point bending plain concrete fracture specimen using the meshes shown in Figure 4.1. The specimens spanned 450 mm, had an overall depth of 100 mm and were notched at midspan as shown in Figure 4.1. The vertically aligned meshes in Figures 4.1a and 4.1b were used to compare the mesh size dependency. The angled meshes in Figures 4.1c and 4.1d are oriented at 72 to the longitudinal direction and were used to test the influence of mesh orientation on crack propagation. The widths of the element above the notch for the coarse and fine meshes are 10 mm and 5 mm, respectively. The material parameters used in the four-point bending example were: Ec = 38 GPa , f ct = 3 MPa and

G f = 60 N/m . For the crack band model, fracture is expected to localize into a single band of elements for which the crack band widths were taken as the element width, i.e. either hc = 10 mm or hc = 5 mm . The characteristic length for the non-local smeared crack model was taken as lch = 50 mm which gives a width of localization of h = 50 mm approximately. Instead of adjusting the softening branch of tensile stress-strain curve according to the size of the element as in the crack band model, the tensile stressstrain curve for the non-local smeared crack model was obtained based on the width of localization (see Section 3.6.1). The load versus midspan deflection results for the analyses with various mesh sizes are shown in Figure 4.2. It is well recognized that the crack band model suffers from stress-locking (Rots, 1988) if the mesh orientation is not in alignment with the direction of crack propagation. This is seen in Figure 4.2a for both the coarse and fine slanted meshes, where the residual loads after cracking maintained at about 1.75 kN although the cracks had completely opened. Stress-locking is greatly reduced for the

142

Thickness = 50 mm 100 mm 450 mm (a) Nodes: 696 Elements: 642 (b) Nodes: 325 Elements: 283 (c) Nodes: 725 Elements: 672 (d) Fig. 4.1 - Finite element meshes: (a) coarse vertical mesh; (b) fine vertical mesh; (c) coarse slanted mesh; (d) fine slanted mesh. Nodes: 314 Elements: 273

5.0 4.5 4.0 3.5

Coarse vertical mesh Fine vertical mesh Coarse slanted mesh Fine slanted mesh

5.0 4.5 4.0 3.5

Coarse vertical mesh Fine vertical mesh Coarse slanted mesh Fine slanted mesh

Load (kN)

Load (kN)
0.3

3.0 2.5 2.0 1.5 1.0 0.5 0.0 0 0.1

3.0 2.5 2.0 1.5 1.0 0.5 0.0

0.2

0.1

0.2

0.3

Midspan deflection (mm)

(a)

Midspan deflection (mm)

(b)

Fig. 4.2 - Load versus midspan deflection diagrams: (a) crack band model; (b) nonlocal smeared crack model.
143

non-local model. The coarse slanted mesh is seen to exhibit a higher softening tail in Figure 4.2b, but when the mesh is refined the softening tail reduces to that of the vertical meshes. The deformed shapes and strain localization of the slanted meshes are shown in Figure 4.3 where the darker shading denotes higher strains. Although the crack band model suffers from stress-locking, it is surprising to observe that the crack propagation is not excessively sensitive to the mesh orientation as reported by Li and Zimmermann (1998). The crack path is corrected as the crack propagates through the specimen (Figure 4.3a). Improved crack path was obtained upon mesh refinement as shown in

(a)

(b)

(c)

(d)

Fig. 4.3 - Deformed meshes and strain localization (plot of principal strain): (a) and (b) coarse and fine slanted meshes for crack band model, respectively; (c) and (d) coarse and fine slanted meshes for non-local smeared crack model, respectively.

144

Figure 4.3b. For the non-local model, as expected, the width of the fracture zone stretches over a number of elements and propagates vertically through the specimen without being affected by the alignment of the mesh (Figures 4.3c and 4.3d). Li and Zimmermann (1998) analysed a notched three-point bending concrete fracture test with slanted meshes (see Figure 4.4) and reported that crack band models based on the rotating crack concept suffer from serious mesh directional bias. They investigated the influence of mesh orientation on crack propagation by setting the meshes at 65 and 45 to the longitudinal direction and the results are illustrated in Figures 4.5a and 4.5b. To investigate the crack band model developed in this study, identical specimens and meshes were analysed. The material parameters were: Ec = 20 GPa , f ct = 2.4 MPa and G f = 100 N/m . The thickness of the specimen has no influence on the direction of crack propagation but, of course, it does influence the

Thickness = 100 mm

200 mm 40 mm 100 mm 800 mm (a) 20 mm 100 mm

(b) Fig. 4.4 - Three-point bending concrete fracture specimen: (a) slanted mesh at 65 to the horizontal; (b) slanted mesh at 45 to the horizontal. (after Li and Zimmermann, 1998).

145

(a)

(b)

(c)

(d) Fig. 4.5 - Displaced shapes and crack patterns for three-point bending concrete fracture specimen: (a) and (b) crack patterns for 65 and 45 (to horizontal) slanted meshes, respectively (after Li and Zimmermann, 1998); (c) and (d) crack patterns for 65 and 45 (to horizontal) slanted meshes, respectively, analysed using the crack band model of this study.

load-deflection response. Therefore the thickness, which is not reported in Li and Zimmermanns work, was arbitrarily taken as one half of the depth of the specimen (i.e. 100 mm). The crack patterns from the analyses are shown in Figure 4.5c and 4.5d. It is obvious in Figure 4.5 that the rotating crack model developed by Li and Zimmermann (1998) exhibits more severe mesh directional bias than the model of this study. It is observed that, if the mesh is oriented at a smaller angle to the crack direction the directional bias is more prominent. This is shown in Figure 4.5c, where the crack

146

path is corrected after the crack has propagated over two elements. If the angle between the mesh and the crack direction is large, the crack tends to propagate diagonally through the elements as shown in Figure 4.5d. Though the crack path can be satisfactorily calculated the model still suffers from stress-locking as discussed earlier.

4.3 Creep of Plain Concrete under Variable Stress


To investigate the capability of the solidification formulation (Baant and Prasannan, 1989a, b) presented in Chapter 3, the creep tests under varying stress histories undertaken by Ross (1958) were analysed and were compared with the experimental results for verification. Ross (1958) loaded cylindrical specimens sized 305mm long and 117.5 mm diameter with various stress histories. Each specimen was accompanied by an unstressed control specimen under the same atmospheric conditions in order to measure the free shrinkage strain. The time-dependent deformation of the specimens was presented as the sum of elastic and creep strains. The cylindrical specimens cannot be modelled exactly using the plane stress elements developed in this study, however, since the specimens were loaded in the elastic range, the geometry of the specimens in fact has little influence to the development of strain. A single plane concrete element was used to analyse the uniaxially loaded creep specimens, as shown in Figure 4.6. The size of the element was arbitrarily taken to be 100 mm width, 200 mm long and a thickness of 65 mm. P/2 P/2

200 mm

100 mm Fig. 4.6 - A single plain concrete element used to model creep under variable stress.

147

The

material

parameters

used

for

the

creep

tests

were: and

E 0 = 1.6 E c.28 = 68.95 GPa ,

q 2 = 46.41 MPa ,

q3 = 11.6 MPa

q 4 = 3.046 MPa . The Dirichlet series was discretized into eight Kelvin chain units for storage of viscoelastic strain history. The corresponding elastic modulus E and retardation time for each Kelvin chain unit are given in Table 4.1 and the negative infinity area of the continuous retardation spectrum was taken as A0 = 13.2 MPa-1. The results calculated by the finite element model are compared with Ross (1958) creep tests in Figures 4.7, 4.8 and 4.9 with a good overall agreement. This demonstrates that the model based on the principle of superposition using the solidification theory computes generally accurate creep deformation of a concrete structure.

Table 4.1 - Kelvin chain properties for solidification creep model.

-th unit 1 2 3 4 5 6
7 8

E (MPa) 0.33892 0.28833 0.24817 0.21628 0.19098 0.17089 0.15494 0.14229

(days) 0.0001 0.001 0.01 0.1 1 10 100


1000

148

Stress (MPa)

20 15 10 5 0 0 700 20 40 60 80 100 120 140 160


15.03

Elastic + creep strains (x10 )

-6

600 Experimental 500 400 300 200 100 0 0 20 40 60 80 100 120 140 160 Age (days) FEM

(a)

Stress (MPa)

15 10 5 0 0 600

13.8 11 8.3 5.5 2.75

20

40

60

80

100

120

140

160

Elastic + creep strains (x10 )

500 400 300 200 100 0 0 20 40 60 80 100

-6

Experimental FEM

120

140

160

Age (days)

(b) Fig. 4.7 - Creep tests of Ross (1958): (a) specimen subjected to sustained constant load over a period of 46 days; (b) specimen subjected to descending stress history.
149

Stress (MPa)

15 10 5 0 0 600 20 40 60 80
5.5 2.75 11 8.3

13.8

100

120

140

160

Elastic + creep strains (x10 )

500 400 300 200 100 0 0 20 40 60 80 100 120 140 160 Age (days) Experimental FEM

-6

(a)

Stress (MPa)

15 10 5 0 0 600

13.8 8.3 2.75 8.3

13.8

20

40

60

80

100

120

140

160

Elastic + creep strains (x10 )

500 400 300 200 100 0 0 20 40 60 80 100 120 140 160 Age (days) Experimental FEM

-6

(b) Fig. 4.8 - Creep tests of Ross (1958): (a) specimen subjected to ascending stress history and a subsequent complete load removal; (b) specimen subjected to combined loading and unloading stress history.
150

Stress (MPa)

20 15 10 5 0 0 700 50 100 150 200 250 300


15.03 7.52

Elastic + creep strains (x10 )

-6

600 500 400 300 200 100 0 0 50 100 150 200 Age (days)

Experimental FEM

250

300

Fig. 4.9 - Creep tests of Ross (1958): specimen subjected to prolonged sustained stress.

4.4 Long-term Flexural Cracking Tests


4.4.1 Introduction

To the authors knowledge, experimental data available for long-term cracking of reinforced concrete structures are scarce in the literature. Two series of long-term tests were conducted by Gilbert and Nejadi, which consist of a series of flexural cracking tests (Gilbert and Nejadi, 2004) and a series of restrained deformation tests (Nejadi and Gilbert, 2004). One of the major aims of this complementary experimental program is to make available the experimental data with specific regard to the time-dependent development of crack spacing and crack width of reinforced concrete structures. The flexural cracking tests comprise a total of twelve one-way singly reinforced simply supported specimens without transverse reinforcement, in which six are beam specimens and the other six are slab specimens. The tests consist of six pairs of identical specimens and each pair was subjected to sustained load up to 50%

151

(designated a) and 30% (designated b) of its ultimate load (see Tables 4.2 and 4.3). The specimens were cast and moist cured for a period of 14 days and tested under sustained loading over a span of 3.5 m for durations up to 400 days. The beam specimens were subjected to sustained loads at the one-third points between the supports and the slab specimens were subjected to sustained uniformly distributed loads over the entire span. The growth of flexural cracking and midspan deflection for each specimen were monitored with time. The loading configurations and the dimensions of the specimens are shown in Figure 4.10. The details of the crosssection and the applied load for each specimen are presented in Tables 4.2 and 4.3. Companion specimens were also cast at the same time to measure creep and shrinkage of the concrete. The creep coefficients were calculated from the data obtained from a creep test which was performed by applying 5 MPa constant stress to several concrete cylinders. Shrinkage strains were measured from cylindrical specimens and two rectangular concrete panels after the commencement of drying after 14 days of age. Table 4.2 - Details of flexural beam specimens (Gilbert and Nejadi, 2004).
Specimen B1-a B1-b B2-a B2-b B3-a B3-b No. of bars 2 2 2 2 3 3 Bar dia. (mm) 16 16 16 16 16 16 cb (mm) 40 40 25 25 25 25 cs (mm) 40 40 25 25 25 25 s (mm) 150 150 180 180 90 90 Load, P (kN) 18.6 11.8 18.6 11.8 27.0 15.2

Table 4.3 - Details of flexural slab specimens (Gilbert and Nejadi, 2004).
Specimen S1-a S1-b S2-a S2-b S3-a S3-b No. of bars 2 2 3 3 4 4 Bar dia. (mm) 12 12 12 12 12 12 cb (mm) 25 25 25 25 25 25 cs (mm) 40 40 40 40 40 40 s (mm) 308 308 154 154 103 103 Load, w (kN/m) 2.9 1.9 4.9 2.9 5.8 3.9

152

A
L/3 L/3 L = 3500 L/3

(a)

B
L = 3500

(b)

400
300

s Ast cs

130 Ast

cs

(c)

(d)

Fig. 4.10 - Flexural cracking tests: (a) beam specimen under four-point sustained bending; (b) slab specimen subjected to sustained uniformly distributed load; (c) typical cross-section for beam specimens (section A-A); (d) typical cross-section for slab specimens (section B-B). (Gilbert and Nejadi, 2004).

153

4.4.2

Analysis of Long-term Flexural Cracking Tests and Material Properties

All the specimens in this series of tests were cast from the same batch of concrete mix. Material tests on 300 mm high by 150 mm diameter cylinder gave the mean 28-day concrete compressive strength as f cm = 24.8 MPa and the 28-day elastic modulus as Ec.28 = 24952 MPa . The tensile strength was obtained from indirect tension (Brazil) tests on 150 mm diameter cylinders tested at ages 14, 21 and 28 days and were f ct.14 = 2.0 MPa , f ct.21 = 2.6 MPa and f ct.28 = 2.8 MPa , respectively. For finite element modelling, the time-dependent development of tensile strength and shrinkage are approximated using the parameters A and B given by Eq. 3.46 in Chapter 3. The parameters for growth of the mean concrete tensile strength were taken as A f ct = 4.2 MPa and B f ct = 15 days and the growth curve is shown in Figure 4.11a. The shrinkage parameters were obtained by fitting the experimental shrinkage data and were Ash = 950 and Bsh = 45 days (see Figure 4.11b). The concrete fracture energy G f was taken as 75 N m and Poissons ratio = 0.2 .

Conc. tensile strength fct (MPa)

3.5 3.0 2.5 2.0 1.5 1.0 0.5 0.0 0

Shrinkage strain ( )

Afct = 4.2 MPa Bfct = 15 days

1000 800 600 400 200 0

Ash = 950 Bsh = 45 days

Experimental data Model 10 20 30

Experimental data Model 0 50 100 150 200 Time since commencement of drying (days)

Age (days)

(a)

(b)

Fig. 4.11 - Test data of companion specimens compared with models: (a) growth of concrete tensile strength; (b) shrinkage strain since commencement of drying. (Gilbert and Nejadi, 2004).

154

For creep modelling using the solidification theory, the asymptotic elastic modulus of concrete E0 is taken as that recommended by Bazant and Baweja (1995a), that is, E0 = 1.6 Ec.28 = 40 GPa . The empirical material constants q 2 , q3 and q 4 were determined by fitting the compliance data obtained from the creep test. The empirical material constants were q 2 = 182.5 MPa , q3 = 1.0 MPa and

q 4 = 23.7 MPa and the approximated compliance curve is shown in Figure 4.12. The Dirichlet series was discretized into eight Kelvin chain units for storage of deformation history of viscoelastic strain. The corresponding elastic modulus E and retardation time for each Kelvin chain unit are given in Table 4.4 and the negative infinity area of the continuous retardation spectrum was taken as A0 = 52.8 MPa-1. For the reinforcing steel, the stress-strain relationship is assumed to be elastic-perfectly plastic with a yield strength of 500 MPa and an elastic modulus of 200 GPa. In this numerical study, no correction has been made to account for size effects of creep and shrinkage of concrete. Considering the cross-section dimensions of the test specimens compared to the size of the control specimens for which the shrinkage and creep relationships with time were determined, this is considered to be a reasonable supposition.

200 Experimental data (t'=14 days) Model (t'=14 days)

J(t,t') (10 / MPa)

150

-6

100

50

0 0.1 1 10 100 1000

Time under load, t-t' (days)

Fig. 4.12 - Compliance curve for flexural cracking tests (Gilbert and Nejadi, 2004).

155

Table 4.4 - Kelvin chain properties for solidification creep model.


-th unit 1 2 3 4 5 6 7 8
E (MPa) 0.08480 0.07214 0.06209 0.05411 0.04778 0.04276 0.03877 0.03560

(days) 0.0001 0.001 0.01 0.1 1 10 100 1000

To compare the finite element models with the experimental results, the experimental average crack widths were calculated by summing the widths of all cracks at the soffit within the constant moment region for the beam specimens or within the region of moment greater than 90% of the maximum moment for the slab specimens. The sum of the crack widths was divided by the number of cracks within the region to give the average crack width. Serviceability of a reinforced concrete structure, however, depends not on the average crack width but on the widest crack. Therefore, the comparison of the experimental and numerical maximum crack widths is also a major interest in this study. Crack widths of the finite element models were computed using the methods discussed in Section 3.6.6. To illustrate the comparisons of the results of the finite element models and the experimental data, the detailed numerical results for four specimens (B1-a, B2-a, S2-a and S3-a) are presented and discussed in the sections that follow. The comparisons of results for the other 8 specimens are presented in tabulated form.

4.4.3

Analysis of Long-term Flexural Cracking Tests using the Distributed Cracking Model Cracked Membrane Model

4.4.3.1

Four-point Bending Beam Tests under Sustained Load

Two types of finite element mesh were used for the beams B1-a and B2-a since the two beams had different thicknesses of bottom concrete cover. Figure 4.13 shows the finite

156

element meshes defining the two different types of beam specimens. Due to symmetry, only one-half of each beam was required for the modelling. The mesh for the specimen with 40 mm cover consists of 199 nodes, 108 concrete elements, 54 reinforced concrete elements and 2 stiff elastic elements for the supports (Figure 4.13a). The mesh for the specimen with 25 mm cover is made up of 171 nodes, 108 concrete elements, 27 reinforced concrete elements and 2 stiff elastic elements at the supports (Figure 4.13b). The shaded mesh represents the concrete elements containing the smeared longitudinal reinforcement. The details of the reinforcement for all beam specimens are given in Table 4.5. Analyses were undertaken for = 0.5 and = 1 representing analysis based on the minimum and maximum crack spacings, respectively, which accordingly give the bounds of the crack opening with time. The numerical crack widths were obtained by summing the average crack width for all elements (the average crack width for all integration points within that element) at the soffit within the constant moment region and divided by the number of element under consideration.

(a)

(b) Plain concrete RC zone


Stiff elastic elements

Fig. 4.13 - Finite element meshes for beam specimens: (a) beam specimen with 40 mm concrete cover; (b) beam specimen with 25 mm concrete cover.

157

Table 4.5 - Reinforcement properties for beam specimens.


Specimen B1-a B1-b B2-a B2-b B3-a B3-b Reinforcement ratio

x 0.01676 0.01676 0.02437 0.02437 0.03655 0.03655

y 0 0 0 0 0 0

The results of the calculated time-dependent midspan deflection are plotted in Figure 4.14 and a good correlation is shown between the numerical results and the experimental data. The comparisons of the calculated and experimental crack widths in the constant moment region are shown in Figure 4.15. The calculated time-dependent crack opening is shown by the shaded region in Figure 4.15 and shows a reasonable agreement with the experimental results. However, the model underestimates the maximum crack width for beam B2-a. A second run of the finite element model was carried out using the crack spacings as observed in the experiment and the results were found to correlate well with the experimental average crack opening as shown in Figure 4.15. This indicates a strong dependence of the numerical crack width on the crack spacing for the distributed cracking model.

16 Midspan deflection (mm) 14 12 10 8 6 4 2 0 0

Beam B1-a
Midspan deflection (mm)

16 14 12 10 8 6 4 2 0 0

Beam B2-a

Experimental FEM ( =1.0) FEM ( =0.5) 100 200 Time (days) 300 400

Experimental FEM ( =1.0) FEM ( =0.5) 100 200 Time (days) 300 400

(a)

(b) FEM and experimental time-dependent midspan

Fig. 4.14 - Comparison

of

deflections: (a) beam B1-a; (b) beam B2-a.

158

0.45 0.40 Beam B1-a

maximum

0.40 0.35 Beam B2-a

maximum average

Crack width (mm)

Crack width (mm)

0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0 100 200 300 400
average

0.30 0.25 0.20 0.15 0.10 0.05 0.00 0 100 200

Experimental FEM (EXP Srm) FEM (TCM Srm)

Experimental FEM (EXP Srm) FEM (TCM Srm) 300 400

Time (days)

Time (days)

(a) beam B1-a; (b) beam B2-a.

(b)

Fig. 4.15 - Comparison of FEM and experimental time-dependent crack openings: (a)

The computed crack width envelopes at 200 days after loading are compared with the experimental results in Figures 4.16a and 4.16b for beam B1-a and beam B2-a, respectively. The crack widths for the primary cracks within the constant moment region were plotted against the depth of the specimens. It is seen that the calculated crack width envelopes have a good agreement with the test data.

300
Height from bottom (mm)

Height from bottom (mm)

Beam B1-a

Experimental FEM

300 250 200 150 100 50 0

Beam B2-a

Experimental FEM

250 200 150 100 50 0 0 0.1

0.2 0.3 Crack width (mm)

0.4

0.5

0.1

(a)

0.2 0.3 Crack width (mm)

0.4

(b)

Fig. 4.16 - Comparison of FEM and experimental crack width envelopes across the depth of specimens at 200 days: (a) beam B1-a; (b) beam B2-a.

159

4.4.3.2

Uniformly Loaded One-way Slabs under Sustained Load

The longitudinal layout of slab S2-a and slab S3-a (Figure 4.10b) are identical to that of the beam specimens. The slabs were subjected to uniformly sustained load up to a period of 400 days. Figure 4.17 shows the finite element mesh for the slabs with the smeared reinforcement depicted by the shaded area. Table 4.6 shows the details of the reinforcement for all slab specimens. Only one half of specimens were modelled due to symmetry. The mesh has 199 nodes, 108 concrete elements, 54 reinforced concrete elements and 2 extra stiff elastic elements for the steel plate support. The numerical crack widths were calculated similar to the procedure explained for the beam specimens but the averaging was undertaken for the region of moment greater than 90% of the maximum moment, as indicated in Figure 4.17.

> 0.9 Mmax Plain concrete


RC zone

Stiff elastic elements

Fig. 4.17 - Finite element mesh for slab specimens.

Table 4.6 - Reinforcement properties for slab specimens.


Specimen S1-a S1-b S2-a S2-b S3-a S3-b Reinforcement ratio

x 0.009121 0.009121 0.013681 0.013681 0.018242 0.018242

y 0 0 0 0 0 0

160

Like the beam specimens, the slabs were modelled for = 0.5 and = 1.0 so that the maximum and minimum crack widths can be calculated by the finite element model. The midspan deflection curves with time are shown in Figures 4.18a and 4.18b for slab S2-a and S3-a, respectively, and the model results show a good correlation with the test data. In Figure 4.19 the development of crack opening with time of the finite element model again shows a close agreement with the experimental cracking opening. The development of crack opening with time, which was computed using the experimentally observed average crack spacing as an input to the finite element analysis, is shown in Figure 4.19 denoted by dashed line. The numerical results closely agree with the experimental results for slab S3-a but are slightly higher than those for specimen S2-a.
35 Midspan deflection (mm) 30 25 20 15 10 5 0 0 100 200 Time (days) 300 400 Experimental FEM ( =1.0) FEM ( =0.5) 35 Midspan deflection (mm) 30 25 20 15 10 5 0 0 100 200 Time (days) 300 400 Experimental FEM ( =1.0) FEM ( =0.5)

Slab S2-a

Slab S3-a

(a)

(b)

Fig. 4.18 - Comparison

of

FEM

and

experimental

time-dependent

midspan

deflections: (a) beam S2-a; (b) beam S3-a.


0.30 Slab S2-a
maximum

0.30

Slab S3-a
maximum

Crack width (mm)

Crack width (mm)

0.25 0.20 0.15


average

0.25 0.20 0.15 0.10 0.05 0.00

0.10 0.05 0.00 0 100 200

Experimental FEM (EXP Srm) FEM (TCM Srm) 300 400

average

Experimental FEM (EXP Srm) FEM (TCM Srm) 200 300 400

100

Time (days)

(a)

Time (days)

(b)

Fig. 4.19 - Comparison of FEM and experimental time-dependent crack openings: (a) beam S2-a; (b) beam S3-a.

161

4.4.3.3

Discussion

In this study the average crack spacing were taken as two third of the maximum crack spacing as recommended by CEB-FIP Model Code 1990 (1993), which is equivalent to taking = 0.67 in the finite element model. For presentation purposes, the timedependent midspan deflections for all other flexural specimens were calculated using

= 0.67 and the analysis results are compared with the test data in Table 4.7. Table 4.8
summarizes the average crack spacing calculated by the model and those measured in the test. The average crack width was calculated, again, with = 0.67 and the crack width for the widest cracks with = 1.0. The calculated crack openings with time are presented in Tables 4.9 and 4.10 for comparison with test data. For the case of the distributed cracking model, the concrete tensile strength is taken as a constant for the entire structure, thus the concrete elements at the soffit begin to crack as soon as the cracking load is reached. This inevitably over-softens the structure and must be compensated by a tension-stiffening model. The cracked membrane model includes tension stiffening via the tension chord model, however, the concrete tension stiffening stress between cracks is assumed to have a linear distribution. This in fact idealizes the actual concrete stress distribution, which increases non-linearly at a decreasing rate from the crack to the point midway between adjacent cracks. A comparison of the linearized and the actual concrete stress distribution is illustrated in Figure 4.20. Obviously, the linearized stress distribution underestimates the concrete tension stiffening which may lead to an over-soft structural behaviour. For the beam specimens that were lightly loaded relative to the cracking load, especially for beam B1-b, the calculated midspan deflections are higher than the experimental results. In the authors view, this is attributed to the limitation discussed above.

162

c fct

Actual stress distribution Cracked membrane model

Fig. 4.20 - Comparison of the concrete stress distribution between cracks for cracked membrane model and that in real structure.

In the test, it may be the case that the instantaneous load was applied approaching the cracking load and cracks formed at the weakest locations (lowest concrete tensile strength) but the crack pattern was yet to develop fully. Secondary cracks might then begin to form with time due to the effects of creep and shrinkage and therefore the increase in deflection with time was gradual. This can be illustrated by examining the experimental crack spacings for the beam specimens loaded at 50% (designated a) and 30% (designated b) of the ultimate loads (see Table 4.8). The specimens subjected to a lower load had wider crack spacings than those subjected to a higher load. In the distributed cracking model, full cracking of the specimens occurred immediately after the cracking load was exceeded and this subsequently leads to a larger deformation. This also explains the less satisfactory correlation between the calculated and the experimental average crack spacings for beam B2-b. In summary, the overall correlation for the average crack spacings is reasonable except for beams B3-a and B3-b, the model calculated a lower average crack spacing than the test data.

163

Table 4.7 - Midspan deflections for = 0.67 at various times t (days) after loading (Distributed cracking model Cracked membrane model).
Specimens Instantaneous B1-b FEM Exp. FEM/Exp. FEM Exp. FEM/Exp. FEM Exp. FEM/Exp. FEM Exp. FEM/Exp. FEM Exp. FEM/Exp. FEM Exp. FEM/Exp. FEM Exp. FEM/Exp. FEM Exp. FEM/Exp. 1.84 1.98 0.93 1.89 2.06 0.92 5.13 5.81 0.88 1.64 1.97 0.83 6.08 7.14 0.85 2.16 2.72 0.79 3.48 4.43 0.79 4.61 5.04 0.91 Midspan deflections (mm) t = 50 t = 200 6.94 4.88 1.42 6.48 5.11 1.27 10.48 10.08 1.04 5.63 5.33 1.06 18.75 18.59 1.01 15.28 12.62 1.21 14.56 14.34 1.02 15.34 15.22 1.01 9.54 6.69 1.43 9.32 7.06 1.32 13.50 12.35 1.09 8.52 7.13 1.19 24.66 22.87 1.08 21.14 17.79 1.19 21.05 19.83 1.06 21.96 20.65 1.06 t = 380 10.33 7.44 1.39 10.19 7.88 1.29 14.62 13.32 1.10 9.61 7.90 1.22 26.66 25.12 1.06 23.06 19.91 1.16 23.14 21.93 1.06 24.26 22.90 1.06

B2-b

B3-a

B3-b

S1-a

S1-b

S2-b

S3-b

An issue arose when the finite element analysis was undertaken for slab S1-b, the most lightly reinforced concrete slab specimen subjected to a sustained load of 30% of its ultimate load. The calculated results show that the specimen was uncracked at instantaneous loading and this does not agree with the observation in the experiment. Consequently, a range of concrete tensile strengths was assumed in the finite element trial simulation in order to predict the gradual cracking in slab S1-b. The results for the trial simulation are shown in Figure 4.21 and it is seen that the results for f ct.14 = 1.6 MPa correlates the test data most satisfactorily. Therefore, this value of concrete tensile strength was used for the modelling of slab S1-b.

164

18 Midspan deflection (mm) 16 14 12 10 8 6 4 2 0 0

Slab S1-b

Experimental FEM (fct=2.0MPa) FEM (fct=1.9MPa) FEM (fct=1.8MPa) FEM (fct=1.7MPa) FEM (fct=1.6MPa)

10

20 30 Time (days)

40

50

Fig. 4.21 - Finite element trial simulation over a range of concrete tensile strength for specimen S1-b.

Table 4.8 - Comparison of FEM and experimental average crack spacings for flexural specimens (Distributed cracking model Cracked membrane model).
Specimens B1-a B1-b B2-a B2-b B3-a B3-b S1-a S1-b S2-a S2-b S3-a S3-b FEM (mm) 195 195 135 135 87 87 158 158 104 104 78 78 Experimental (mm) 190 220 170 320 160 170 130 130 120 110 110 130 FEM/Exp. 1.03 0.89 0.79 0.42 0.55 0.51 1.21 1.21 0.87 0.95 0.71 0.60

The underestimation of the numerical model for the cracking state of slab S1-b is probably because the applied load was close to the cracking load of the specimen. Since cracking is a gradual process, slab S1-b could have been mildly cracked when subjected to the applied load and the cracks remained fine. The finite element model may not be able to capture this sensitive behaviour properly. In addition, due to the random nature of cracking and the heterogeneous properties of concrete, there might be some locations

165

in the specimens where the concrete tensile strengths were much lower than the mean concrete tensile strength obtained from a laboratory test. Consequently, this could also be one of the causes resulting in the specimen cracking at a lower load. As mentioned earlier the crack widths obtained by the distributed cracking model have a strong dependence on the crack spacing. As expected, for specimen B1-b (beam that was loaded up to about the cracking moment) the calculated crack widths are higher than the test data (see Table 4.9) since the model overestimated the deformation of the specimen. One would expect the same to happen for specimen B2-b (the other

Table 4.9 - Crack widths for beam specimens at various times t (days) after loading (Distributed cracking model Cracked membrane model).
Beam specimens t=7 B1-b FEM
avg

Crack widths (mm) t = 50 t = 200 0.164 0.097 1.69 0.245 0.127 1.93 0.073 0.110 0.66 0.126 0.127 0.99 0.108 0.127 0.85 0.163 0.254 0.64 0.042 0.082 0.51 0.067 0.127 0.53 0.218 0.122 1.79 0.312 0.152 2.05 0.117 0.127 0.92 0.172 0.152 1.13 0.124 0.149 0.83 0.177 0.279 0.63 0.065 0.102 0.64 0.100 0.127 0.79

t = 380 0.234 0.137 1.71 0.332 0.178 1.87 0.130 0.152 0.86 0.185 0.178 1.04 0.132 0.184 0.72 0.187 0.279 0.67 0.073 0.112 0.65 0.110 0.127 0.87

0.078 0.046 1.70 0.149 0.076 1.96 0.015 0.042 0.36 0.023 0.076 0.30 0.074 0.066 1.12 0.106 0.102 1.04 0.015 0.030 0.50 0.024 0.051 0.47

B2-b

B3-a

B3-b

Exp.avg FEM/Exp.avg FEMmax Exp.max FEM/Exp.max FEMavg Exp.avg FEM/Exp.avg FEMmax Exp.max FEM/Exp.max FEMavg Exp.avg FEM/Exp.avg FEMmax Exp.max FEM/Exp.max FEMavg Exp.avg FEM/Exp.avg FEMmax Exp.max FEM/Exp.max

166

Table 4.10 - Crack widths for slab specimens at various times t (days) after loading (Distributed cracking model Cracked membrane model).
Slab specimens t=7 S1-a FEM Exp.avg FEM/Exp.avg FEMmax Exp.max FEM/Exp.max FEMavg Exp.avg FEM/Exp.avg FEMmax Exp.max FEM/Exp.max FEMavg Exp.avg FEM/Exp.avg FEMmax Exp.max FEM/Exp.max FEMavg Exp.avg FEM/Exp.avg FEMmax Exp.max FEM/Exp.max
avg

Crack widths (mm) t = 50 t = 200 0.223 0.130 1.72 0.299 0.203 1.47 0.151 0.078 1.94 0.240 0.127 1.89 0.083 0.092 0.90 0.140 0.127 1.10 0.065 0.094 0.69 0.101 0.127 0.80 0.267 0.155 1.72 0.365 0.254 1.44 0.211 0.105 2.01 0.305 0.178 1.71 0.124 0.117 1.06 0.188 0.152 1.24 0.092 0.124 0.74 0.138 0.178 0.78

t = 380 0.282 0.168 1.68 0.386 0.254 1.52 0.229 0.114 2.01 0.327 0.178 1.84 0.137 0.130 1.05 0.202 0.178 1.13 0.101 0.137 0.74 0.150 0.203 0.74

S1-b

S2-b

S3-b

0.158 0.066 2.39 0.206 0.127 1.62 0.059 0.044 1.34 0.141 0.102 1.38 0.033 0.058 0.57 0.062 0.076 0.82 0.035 0.043 0.81 0.054 0.076 0.71

beam that was loaded up to about the cracking moment) as the model also computed a larger deformation than the observed deformation. However, the crack widths have been compensated by the smaller crack spacing calculated by the model than that observed in the experiment, therefore the apparent good correlation obtained by the model, in the authors view, was a mere coincidence. The model also calculated larger crack widths for slabs S1-a and S1-b as shown in Table 4.10, which are due mainly to the overestimation of crack spacing and the slightly higher deformation calculated by the model. For other specimens the model results have a reasonable correlation with the test data.

167

To investigate the correlation between the calculated and the experimental crack widths, the test data were plotted against the corresponding computed results for 7 days, 50 days, 200 days and 380 days after loading in Figures 4.22a and 4.22b, for both average and maximum crack widths, respectively. Three reference lines were plotted on the same diagram to show the lines with ratio of experimental to calculated results r of 1.5, 1 and 1.5-1. The line r = 1 depicts a perfect correlation between the model and the test results while the lines denoted r = 1.5 and r = 1.5-1 represent a deviation of 50% of the calculated results from the test data. It is seen in Figures 4.22a and 4.22b that most calculated crack widths fall within 50% of the measured crack widths except for the data points of beam B1-a, slab S1-a and slab S1-b, which were mentioned previously in the comparisons of results. The distributed cracking cracked membrane model, on the whole, gives reasonable results as shown in the comparisons described above. Nevertheless, the strong dependence of the numerical results on the crack spacing is the biggest shortcoming of the distributed cracking model. This may be overcome to some extent by calibrating the bond shear stress b0 , which has a considerable influence on the calculation of crack spacing.

0.4 Exp. avg. crack width, wa.Exp (mm)


r = 1.5

0.5 Exp. max. crack width, wm.Exp (mm)


r=1 r = 1.5 r=1

0.3
r = 1.5 -1

0.4 0.3 0.2 0.1 r = wm.Exp / wm.FEM 0 0 0.1 0.2 0.3 0.4 0.5 FEM max. crack width, wm.FEM (mm)

r = 1.5

-1

0.2

0.1 r = wa.Exp / wa.FEM 0 0 0.1 0.2 0.3 0.4 FEM avg. crack width, wa.FEM (mm)

(a)

(b)

Fig. 4.22 - Correlation diagrams for distributed cracking model cracked membrane model: (a) average crack widths; (b) maximum crack widths.

168

4.4.4

Analysis of Long-term Flexural Cracking Tests using the Localized Cracking Model Crack Band Model

In the preceding section, analyses were performed using the distributing cracking approach, which can be conceived as an average model that gives reasonably good results in a global sense. In the localized cracking model using the crack band approach, the cracking mechanism in a reinforced concrete member is decomposed into finer individual mechanisms such as fracture of concrete and stress transfer between steel and concrete. To correctly capture the localized fracturing of concrete, a finer mesh is required in comparison to the distributed cracking model. The size of the mesh must be sufficiently fine to capture the strain localization at discrete locations with a finite crack spacing and also to allow for the bond stress to develop over the bond development length. To calculate crack widths, the cracked concrete element (at discrete locations) at the soffit is identified and the element crack width is calculated by averaging the crack widths computed at the integration points of that element. The average crack width calculated by the finite element model is determined by summing the crack widths of the cracked elements and is then divided by the total number of cracked elements. The treatment of a crack as localized fracture enables the localized cracking models to trace the locations and propagation of the crack. To improve the simulation of the random cracking phenomena in real structures, stochastic fluctuations of the concrete tensile strength are introduced. Clark and Spiers (1978) suggested that the first major crack to form at about 90% of the mean concrete tensile strength and the last major crack at about 110% of the mean concrete tensile strength. In the following examples a 10% random fluctuation of the mean concrete tensile strength f ct is assigned to the concrete elements. The random concrete tensile strengths were generated using the random number generator of the Compaq FORTRAN 90 compiler used in this study. Different sets of random numbers may be generated using different seed numbers. In the examples that follow the size of each finite element was taken as three times the maximum aggregate size as recommended by Baant and Oh (1983). The parameters of the bond model adopted were as per the recommendation of CEB169

FIP Model Code 1990 (1993) for good bond conditions for unconfined concrete (also see Table 3.1) and these were: s1 = s2 = 0.6 mm , s3 = 1.0 mm ,

max = 10.0 MPa , f = 1.5 MPa


k u = 100 MPa mm .

and the unloading modulus was taken as

4.4.4.1

Four-point Bending Beam Tests under Sustained Load

The typical finite element mesh used to model all the beam specimens is shown in Figure 4.23a (hc = 35 mm) . The bottom concrete cover of the mesh is adjusted accordingly for each beam specimen. In addition, a fine mesh (hc = 17.5 mm) and a slanted mesh (hc = 35 mm) were used to model beam B1-a, as shown in Figures 4.23b and 4.23c, respectively, in order to evaluate the mesh size and mesh directional

(a)

(b)

(c) Fig. 4.23 - Finite element meshes for beam specimens: (a) coarse mesh for all beam specimens; (b) fine mesh for beam B1-a; (c) slanted mesh for beam B1-a.

170

dependency of the crack band model when considering time effects and with reinforcing steel. The coarse mesh has 665 nodes, 540 concrete elements, 54 steel elements, 54 bond-slip elements and 4 stiff elastic support elements. The fine mesh consists of 2405 nodes, 2168 concrete elements, 108 steel elements, 108 bond-slip elements and 8 stiff elastic support elements. The slanted mesh is made up of 625 nodes, 510 concrete elements, 51 steel elements and 51 bond-slip elements and 2 stiff support elements. The steel elements were connected to the concrete element via bondslip interface elements. The far left end node of the steel bar element was rigidly connected to the concrete membrane node to simulate anchorage of the bar. In addition, to examine the significance of the bond-slip interface element, beam B1-a was also modelled using the coarse mesh but with the steel truss elements directly connected to the nodal points of the concrete elements. The calculated midspan deflection versus time curves are compared with the experimental results in Figure 4.24. The coarse, fine and slanted meshes gave very similar deformation with time for beam B1-a but with a very slight overestimation compared to the experimental results. They are, however, considered to have a good agreement with the experimental data. It is seen that the calculated time-dependent midspan deflection for beam B1-a with a perfect steel-concrete bond assumption is not dissimilar from those calculated using the bond-slip interface elements. Despite this, a realistic description of steel-concrete bond is crucial and the importance of the bondslip interface elements will be seen subsequently when it comes to determining the
14 Midspan deflection (mm) 12 10 8 6 4 2 0 0 100 200 Time (days) 300 400 Experimental FEM (coarse mesh) FEM (fine mesh) FEM (slanted mesh) FEM (perfect bond) 14 Midspan deflection (mm) 12 10 8 6 4 2 0 0 100 200 Time (days) 300 400 Experimental FEM

Beam B1-a

Beam B2-a

(a)

(b)

Fig. 4.24 - Comparison

of

FEM

and

experimental

time-dependent

midspan

deflections: (a) beam B1-a; (b) beam B2-a.

171

crack opening of the specimens. For beam B2-a the correlation of the calculated midspan deflection and the experimental results is excellent. The displaced shape of beam B1-a is shown in Figure 4.25a for which the crack openings are clearly visible at a number of discrete locations. The bond-slip between steel and concrete is evidenced by the dislocation of the overlapping nodes of the steel and concrete elements, as depicted in Figure 4.25b.

Crack locations

(a)

(b)

Fig. 4.25 - Cracking of numerical beam B1-a: (a) FEM deflection at 380 days (coarse mesh: scale 40); (b) dislocation of nodes due to bond slip.

Figures 4.26a to 4.26l show the crack formation with time for beam B1-a and compare the numerical results for the coarse and fine meshes at instantaneous loading and at 380 days. The numerically calculated slips increased with time due to the effect of bond creep at the interface of reinforcing steel and concrete. At instantaneous loading, the bond stresses away from the dominantly cracked region (approaching the support) were nearly zero as no significant bond-slip has taken place due the absence of cracks. However, bond stresses developed with time in this region (Figures 4.26g and 4.26h) as slippage between the concrete and the steel increased due to shrinkage of concrete and the subsequent restraint imposed by the steel. This can be well illustrated by the development of tensile stress in the concrete and increase of compressive stress in the steel near the support as indicated in Figures 4.26i to 4.26l. Furthermore, the simulation of fracture of reinforced concrete with the use of bond-slip interface

172

(a)

(b)

(c)
0.25 Slip (mm) 0.05 -0.05 -0.15 -0.25
Bond stress (MPa) 5.0 3.0 1.0 -1.0 -3.0 -5.0 3.0 2.0 1.0 0.0

(d)
0.25 Slip (mm)

0.15

Instantaneous

0.15 0.05 -0.05 -0.15 -0.25

Instantaneous

380 days

(e)

380 days

(f)

Instantaneous

Bond stress (MPa)

5.0 3.0 1.0 -1.0 -3.0 -5.0 3.0 2.0 1.0 0.0
300 200 100 0

Instantaneous

380 days

(g)
380 days Instantaneous

380 days

(h)
380 days Instantaneous

Conc. stress (MPa)

(i)
380 days

Conc. stress (MPa)

(j)
380 days

Steel stress (MPa)

200 100 0 -100 -200

Steel stress (MPa)

300

Instantaneous

-100 -200

Instantaneous

(k)

(l)

Fig. 4.26 - Formation of crack for beam B1-a: (a) and (b) instantaneous crack pattern (cracking strain plot) for coarse and fine meshes, respectively; (c) and (d) crack pattern (cracking strain plot) for coarse and fine meshes at 380 days, respectively; (e) and (f) longitudinal slip between steel and concrete; (g) and (h) longitudinal bond stress; (i) and (j) longitudinal stress of concrete adjacent to reinforcing steel; (k) and (l) longitudinal reinforcing steel stress.

173

elements accounts for the tension stiffening effect in a more realistic manner. This is evidenced by the development of concrete stresses between the cracks at the constant moment region as a result of bond action between the concrete and the steel (Figures 4.26i and 4.26j). Figures 4.27a and 4.27b show the crack patterns for the slanted mesh at instantaneous loading and at 380 days, respectively. The experimental crack pattern for beam B1-a at 380 days is shown in Figure 4.27g. It is seen that the computed crack patterns for the three meshes agree well with that observed in the test. For the slanted mesh, the cracks propagated in the direction of the mesh orientation. The mesh directional bias can be alleviated if a finer mesh is used, as demonstrated in Figure 4.3b for the fracture tests presented in Section 4.2. The computed crack patterns for beam B1-a assuming perfect bond between steel and concrete (without bond-slip elements) are shown in Figures 4.27c and 4.27d for instantaneous loading and for 380 days after loading, respectively. It is apparent that the crack pattern computed without using the bond-slip interface elements is unobjective. Although localized cracking was computed, the spacing of the cracks is evidently incorrect compared to the experimental crack pattern. In addition, no distinct crack opening was obtained at the soffit since cracking has smeared over the bottom concrete cover of the beam. Figure 4.28 compares the computed crack patterns to the experimental crack pattern for beam B2-a and, again, a good agreement is obtained. In fact there is a degree of confusion regarding the significance of the modelling of bond between concrete and reinforcing steel. It is commonly conceived in finite element modelling of reinforced concrete structures that, the influence of bond is not important if the primary interest is to obtain a monotonic load-deflection response of a reinforced concrete member (Darwin, 1993). Stevens et al. (1991) developed a finite element code based on an extended version of the modified compression field theory of Vecchio and Collins (1986) and modelled one of the simply supported reinforced concrete beams of Bresler and Scordelis (1963). They concluded that the global loaddeflection behaviour of a reinforced concrete member was not sensitive to bond-slip except for cases where bond failure was critical. Balakrishnan et al. (1988) pointed out

174

(a)

(b)

(c)
Conc. stress (MPa)

(d)
380 days Instantaneous
Steel stress (MPa) 300 200 100 0 -100 -200

3.0 2.0 1.0 0.0

380 days

Instantaneous

(e)

(f)

(g) Fig. 4.27 - Crack pattern of beam B1-a: (a) and (b) crack patterns (cracking strain plot) for slanted mesh at instantaneous loading and at 380 days, respectively; (c) and (d) crack patterns (cracking strain plot) for coarse mesh without bondslip interface elements (perfect bond) at instantaneous loading and at 380 days, respectively; (e) and (f) concrete stress and steel stress distribution for perfect bond example, respectively; (g) experimental crack pattern at 380 days.

175

(a)

(b)

(c) Fig. 4.28 - Crack pattern of beam B2-a: (a) and (b) crack patterns (cracking strain plot) at instantaneous loading and at 380 days, respectively; (c) experimental crack pattern at 380 days.

that the inclusion of bond-slip had a great effect on shear critical beams while had a very little improvement on the load-deflection response on other specimens. Based on the comments given above, one would naturally reach a conclusion to disregard bond action in a finite element analysis. However, if crack widths and crack spacings are of primary interest, the modelling of bond-slip is crucial. The bond action between steel and concrete can generally be divided into two major approaches as described in Section 2.6.7. Bond can be accounted for in an average sense in terms of concrete tension stiffening stress that develops due to stress transfer via bond or, it can be modelled in a discrete manner by using a specific type of bond element in conjunction with discrete steel elements. Both the finite element models developed by Stevens et al. (1991) and Balakrishnan et al. (1988), in fact, incorporated a concrete tension stiffening model, which infers that the effect of steelconcrete bond action was considered indirectly in their models. Therefore, it is incorrect to say that bond action is unimportant in simulating the global behaviour of a reinforced concrete structure if one is including a tension stiffening model that significantly affects that behaviour.

176

In addition, it is redundant to model discretely the bond-slip between concrete and reinforcing steel together with a tension stiffening model as done by Steven et al. (1991) and Balakrishnan et al. (1988) since this attempts to model the effect of bond twice. A realistic discrete bond model should be able to capture the tension stiffening effect through stress transfer between concrete and reinforcing steel (see Figures 4.26i and 4.26j) and no additional average tension stiffening stress should be superimposed onto the finite element model. Furthermore, a finite element model employing a tension stiffening model belongs to the distributed cracking models, for which cracking occurs in a smeared manner throughout the region where tensile stress is higher than the concrete tensile strength. In these models, the bond-slip element cannot work effectively since the slip between concrete and steel is small due to the smeared nature of cracking. Therefore, the discrete bond representation in modelling reinforced concrete member must be used in conjunction with a concrete fracture model as the discrete bond model can only be engaged effectively when the cracking of concrete is localized. The question remains, why the simulation for beam B1-a without using either the bond-slip interface elements or a tension stiffening model gives a good correlation with the experimental deflection as shown in Figure 4.24a? The reinforcing steel elements were rigidly connected to the nodes of the concrete elements. With the use of a concrete fracture model, the concrete elements adjacent to the steel elements were free to deform longitudinally through cracking and shearing along the reinforcing steel. This apparent bond-slip phenomenon assists the transfer of stress from the steel to the concrete and it can be observed in Figures 4.27e and 4.27f, where the distribution of concrete tension stiffening stress and the steel stress along the beam are shown. Although the apparent bond-slip cannot correctly simulate the bond mechanisms, it does induce a degree of concrete tension stiffening in the crack region which prevents the specimen from becoming over-soft in its structural behaviour. This also explains the good correlation between the calculated and the experimental deflections for the perfect bond example. Comparisons are made in Figures 4.29a, 4.29b and 4.29c for the three meshes between the calculated and measured crack widths at the soffit of beam B1a with increasing time in the constant moment region. The crack opening with time for beam

177

0.50

Crack width (mm)

0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0 100

Crack width (mm)

0.45 0.40

Beam B1-a

maximum

0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0

Beam B1-a

maximum

average

average

Experimental FEM (coarse mesh) 200 300 400

Experimental FEM (fine mesh) 100 200 300 400

Time (days)

(a)

Time (days)

(b)

0.45 0.40

Beam B1-a

maximum

0.40 0.35

Beam B2-a

maximum

Crack width (mm)

Crack width (mm)

0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0 100 200 300 400 Experimental FEM (slanted mesh)
average

0.30 0.25 0.20 0.15 0.10 0.05 0.00 0 100 200 300 400
average Experimental FEM (coarse mesh)

Time (days)

(c)

Time (days)

(d)

Fig. 4.29 - Comparison of FEM and experimental time-dependent crack openings: (a), (b) and (c) beam B1-a coarse mesh, fine mesh and slanted mesh, respectively; (d) beam B2-a coarse mesh.

B2-a is shown in Figure 4.29d. Both maximum and average crack widths are presented for comparisons between calculated and measured crack widths. It is seen that the calculated results are not very sensitive to the mesh configurations. An overall good correlation with the test results is achieved with a deviation less than 15%. The comparisons of the calculated and experimental crack envelopes for the primary cracks at 200 days after loading are shown in Figure 4.30a and 4.30b for beam B1-a and beam B2-a, respectively. Good agreement is observed for both specimens. The model also obtained accurate heights of crack penetration up into the beams, which approximately gives the position of the neutral axis.

178

300
Height from bottom (mm)

250 200 150 100 50 0 0 0.1

Height from bottom (mm)

Beam B1-a

Experimental FEM

300 250 200 150 100 50 0

Beam B2-a

Experimental FEM

0.2 0.3 Crack width (mm)

0.4

0.5

0.1

0.2 0.3 Crack width (mm)

0.4

(a)

(b)

Fig. 4.30 - Comparison of FEM (coarse mesh) and experimental crack width envelopes across the depth of specimens at 200 days: (a) beam B1-a; (b) beam B2-a.

In addition, from the crack patterns presented in Figures 4.26, 4.27 and 4.28, it is evident that the height of the crack penetration (or position of the neutral axis) has shifted to a lower position with increasing time. This agrees well with the fact that neutral axis of a beam moves downwards and the compressive zone gradually becomes larger due to the effect of creep of concrete with time (Gilbert, 1988).

4.4.4.2

Uniformly Loaded One-way Slabs under Sustained Load

The finite element mesh used to model the slab specimens is shown in Figure 4.31 with one half of the slab modelled due to symmetry. The finite element mesh consists of 390 nodes, 270 concrete elements, 54 steel elements, 54 bond-slip interface elements and 4 stiff elastic support elements. The steel and concrete were bonded via interface element with full anchorage provided for the steel node at the far left end of the specimen. The calculated midspan deflection with time curves are shown in Figures 4.32a and 4.32b for slab S2-a and slab S3-a, respectively, and both show an excellent correlation with the experimental data. The calculated crack patterns and cracking strain plots are shown in Figures 4.33a and 4.33b for slab S2-a and in Figures 4.34a and 4.34b for slab S3-a both at first loading and at 380 days after loading. Good agreements were obtained between the numerical results and the experimental crack patterns as presented in Figures 4.33c and 4.34c.
179

> 0.9 Mmax Fig. 4.31 - Finite element mesh for slab specimens.

40 Midspan deflection (mm) 35 30 25 20 15 10 5 0 0

Slab S2-a Midspan deflection (mm)

35 30 25 20 15 10 5 0

Slab S3-a

Experimental FEM 100 200 Time (days) 300 400

Experimental FEM 0 100 200 Time (days) 300 400

(a) Fig. 4.32 - Comparison of FEM and experimental deflections: (a) slab S2-a; (b) slab S3-a.

(b) time-dependent midspan

(a)

(b)

> 0.9 Mmax

(c) Fig. 4.33 - Crack pattern of slab S2-a: (a) and (b) crack patterns (cracking strain plot) at instantaneous loading and at 380 days, respectively; (c) experimental crack pattern at 380 days.

180

(a)

(b)

> 0.9 Mmax

(c) Fig. 4.34 - Crack pattern of slab S3-a: (a) and (b) crack patterns (cracking strain plot) at instantaneous loading and at 380 days, respectively; (c) experimental crack pattern at 380 days.

In Figures 4.35a and 4.35b the crack widths calculated by the finite element model for slab S2-a and slab S3-a are compared with the experimental measurement for the cracks within the region of moment greater than 90% of the midspan moment. The model calculated a slightly higher average crack width for both the slab specimens compared to the observed test results. An excellent correlation is obtained for the calculated and experimental maximum crack width for slab S2-a. For slab S3-a, the numerical maximum crack width development correlates well with the test data, however, the experimental crack width increased abruptly at about 280 days. Overall, the agreement of the numerical and experimental crack opening with time is reasonable.

0.30 0.25

Slab S2-a

maximum

0.30 0.25

Slab S3-a

maximum

Crack width (mm)

Crack width (mm)

0.20 0.15 0.10 0.05 0.00 0 100 200 300 400


average

0.20 0.15 0.10 0.05 0.00 0 100 200 300 400


average

Experimental FEM

Experimental FEM

Time (days)

Time (days)

(a) slab S2-a; (b) slab S3-a.


181

(b)

Fig. 4.35 - Comparison of FEM and experimental time-dependent crack openings: (a)

4.4.4.3

Discussion

Table 4.11 shows a summary of the comparison of the numerical and experimental midspan deflections at various times for specimens B1-b, B2-b, B3-a, B3-b, S1-a, S1-b, S2-b and S3-b. The numerical and experimental average crack spacings for the primary cracks are summarized in Table 4.12. The the time-dependent crack widths for the beam and slab specimens are given in Tables 4.13 and 4.14, respectively.

Table 4.11 - Midspan deflections at various times t (days) after loading (Localized cracking model Crack band model).
Specimens Instantaneous B1-b FEM Exp. FEM/Exp. FEM Exp. FEM/Exp. FEM Exp. FEM/Exp. FEM Exp. FEM/Exp. FEM Exp. FEM/Exp. FEM Exp. FEM/Exp. FEM Exp. FEM/Exp. FEM Exp. FEM/Exp. 1.93 1.98 0.97 2.09 2.06 1.01 6.05 5.81 1.04 1.79 1.97 0.91 8.23 7.14 1.15 2.46 2.72 0.90 3.27 4.43 0.74 6.11 5.04 1.21 Midspan deflections (mm) t = 50 t = 200 5.28 4.88 1.08 4.98 5.11 0.97 11.15 10.08 1.11 5.45 5.33 1.02 18.85 18.59 1.01 8.73 12.62 0.69 14.22 14.34 0.99 17.19 15.22 1.13 8.45 6.69 1.26 7.77 7.06 1.10 13.97 12.35 1.13 8.61 7.13 1.21 23.90 22.87 1.05 17.85 17.79 1.00 20.70 19.83 1.04 23.80 20.65 1.15 t = 380 9.77 7.44 1.31 8.76 7.88 1.11 15.03 13.32 1.13 9.61 7.90 1.22 26.78 25.12 1.07 21.28 19.91 1.07 23.39 21.93 1.07 26.37 22.90 1.15

B2-b

B3-a

B3-b

S1-a

S1-b

S2-b

S3-b

182

Table 4.12 - Comparison of FEM and experimental average crack spacings for flexural specimens (Localized cracking model Crack band model).
Specimens B1-b B2-b B3-a B3-b S1-a S1-b S2-b S3-b FEM (mm) 163 160 147 147 112 114 119 119 Experimental (mm) 220 320 160 170 130 130 110 130 FEM/Exp. 0.74 0.50 0.92 0.86 0.86 0.88 1.08 0.92

Table 4.13 - Crack widths for beam specimens at various times t (days) after loading (Localized cracking model Crack band model).
Beam specimens t=7 B1-b FEM
avg

Crack widths (mm) t = 50 t = 200 0.077 0.097 0.79 0.126 0.127 0.99 0.071 0.110 0.65 0.119 0.127 0.94 0.172 0.127 1.35 0.211 0.254 0.83 0.048 0.082 0.59 0.057 0.127 0.45 0.148 0.122 1.21 0.208 0.152 1.37 0.131 0.127 1.03 0.179 0.152 1.18 0.202 0.149 1.36 0.250 0.279 0.90 0.096 0.102 0.94 0.116 0.127 0.91

t = 380 0.172 0.137 1.26 0.223 0.178 1.25 0.148 0.152 0.97 0.191 0.178 1.07 0.209 0.184 1.14 0.264 0.279 0.95 0.109 0.112 0.97 0.130 0.127 1.02

0.027 0.046 0.59 0.040 0.076 0.53 0.018 0.042 0.43 0.031 0.076 0.41 0.135 0.066 2.05 0.163 0.102 1.60 0.027 0.030 0.90 0.032 0.051 0.63

B2-b

B3-a

B3-b

Exp. FEM/Exp.avg FEMmax Exp.max FEM/Exp.max FEMavg Exp.avg FEM/Exp.avg FEMmax Exp.max FEM/Exp.max FEMavg Exp.avg FEM/Exp.avg FEMmax Exp.max FEM/Exp.max FEMavg Exp.avg FEM/Exp.avg FEMmax Exp.max FEM/Exp.max

avg

183

Table 4.14 - Crack widths for slab specimens at various times t (days) after loading (Localized cracking model Crack band model).
Slab specimens t=7 S1-a FEM Exp.avg FEM/Exp.avg FEMmax Exp.max FEM/Exp.max FEMavg Exp.avg FEM/Exp.avg FEMmax Exp.max FEM/Exp.max FEMavg Exp.avg FEM/Exp.avg FEMmax Exp.max FEM/Exp.max FEMavg Exp.avg FEM/Exp.avg FEMmax Exp.max FEM/Exp.max
avg

Crack widths (mm) t = 50 t = 200 0.157 0.130 1.21 0.220 0.203 1.08 0.035 0.078 0.45 0.136 0.127 1.07 0.126 0.092 1.37 0.163 0.127 1.28 0.100 0.094 1.06 0.127 0.127 1.00 0.183 0.155 1.18 0.249 0.254 0.98 0.133 0.105 1.27 0.242 0.178 1.36 0.165 0.117 1.41 0.222 0.152 1.46 0.129 0.124 1.04 0.161 0.178 0.90

t = 380 0.192 0.168 1.14 0.254 0.254 1.00 0.154 0.114 1.35 0.250 0.178 1.40 0.177 0.130 1.36 0.234 0.178 1.31 0.140 0.137 1.02 0.173 0.203 0.85

S1-b

S2-b

S3-b

0.129 0.066 1.95 0.178 0.127 1.40 0.011 0.044 0.25 0.022 0.102 0.22 0.064 0.058 1.10 0.030 0.076 0.39 0.067 0.043 1.56 0.086 0.076 1.13

For the lightly loaded beam specimens, B1-b and B2-b, the distributed cracking model computed a generally higher midspan deflection than that measured from the experiment and the associated limitations were discussed in Section 4.4.3.3. In the localized cracking model, attempts were made to eliminate these limitations. The introduction of a random fluctuation of concrete tensile strength provides not only a closer simulation for a real structure but also facilitates a stable numerical analysis that prevents bifurcation stemming from the existence of multiple equilibrium paths. The coupling of a concrete fracture model with the bond-slip interface element enables a more realistic description of the tension stiffening effect of the cracked concrete and it also allows for the progressive formation of cracks as load increases.
184

For beam B1-b and beam B2-b, the midspan deflections calculated by this model have a closer agreement with the test results than those calculated by the distributed cracking model. Moreover, the issue discussed in Section 4.4.3.3 for slab S1-b still applies for the current model. Therefore, the concrete tensile strength at 14 days of age was taken as f ct.14 = 1.6 MPa in the analysis in order to obtain a pre-cracked specimen upon instantaneous loading as observed in the experiment. For slab S1-b, it is seen that the model underestimated the midspan deflection of the specimen at 50 days after loading (see Table 4.11), however the calculated results is again in good agreement from 200 days onwards. This is because the model computed less cracking before 50 days and further cracking occurred with time due mainly to the effect of shrinkage of concrete. In Table 4.12, it is seen that the computed average crack spacings have an overall satisfactory correlation with the test results except for beam B2-b, where the model computed a much lower average crack spacing than that observed in the experiment ( s rm. exp = 320 mm) . In the authors view, the experimental average crack spacing reported for beam B2-b (lightly loaded at 30% of the ultimate load with a bottom concrete cover of 25 mm) is not representative. Perhaps some fine cracks were overlooked during the experiments. Comparing with the experimental results of beam B1-b ( s rm. exp = 220 mm) which is also loaded at 30% of the ultimate load but with a thicker bottom concrete cover of 40 mm, the average crack spacing of beam B2-b is way too large. It is theoretically well established that a specimen with thicker concrete cover generally has a larger crack spacing, which clearly contradicts the overly large average crack spacing obtained experimentally for beam B2-b. It is seen in Tables 4.13 and 4.14 that the calculated crack widths are in reasonable agreement with the test results especially for the final crack widths, except for slab S1-b and slab S2-b, the model calculated a slightly larger crack width than the test results. Figures 4.36a and 4.36b show the correlation diagrams for the average crack widths and the maximum crack widths of all flexural specimens at 7 days, 50 days, 200 days and 380 days after loading. Similar to that for the distributed cracking model, the 0% (denoted by r = 1) and 50% (denoted by r = 1.5 and r = 1.5-1) deviation lines are plotted on the same diagram for comparison purposes. It is seen that the great majority of the data points are well within the 50% deviation lines. Comparing to the

185

distributed cracking model, the localized cracking crack band model obtained a better correlation with the test data as the data points are less scattered and also, are closer to the 0% deviation line.

0.4 Exp. avg. crack width, wa.Exp (mm)


r = 1.5

0.5 Exp. max. crack width, wm.Exp (mm)


r=1 r = 1.5 r=1

0.3
r = 1.5 -1

0.4 0.3 0.2 0.1 r = wm.Exp / wm.FEM 0 0 0.1 0.2 0.3 0.4 0.5 FEM max. crack width, wm.FEM (mm)

r = 1.5

-1

0.2

0.1 r = wa.Exp / wa.FEM 0 0 0.1 0.2 0.3 0.4 FEM avg. crack width, wa.FEM (mm)

(a) average crack widths; (b) maximum crack widths.

(b)

Fig. 4.36 - Correlation diagrams for localized cracking model crack band model: (a)

The localized cracking model with the crack band fracture model allows for a more detailed investigation of time-dependent cracking in reinforced concrete structures than the distributed cracking model, however, the numerical results of the model are still sensitive to some parameters, in particular the mean concrete tensile strength. Although a stochastic fluctuation of concrete tensile strength is used in the model, the mean concrete tensile strength is still the primary governing factor to the cracking load of a reinforced concrete structure. An accurate computation of the cracking load in turn influences the time-dependent behaviour of the structure, as was discussed in Section 4.4.3.3. The bond stress-slip model is also a factor that has a great effect on the computed crack widths. In this study, the bond model of the CEB-FIP Model Code 1990 (1993) for good bond conditions for unconfined concrete (Table 3.1) was adopted in the analyses for all specimens. If a different bond model or the other bond conditions of the CEB-FIP bond model were used in the analyses, a different computation outcome

186

would be expected. The computed crack widths would be narrower if a stiffer bond model was used or, they would be wider if a less stiff bond model were employed. It has been shown in Section 4.2 regarding mesh sensitivity, the crack band model suffers from stress locking if the orientation of the mesh is not aligned with the direction of crack propagation. Surprisingly, the stress-locking issue is not significant when modelling a reinforced concrete member. As seen in Figure 4.24, the midspan deflection with time diagram for beam B1-a using the slanted mesh in fact is very similar to those of the vertical meshes. In addition, it is also encouraging to obtain a numerical time-dependent crack opening for the slanted mesh having a very good correlation with the test data. It can, therefore, be concluded that stress-locking is merely a plain fracture problem and it becomes insignificant as soon as cracking of concrete is stabilized by the presence of reinforcement.

4.4.5

Analysis of Long-term Flexural Cracking Tests using the Localized Cracking Model Non-local Smeared Crack Model

It was seen in Section 4.2 that the non-local smeared crack model is more advantageous than the crack band model in modelling fracture of concrete. The model completely alleviates the mesh sensitivity problem usually associated with a standard smeared crack model. In addition, the non-local smeared crack model also overcomes the stresslocking phenomenon that persists in the crack band model. In this section, the non-local smeared crack model is employed in conjunction with the bond-slip interface element and steel truss element to simulate time-dependent cracking in reinforced concrete structures. An evaluation is made to scrutinize the applicability of the non-local model in reinforced concrete structures. As in the localized cracking model using the crack band approach, the concrete tensile strength assigned to each element making up a structure is randomized at a 10% limit of the mean concrete tensile strength f ct . For the following study, the characteristic length for the non-local smeared crack model is taken as lch = 50 mm and a width of localization of h = 50 mm is used to calculate the tensile stress-strain

187

curve for the concrete. The CEB-FIP (1993) bond stress-slip relationship for good bond conditions and unconfined concrete is used, giving s1 = s2 = 0.6 mm , s3 = 1.0 mm ,

max = 10.0 MPa ,

f = 1.5 MPa

and

an

unloading

modulus

k u = 100 MPa mm .

4.4.5.1

Four-point Bending Beam Tests under Sustained Load

The effectiveness of spatial averaging of the non-local smeared crack model depends primarily on the total number of integration points contained within the averaging neighbourhood of the point of interest in a structure. This inevitably calls for the need of a finer finite element mesh discretization. The finite element meshes used for the four-point bending specimens with bottom cover 40 mm and 25 mm are shown respectively in Figures 4.37a and 4.37c, for which the mesh in the constant moment region is 10 mm wide. A coarse mesh, as shown in Figure 4.37b, was also generated for beam B1-a to test the mesh size sensitivity. The mesh shown in Figure 4.37a consists of 2826 nodes, 2576 concrete elements, 112 steel truss elements, 112 bond-slip interface elements and 2 stiff elastic support elements. The coarse mesh shown in Figure 4.37b consists of 2176 nodes, 1978 concrete elements, 86 steel truss elements, 86 bond-slip interface elements and 2 stiff elastic support elements. The mesh, for specimens with 25 mm concrete cover (Figure 4.37c), is made up of 2713 nodes, 2464 concrete elements, 112 steel truss elements, 112 bond-slip elements and 2 stiff elastic elements. The steel truss elements were overlaid onto the concrete elements via bond-slip interface elements. A rigid fixity was provided at the node of steel-concrete overlay at the free end of the specimens near the support to simulate full anchorage. The experimental and numerical midspan deflections with time after loading are compared in Figures 4.38a and 4.38b for beams B1-a and B2-a, respectively. The finite element model is in reasonable agreement with the test data but slightly overestimates the deflections. In addition, the model is relatively insensitive to mesh size, which can be seen in Figure 4.38a where the deflection curves for the coarse and fine meshes are very similar.

188

(a)

(b)

(c) Fig. 4.37 - Finite element meshes for beam specimens: (a) fine mesh for beam specimens with 40 mm bottom cover; (b) coarse mesh for beam B1-a; (c) fine mesh for beam specimens with 25 mm bottom cover.

16 Midspan deflection (mm) 14 12 10 8 6 4 2 0 0

Beam B1-a Midspan deflection (mm)

16 14 12 10 8 6 4 2 0 0

Beam B2-a

Experimental FEM (fine mesh) FEM (coarse mesh) 100 200 Time (days) 300 400

Experimental FEM 100 200 Time (days) 300 400

(a) Fig. 4.38 - Comparison of FEM and experimental deflections: (a) beam B1-a; (b) beam B2-a.

(b) time-dependent midspan

189

Instead of presenting the crack pattern in a cracking strain plot as for the crack band model, the crack pattern computed by the non-local smeared crack model is shown as principal strain contours. Figures 4.39a to 4.39l show the crack patterns, the longitudinal slips and the stresses for beam B1-a. The results computed for the coarse and fine meshes are also compared. The model obtained a similar crack pattern for the coarse and fine meshes and the computed crack patterns compare favourably with the experimental crack pattern as shown previously in Figure 4.27g. The finite element computed crack patterns for beam B2-a are presented in Figure 4.40 and are similar to that observed in the test (Figure 4.28c). It is observed that the non-local smeared crack model is insensitive to mesh size and the widths of the fracture zones are consistent, irrespective of the size of the mesh. This is a prominent characteristic of the non-local model. However, due to the effect of spatial averaging the model is particularly sensitive to regions of high tensile stress. This can be seen in Figure 4.39a to 4.39d where cracking of concrete occurred not only in the transverse direction but also in the longitudinal direction at the level of the reinforcing steel as a result of the splitting tension induced by the bond action between steel and concrete. The sensitivity of spatial averaging in the model also results in a discontinuity in crack propagation from the soffit across the reinforcement layer causing fracture of concrete not being able to be localized nicely into a single crack. Unlike the crack band model, fracture in the non-local smeared crack model localizes into a number of element widths and this leads to a small slip between the steel and concrete. Comparing the magnitude of slip in Figures 4.39e and 4.39f to that in Figures 4.26e 4.26f, the non-local smeared crack model predicted only about 10% of that computed by the crack band model. The small slip gives rise to a low bond stress and this influences the distribution of the concrete tension stiffening stress and the steel stress. In addition, due to the influence of the dispersed fracture at the soffit, the distribution of concrete tensile stiffening stresses are also in a rather irregular pattern.

190

(a)

(b)

(c)
0.03 0.02 0.01 0 -0.01 -0.02 -0.03

(d)
0.03 0.02 0.01 0 -0.01 -0.02 -0.03 3.0 2.0 1.0 0.0 -1.0 -2.0 -3.0

Slip (mm)

Instantaneous

Slip (mm)

380 days

380 days

Instantaneous

(e)

(f)
Instantaneous

3.0 2.0 1.0 0.0 -1.0 -2.0 -3.0

Bond stress (MPa)

Instantaneous 380 days

Bond stress (MPa)

(g)
380 days Instantaneous

380 days

(h)
380 days Instantaneous

Conc. stress (MPa)

3.0 2.0 1.0 0.0

Conc. stress (MPa)

4.0

4.0 3.0 2.0 1.0 0.0


300 200 100 0

(i)
380 days

(j)
380 days

Steel stress (MPa)

200 100 0 -100 -200

Steel stress (MPa)

300

Instantaneous

-100 -200

Instantaneous

(k)

(l)

Fig. 4.39 - Formation of crack for beam B1-a: (a) and (b) instantaneous crack pattern (principal strain contours) for coarse and fine meshes, respectively; (c) and (d) crack pattern (principal strain contours) for coarse and fine meshes at 380 days, respectively; (e) and (f) longitudinal slip between steel and concrete; (g) and (h) longitudinal bond stress; (i) and (j) longitudinal stress of concrete adjacent to reinforcing steel; (k) and (l) longitudinal reinforcing steel stress.
191

(a)

(b)

Fig. 4.40 - Crack pattern of beam B2-a: (a) and (b) crack patterns (principal strain contours) at instantaneous loading and at 380 days, respectively.

The variation of the numerical crack width with time within the constant moment region for beam B1-a are compared to the experimental observations in Figures 4.41a and 4.41b for the coarse and fine meshes, respectively. Figure 4.41c shows the

0.50 0.45

Beam B1-a

maximum

Crack width (mm)

Crack width (mm)

0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0 100 200 300 400
average

0.50 0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0

Beam B1-a

maximum

average

Experimental FEM (coarse mesh)

Experimental FEM (fine mesh) 100 200 300 400

Time (days)

Time (days)

(a)
0.45 0.40 Beam B2-a
maximum

(b)

Crack width (mm)

0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0 100 200 300 400
average

Experimental FEM

Time (days)

(c) Fig. 4.41 - Comparison of FEM and experimental time-dependent crack openings: (a) beam B1-a coarse mesh; (b) beam B1-a fine mesh; (c) beam B2-a fine mesh.

192

comparison of the numerical and experimental time-dependent crack widths for beam B2-a within the constant moment region. It is seen that the model calculated a generally higher instantaneous crack width than those observed in the tests. On the whole, the model results have an overall good agreement with the test data. The crack width envelopes across the depth of beams calculated by the model for the constant moment region at 200 days are compared with the test data in Figures 4.42a and 4.42b for beams B1-a and B2-a, respectively, and show a satisfactory correlation with the test data.

300
Height from bottom (mm)

250 200 150 100 50 0 0 0.1

Height from bottom (mm)

Beam B1-a

Experimental FEM

300 250 200 150 100 50 0

Beam B2-a

Experimental FEM

0.2 0.3 Crack width (mm)

0.4

0.5

0.1

(a)

0.2 0.3 Crack width (mm)

0.4

(b)

Fig. 4.42 - Comparison of FEM (fine mesh) and experimental crack width envelopes across the depth of specimens at 200 days: (a) beam B1-a; (b) beam B2-a.

4.4.5.2

Uniformly Loaded One-way Slabs under Sustained Load

For the uniformly loaded slab specimens, cracking occurred in a distributed manner throughout the span of the specimens. To capture properly cracking of the specimens, fine meshes were used throughout the span of the slabs, as shown in Figure 4.43. A total of 2091 nodes were used to define the mesh. The mesh consists of 1730 concrete elements, 173 steel truss elements, 173 bond-slip interface elements and 4 stiff elastic support elements. As for all the localized cracking examples, steel truss elements were overlaid onto concrete elements via bond-slip interface elements and a full anchorage was provided to the steel element node at the free end of the slab.

193

> 0.9 Mmax

Fig. 4.43 - Finite element mesh for slab specimens.

The midspan deflection versus time diagrams for slabs S2-a and S3-a are presented in Figures 4.44a and 4.44b, respectively. In comparison to the test data, the model obtained results that are in a very close agreement with the experimental results over the duration of the test. The crack patterns are shown in Figures 4.45a and 4.45b for slab S2-a and in Figures 4.45c and 4.45d for slab S3-a at instantaneous loading and at 380 days after first loading. Although strain localization still tends to occur at locations of high bond stresses, it is seen that the strain localizations for the slab specimens have better defined individual cracks than those in the beam specimens (see Figures 4.39a to 4.39d and Figures 4.40a and 4.40b) as the propagation of cracks is continuous from the soffit through the reinforcement layer. The crack patterns computed by the model for the two slabs are similar to those observed in the test as shown in Figures 4.33c and 4.34c for slab S2-a and slab S3-a, respectively. The average

40 Midspan deflection (mm) 35 30 25 20 15 10 5 0 0

Slab S2-a Midspan deflection (mm)

35 30 25 20 15 10 5 0

Slab S3-a

Experimental FEM 100 200 Time (days) 300 400

Experimental FEM 0 100 200 Time (days) 300 400

(a) Fig. 4.44 - Comparison of FEM and experimental deflections: (a) slab S2-a; (b) slab S3-a.

(b) time-dependent midspan

194

(a)

(b)

(c)

(d)

Fig. 4.45 - Crack patterns of slab specimens (principal strain contours): (a) and (b) slab S2-a at instantaneous loading and at 380 days, respectively; (c) and (d) slab S3-a at instantaneous loading and at 380 days, respectively. crack spacings were calculated for slab S2-a and slab S3-a within the region greater than 90% of the maximum moment and are 110 mm and 92 mm, respectively, which correlate quite well with the measured average crack spacings given in Table 4.8 (120 mm for slab S2-a and 110 mm for slab S3-a). The calculated crack widths with time for slab S2-a and slab S3-a are compared with the test data in Figures 4.46a and 4.46b, respectively. The model results have a good correlation with the experimental results for the widest crack of both slabs and for the average crack width of slab S3-a. However, the average crack width calculated by the model for slab S2-a is higher than the experimental results. Similar to the results for the beam specimens, the model consistently calculated a high instantaneous crack width for the two slab specimens.
0.35 0.30 Slab S2-a
maximum

0.30

Slab S3-a

maximum

Crack width (mm)

0.25 0.20 0.15 0.10 0.05 0.00 0 100 200 300 400
average Experimental

Crack width (mm)

0.25 0.20
average

0.15 0.10 0.05 0.00 0 100 200 300 400 Experimental FEM

FEM

Time (days)

(a)

Time (days)

(b)

Fig. 4.46 - Comparison of FEM and experimental time-dependent crack openings: (a) slab S2-a; (b) slab S3-a.
195

4.4.5.3

Discussion

The comparisons of the computed and the test results for the other specimens are summarized in Table 4.15 for time-dependent midspan deflection, Table 4.16 for average crack spacing and Tables 4.17 and 4.18 for crack widths with time for beam specimens and slab specimens, respectively.

Table 4.15 - Midspan deflections at various times t (days) after loading (Localized cracking model Non-local smeared crack model).
Specimens Instantaneous B1-b FEM Exp. FEM/Exp. FEM Exp. FEM/Exp. FEM Exp. FEM/Exp. FEM Exp. FEM/Exp. FEM Exp. FEM/Exp. FEM Exp. FEM/Exp. FEM Exp. FEM/Exp. FEM Exp. FEM/Exp. 1.96 1.98 0.99 2.31 2.06 1.12 6.24 5.81 1.07 2.22 1.97 1.13 10.32 7.14 1.45 3.49 2.72 1.28 4.87 4.43 1.10 4.90 5.04 0.97 Midspan deflections (mm) t = 50 t = 200 4.48 4.88 0.92 5.18 5.11 1.01 11.32 10.08 1.12 6.39 5.33 1.20 19.40 18.59 1.04 12.30 12.62 0.97 14.08 14.34 0.98 15.12 15.22 0.99 7.05 6.69 1.05 7.34 7.06 1.04 14.41 12.35 1.17 8.76 7.13 1.23 24.56 22.87 1.07 16.72 17.79 0.94 19.27 19.83 0.97 21.23 20.65 1.03 t = 380 8.76 7.44 1.18 8.60 7.88 1.09 15.88 13.32 1.19 9.86 7.90 1.25 26.93 25.12 1.07 18.72 19.91 0.94 21.62 21.93 0.99 23.94 22.90 1.05

B2-b

B3-a

B3-b

S1-a

S1-b

S2-b

S3-b

196

Table 4.16 - Comparison of FEM and experimental average crack spacings for flexural specimens (Localized cracking model Non-local smeared crack model).
Specimens B1-b B2-b B3-a B3-b S1-a S1-b S2-b S3-b FEM (mm) 170 156 153 124 108 106 106 105 Experimental (mm) 220 320 160 170 130 130 110 130 FEM/Exp. 0.77 0.49 0.96 0.73 0.83 0.82 0.96 0.81

Table 4.17 - Crack widths for beam specimens at various times t (days) after loading (Localized cracking model Non-local smeared crack model).
Beam specimens t=7 B1-b FEMavg Exp. FEM/Exp.avg FEMmax Exp.max FEM/Exp.max FEMavg Exp.avg FEM/Exp.avg FEMmax Exp.max FEM/Exp.max FEMavg Exp.avg FEM/Exp.avg FEMmax Exp.max FEM/Exp.max FEMavg Exp.avg FEM/Exp.avg FEMmax Exp.max FEM/Exp.max
avg

Crack widths (mm) t = 50 t = 200 0.043 0.097 0.44 0.054 0.127 0.43 0.053 0.110 0.48 0.072 0.127 0.57 0.141 0.127 1.11 0.261 0.254 1.03 0.049 0.082 0.60 0.080 0.127 0.63 0.100 0.122 0.82 0.132 0.152 0.87 0.067 0.127 0.53 0.096 0.152 0.63 0.155 0.149 1.04 0.289 0.279 1.04 0.057 0.102 0.56 0.094 0.127 0.74

t = 380 0.158 0.137 1.15 0.221 0.178 1.24 0.094 0.152 0.62 0.151 0.178 0.85 0.162 0.184 0.88 0.304 0.279 1.09 0.067 0.112 0.60 0.109 0.127 0.86

0.023 0.046 0.50 0.031 0.076 0.41 0.043 0.042 1.02 0.052 0.076 0.68 0.132 0.066 2.00 0.236 0.102 2.31 0.044 0.030 1.47 0.069 0.051 1.35

B2-b

B3-a

B3-b

197

Table 4.18 - Crack widths for slab specimens at various times t (days) after loading (Localized cracking model Non-local smeared crack model).
Slab specimens t=7 S1-a FEM Exp.avg FEM/Exp.avg FEMmax Exp.max FEM/Exp.max FEMavg Exp.avg FEM/Exp.avg FEMmax Exp.max FEM/Exp.max FEMavg Exp.avg FEM/Exp.avg FEMmax Exp.max FEM/Exp.max FEMavg Exp.avg FEM/Exp.avg FEMmax Exp.max FEM/Exp.max
avg

Crack widths (mm) t = 50 t = 200 0.181 0.130 1.39 0.264 0.203 1.30 0.059 0.078 0.76 0.098 0.127 0.77 0.072 0.092 0.78 0.137 0.127 1.08 0.067 0.094 0.71 0.119 0.127 0.94 0.207 0.155 1.34 0.296 0.254 1.17 0.073 0.105 0.70 0.125 0.178 0.70 0.089 0.117 0.76 0.171 0.152 1.13 0.087 0.124 0.70 0.155 0.178 0.87

t = 380 0.221 0.168 1.32 0.313 0.254 1.23 0.084 0.114 0.74 0.143 0.178 0.80 0.104 0.130 0.80 0.196 0.178 1.10 0.101 0.137 0.74 0.181 0.203 0.89

S1-b

S2-b

S3-b

0.125 0.066 1.89 0.199 0.127 1.57 0.020 0.044 0.45 0.028 0.102 0.27 0.034 0.058 0.59 0.066 0.076 0.87 0.029 0.043 0.67 0.054 0.076 0.71

In Table 4.15, it is seen that the midspan deflections calculated by the non-local smeared crack model for the specimens have an overall good correlation with the measured results. In comparison with the distributed cracking model and the localized cracking model with the crack band approach, the non-local smeared crack model obtained results in better agreement with the experimental midspan deflections. This is especially noticeable for beam specimens subjected to 30% of the ultimate load (beams designated b). The distributed cracking model and the localized cracking crack band model obtained a final deflection higher than that observed by up to 40% and 30%, respectively, while the localized cracking non-local smeared crack model predicted deflections higher than the test data by only up to 25%. In addition, similar to the

198

previous two finite element models, the concrete tensile strength for slab S1-b at age of loading (age 14 days) was taken as f ct.14 = 1.6 MPa in order to obtain a cracked the specimen at instantaneous loading as observed in the experiment. The midspan deflection calculated using this concrete tensile strength is in reasonable agreement with the experimental results. Table 4.16 summarizes the calculated and experimental average crack spacing and the comparison shows a reasonable agreement. All calculated crack spacings are within 30% of the test results except for beam B2-b, where the calculated result is only half of that observed in the test. A discussion has been given in Section 4.4.4.3 regarding the experimental average crack spacing for beam B2-b. For the same specimen, the non-local smeared crack model computed an average crack spacing that is in close agreement with that obtained by the crack band model ( s rm = 160 mm) , which indicates that, although a different fracture model is used the modelling approach would give a similar result. In the comparisons of the time-dependent crack widths in Tables 4.17 and 4.18, it is seen that the final crack widths calculated by the model have a reasonable correlation with the experimental results. Nevertheless, the correlation for the calculated and observed development of crack widths over the test period is merely acceptable. As presented before for specimens B1-a, B2-a, S2-a and S3-a, the crack width calculated by the model at first loading is significantly larger than that observed. For the specimens subjected to 50% of the ultimate load (beam B3-a and slab S1-a) as shown in Tables 4.17 and 4.18, the same trend of crack opening with time is observed in the computed results. The instantaneous crack opening is larger than the test data, but is corrected as time approaches 380 days. For the lightly loaded specimens (designated b) except beam B3-b, a different trend in numerical crack width is observed. The model calculated a gradual increase in crack opening with time, but with an initial underestimation and a closer agreement for the final crack widths. It is noticed that the crack patterns computed by the model at low load levels are less well defined. This can be seen, for example, in the principal strain contours diagrams of beam B1-b as shown in Figure 4.47. Numerical analysis shows that the beam was cracked at instantaneous

199

(a)

(b)

Fig. 4.47 - Crack pattern of beam B1-b: (a) and (b) crack patterns (principal strain contours) at instantaneous loading and at 380 days, respectively.

loading, however, no obvious crack pattern is observed in Figure 4.47a. A stabilized crack pattern was only established with time. The crack pattern at 380 days is shown in Figure 4.47b. In the authors view, a number of factors are contributing to the discrepancy of the agreement between the calculated and experimental time-dependent crack openings. The nature of the modelling of fracture zones in the non-local smeared crack model leads to a few undesirable issues in the application to reinforced concrete structures. The non-uniform distribution of cracking strains across the width of the fracture zone causes difficulties in the determination of the crack width. The method used in this study to calculate the crack width is that presented in Section 3.6.6.3, which is simply to compute the difference in displacements of the two nodes adjacent to the fracture zone. However, this method only gives an approximate crack width in the fracture zone. In addition, the issue of the discontinuous fracture across the reinforcement layer from the soffit, which has been mentioned earlier, further handicaps the determination of crack width. Therefore, the computation of crack width depends not only on the magnitude of time-dependent deformation of concrete but also on the selection of the nodal points adjacent to the fracture zone (or crack). This undoubtedly leads to an additional variable in the determination of crack width. Another influencing factor is the effectiveness of the combined usage of the bond-slip interface element together with the non-local smeared crack model. It has been mentioned earlier that the bond slip at cracks is particularly low due to the distribution of fracture over a number of elements, as illustrated in Figure 4.39. It is questionable whether the behaviour of bondslip is realistically described in such a condition, which subsequently has a great influence on the distribution of steel and concrete stresses in the tension region.
200

The correlation diagrams for the localized cracking non-local smeared crack model are shown in Figures 4.48a and 4.48b for the average crack widths and the maximum crack widths at 7 days, 50 days, 200 days and 380 days after loading. It is seen that the model obtained a better correlation with the test data for data points of larger crack widths, which are made up mostly by data points with a longer period under load. On the whole, the model results are in a reasonable agreement with the experimental results.

0.4 Exp. avg. crack width, wa.Exp (mm)


r = 1.5

0.5 Exp. max. crack width, wm.Exp (mm)


r=1 r = 1.5 r=1

0.3
r = 1.5 -1

0.4 0.3 0.2 0.1 r = wm.Exp / wm.FEM 0 0 0.1 0.2 0.3 0.4 0.5 FEM max. crack width, wm.FEM (mm)

r = 1.5

-1

0.2

0.1 r = wa.Exp / wa.FEM 0 0 0.1 0.2 0.3 0.4 FEM avg. crack width, wa.FEM (mm)

(a)

(b)

Fig. 4.48 - Correlation diagrams for localized cracking model non-local smeared crack model: (a) average crack widths; (b) maximum crack widths.

4.4.6

Summary for Analysis of Long-term Flexural Cracking Tests

Comparing the correlation diagrams of the three finite element models (Figures 4.22, 4.36 and 4.48), it is obvious that the localized cracking crack band model has the best correlation with the test data. The data points obtained from this model, for both average and maximum crack widths, are consistently close to the line of 0% deviation. For the localized cracking non-local smeared crack model, the results correlate well with the test data for data points of large crack widths, while for data points of small crack widths, the model calculated lower crack widths than the test data. The crack widths calculated by the distributed cracking cracked membrane model have a uniform dispersion between the 50% deviation lines and have a constant tendency of overestimating the crack widths.
201

In fact, all three models have their respective merits and drawbacks. The distributed cracking model is the simplest model and requires the least computational resources as the model simulates reinforced concrete structures in an average manner. Therefore a coarse finite element mesh is usually sufficient to model most types of structural problems. Another advantage of the model is the smeared description of reinforcement, which greatly simplifies the mesh generation process. However, the shortcoming of the model is that, it only provides average results for a reinforced concrete structure, such as average stress and average strain, where the determination of a localized mechanism of the structure is not possible. Moreover, the reliability of the computed crack openings of a structure depends strongly on the accuracy of the calculated crack spacing obtained using the tension chord model. The original motives of a non-local model are to provide a realistic physical description of the fracture process and to regularize mesh sensitivity of tension softening materials such as concrete. Despite having a sound mathematical formulation that can effectively prevent both mesh size and mesh directional sensitivity, the computational cost is inevitably expensive due to the spatial averaging involved in the computational algorithm. In addition, a fine finite element mesh with mesh size smaller than the fracture width is required for an accurate non-local modelling of a structure. Because of the effect of spatial averaging, the width of a fracture zone generally stretches over several elements. This reduces the practical applicability of the non-local model to engineering problems since concrete structures normally contain a large amount of reinforcement and cracking are densely distributed due to the stabilizing effect of the reinforcement. If the fracture zone is wider than the crack spacing of a structure, the computed neighbourhood fracture zones may coalesce and lead to erroneous numerical results. In addition, the presence of a finite fracture width also reduces the effectiveness of the bond-slip interface element, as has been shown previously in the numerical examples. This further complicates the process of determining the crack openings of a reinforced concrete structure. In contrast, the crack band model offers several advantages over the non-local smeared crack model. The crack band model requires only an energy-based adjusted tensile stress-strain curve and the rest of the computational algorithm is exactly identical to that of a standard finite element model. To properly capture the localized
202

cracking in a reinforced concrete structure, the crack band model inevitably requires a finer finite element discretization than the distributed cracking model. Although some researchers (for example, de Borst, 1997) have pointed out that severe convergence instability may occur if mesh size for the crack band model is refined, the author has not encountered any such problems during the numerical studies undertaken in this study. One very encouraging finding of this study relating to the crack band model is that, the stress locking issue prevailing in concrete fracture problems appears to be insignificant in the modelling of reinforced concrete structures. Furthermore, comparing to the non-local smeared crack model, the crack band model works more effectively with the bond-slip interface elements in describing the bond-slip phenomenon, which in turn facilitates a more realistic stress transfer between concrete and reinforcing steel.

4.5 Long-term Restrained Deformation Cracking Tests


4.5.1 Introduction

This experimental program is the second series of long-term tests performed by Nejadi and Gilbert (2004). Eight fully restrained reinforced concrete slab specimens were fabricated and monitored up to a period of 130 days and the growth of cracking caused by the development of tensile stresses in the restrained specimens due to the effects of drying shrinkage was recorded. Each specimen was 600 mm wide, 100 mm deep and 2000 mm long and each end of the specimen was rigidly restrained by a 1000 mm by 1000 mm by 600 mm thick concrete block clamping to the reaction floor with high strength alloy steel rods so as to prevent longitudinal movement. The setup of the experiment is shown in Figure 4.49. The details of the cross-sections of each specimen are given in Table 4.19. A letter R is added to the original designation given by Nejadi and Gilbert (2004) for each specimen to differentiate the restrained slab specimens from the flexural specimens. The specimens were cured in the formwork and

203

Table 4.19 - Details of restrained slab specimens (Nejadi and Gilbert, 2004).
Specimen RS1-a RS1-b RS2-a RS2-b RS3-a RS3-b RS4-a RS4-b No. of bars 3 3 3 3 2 2 4 4 Bar diameter (mm) 12 12 10 10 10 10 10 10 cs (mm) 109 109 110 110 145 145 115 115 s (mm) 185 185 185 185 300 300 120 120

A
1000

A
2000 1000 2660 1000

(a)

600

(b)
cs

Metal sheet to induce first crack


75 600 75

Fig 4.49 - Restrained deformation cracking tests: (a) typical cross-section through mid-depth of the specimen; (b) side view; (c) typical cross-section (section A-A). (Nejadi and Gilbert, 2004).

204

kept moist by a covering of wet hessian for 3 days in order to prevent loss of moisture from the specimens. Drying shrinkage commenced after the removal of the wet hessian. Companion specimens were cast from the same batch of concrete for measuring the creep and shrinkage properties of the concrete. The creep coefficients were obtained from a creep test by loading a 5 MPa sustained stress longitudinally to concrete cylinders mounted in a standard creep rig at age 3 days. The development of shrinkage strain was measured from plain concrete specimens with the same cross-section as the restrained slab specimens and subjected to the same environmental, curing and drying conditions.

4.5.2

Analysis of Restrained Deformation Cracking Tests and Material Properties

The crack band model with bond-slip interface element was used to simulate these tests. For modelling purposes, it was assumed that the end concrete blocks supporting the slab was rigidly clamped to the floor and the junction connecting the slab to the end concrete block was fully restrained therefore no movement was possible. A typical finite element mesh for the tests is shown in Figure 4.50 (specimen RS1-a or b shown). The mesh consists of 1366 nodes, 1040 concrete elements, 247 steel truss elements and 247 bond-slip interface elements. The steel truss elements were connected to the concrete elements via bond-slip interface elements. The numbers of node and element differ slightly for each specimen since they contained different amount of reinforcement. Instead of exploiting the double symmetric nature for the simulation of only a quarter of the slab, the entire slab was modelled with a 10 percent stochastic fluctuation of the mean concrete tensile strength in order to capture the random formation of cracks in the reinforced concrete member. The boundary conditions were introduced such that the slabs were only free to shrink laterally. In addition, the symmetry line in the longitudinal direction of the slabs was restrained from deforming laterally by providing longitudinal roller supports to the concrete element nodes on the symmetry line. The introduction of this boundary condition is to prevent excessive

205

Softened element to induce first crack

Fig. 4.50 - Typical finite element mesh for restrained slab specimens (slab RS1-a/b shown).

in-plane rotation of the slab due to non-uniform shrinkage across the cross section caused by random cracking results from the stochastic concrete tensile strength. Moreover, this also ensures a better numerical stability for the analyses. All specimens, except slab RS2-b, were cast from the same batch of concrete. Since the properties of the two batches of concrete were similar, only one set of the material parameters was used in the numerical study. The material properties for the concrete were: f cm = 24.3 MPa , E0 = 1.6 Ec.28 = 36.5 GPa , f ct.28 = 1.97 MPa ,

= 0.2 , Ash = 603 and B sh = 44 days . The growth of the mean concrete tensile
strength for aging concrete was approximated with A f ct = 2.1 MPa and

B f ct = 2.6 days (10%), as presented in Figure 4.51a. The development shrinkage strain with time is shown in Figure 4.51b. Similar to that in the flexural cracking tests, the concrete fracture energy G f was taken as 75 N m . The concrete tensile softening stress-strain curve was scaled with a crack band width of hc = 36 mm . The empirical material constants for the solidification creep model were obtained by fitting the compliance data obtained from the companion creep test. The solidification creep parameters were: q 2 = 206.6 MPa , q3 = 0.01 MPa , q 4 = 17.1 MPa and A0 = 58.8 MPa 1 . The approximated compliance curves are shown in Figure 4.52. Eight Kelvin chain units were adopted and the elastic moduli E

206

Conc. tensile strength fct (MPa)

2.5 2.0 1.5 1.0 0.5 0.0 0

Shrinkage strain ( )

Afct = 2.1 MPa Bfct = 2.6 days

600

Ash = 603 Bsh = 44 days

400

200 Experimental data Model 0 0 50 100 150 Time since commencement of drying (days)

Experimental data Model 10 20 30

Age (days)

(a)

(b)

Fig. 4.51 - Test data of companion specimens compared with models (Nejadi and Gilbert, 2004): (a) growth of concrete tensile strength; (b) shrinkage strain since commencement of drying.

and retardation times are those as given in Table 4.20. An elastic-perfectly plastic stress-strain relationship was assumed for the reinforcing steel with a yield strength of 500 MPa and an elastic modulus of 200 GPa. The parameters defining the bond model were: s1 = s2 = 0.6 mm , s3 = 1.0 mm , max = 9.9 MPa , f = 1.5 MPa and k u = 100 MPa mm .

250 Experimental data (t'=3 days) 200 J(t,t') (10-6 / MPa) 150 100 50 0 0.1 1 10 100 1000 Time under load, t-t' (days) Model (t'=3 days)

Fig. 4.52 - Compliance curve for restrained shrinkage tests (Nejadi and Gilbert, 2004).

207

Table 4.20 - Kelvin chain properties for solidification creep model.


-th unit 1 2 3 4 5 6 7 8
E (MPa) 0.07614 0.06477 0.05575 0.04859 0.04290 0.03839 0.03481 0.03197

(days) 0.0001 0.001 0.01 0.1 1 10 100 1000

4.5.3

Comparisons of Numerical and Experimental Results

The restrained deformation tests are fundamentally different from the flexural cracking tests. No external load was applied to the specimens except the tensile reactions induced at the supports and therefore no explicit displacement could be measured to characterize the response of the structure. The deformation was solely due to the development of drying shrinkage of concrete with time, which in turn induced the tensile stresses in concrete. Cracking occurs as soon as the induced tensile stresses exceed the concrete tensile strength. In the following presentation of results, a detailed comparison between the numerical and the experimental results is made only for slabs RS1-a and RS1-b, the results for all other specimens are presented only for the final age 122 days. The formation of cracks computed by the model and the crack widths at age 122 days for slabs RS1-a and RS1-b are shown in Figures 4.53a to 4.53d. The development of concrete tensile stress in the specimen between age 6 days and 50 days are shown in Figures 4.54 and 4.55. The age of the formation of the first crack calculated by the finite element model has a good correlation with the experimental observation (at age 7 days). The shrinkage-induced tension in the slab was relieved immediately after the formation of the first crack. The slab remained in tension after cracking and the reinforcing steel carried the entire induced tensile force at the crack. After the first

208

Age = 6 days

(a)

Age = 40 days

(b)

Age = 50 days

(c)

Age = 122 days

0.17 mm

0.19 mm

0.34 mm (d)

0.14 mm

Fig. 4.53 - FEM crack formation for slabs RS1-a and RS1-b (scale 40): (a) at age 6 days; (b) at age 40 days; (c) at age 50 days; (d) at age 122 days.

209

Age = 6 days

(a)

Age = 20 days

(b)

Age = 30 days

Concrete stress
(kPa)

c1
x (c)

Fig. 4.54 - FEM shrinkage induced concrete tensile stresses in slabs RS1-a and RS1-b: (a) at age 6 days; (b) at age 20 days; (c) at age 30 days.

210

Age = 40 days

(a)

Age = 50 days

Concrete stress
(kPa)

c1
x (b)

Fig. 4.55 - FEM shrinkage induced concrete tensile stresses in slabs RS1-a and RS1-b: (a) at age 40 days; (b) at age 50 days.

cracking, the concrete slab split into two intact pieces and the reinforcing steel at the crack served as a bridge to transfer the shrinkage-induced tensile force to the concrete slabs via bond action. As time increases, concrete underwent further shrinkage and concrete tensile stress began to develop again as shown in Figures 4.54 and 4.55. The regions in dark red indicate the concrete with high tensile stresses. It can be seen that the concrete tensile stress concentrated in the region near the support junctions having a change in width of the slab and also in the regions at a distance away from the first crack. The stress concentration at the supports was caused by the change in boundary conditions which can be thought of as a geometric imperfection. For the regions at a distance away from the crack, the stress concentration occurred in the concrete surrounding the reinforcing steels. At a first glance one may wonder why concrete
211

tensile stress concentrated only at these regions, and why it was not uniformly distributed throughout the section when the stresses were fully transferred between concrete and steel. By examining closely Figures 4.54 and 4.55, the concrete between the reinforcing steels and at the two sides of the slab shrinks more freely with less restraint from the steel bars than the concrete surrounding the steel bars and, therefore the concrete tensile stresses are relatively lower. This is particularly obvious for the side face concrete of the slab for which the low tensile stress region (denoted by green to yellow stress contours, for example, as can be seen in Figures 4.54b, 4.54c and 4.55a) stretches a distance up to about 10 elements width away from the midspan crack. Since tensile stresses are low at the each side of the slab and between the reinforcing bars, considering the equilibrium at the section where the concrete stress is fully developed through bond (about 8 elements width away from the midspan crack), the concrete having a higher restraint from reinforcing steel must have a higher tensile stress concentration to maintain an equilibrium state in the section. The computed crack pattern was fully developed at 50 days and no further cracking was computed up to the age 122 days as the specimen was relieved from the restraining tension and the development of shrinkage of concrete was not sufficient to further produce a tensile stress higher than the concrete tensile strength. Figure 4.56 shows the experimental crack patterns and crack widths at age 122 days. Comparing Figures 4.53d and 4.56, it is seen that the crack pattern and crack widths computed by the model agrees well with the test results. The comparison of the experimental and the calculated crack widths, steel stresses and concrete stress at age 122 days are presented in Table 4.21. A good correlation is obtained except for the steel stress at crack, for which the model calculated a lower value than that obtained from the experiment. The comparisons of the numerical and experimental results for all other specimens are presented in Figures 4.57 to 4.59 and Tables 4.22 to 4.24. In addition to the comparison of the average crack widths, the sum of all crack widths is also presented for the comparisons of the numerical and experimental results. The average crack widths are calculated based on the total number of cracks and the number of cracks can be quite different for two specimens with the same cross-section. For example, the experimental average crack widths of slabs RS3-a and RS3-b are very different (see Table 4.23) because of the difference in number of cracks observed in the
212

test. Therefore, in this case, the sum of all crack widths gives a better understanding of the overall deformation of the specimens and also provides a more meaningful comparison with the numerical results. Overall, the calculated results agree well with the test results.

0.21 mm

0.13 mm 0.37 mm (a) 0.15 mm

0.10 mm

0.13 mm

0.34 mm 0.12 mm 0.11 mm (b) Fig. 4.56 - Experimental crack patterns and crack widths for slabs RS1-a and RS1-b at age 122 days.

Table 4.21 - Comparison of experimental and FEM results for slab RS1-a and RS1-b at age 122 days. Description Average crack width (mm) Sum of all crack widths (mm) Steel stress at crack (MPa) Steel stress away from crack (MPa) Concrete stress away from crack (MPa) Slab RS1-a 0.22 0.86 273 -47.9 1.77 Slab RS1-b 0.18 0.68 190 -57.9 1.41 FEM 0.21 0.85 147 -57.8 1.40

213

0.27 mm

0.49 mm (a)

0.28 mm

0.22 mm

0.25 mm 0.43 mm (b)

0.21 mm

0.45 mm (c)

0.28 mm

Fig. 4.57 - Crack patterns and crack widths for slabs RS2-a and RS2-b at age 122 days: (a) FEM results; (b) test results for RS2-a; (c) test results for RS2-b.

Table 4.22 - Comparison of experimental and FEM results for slab RS2-a and RS2-b at age 122 days. Description Average crack width (mm) Sum of all crack widths (mm) Steel stress at crack (MPa) Steel stress away from crack (MPa) Concrete stress away from crack (MPa) Slab RS2-a 0.30 0.90 250 -41.0 1.13 Slab RS2-b 0.31 0.94 290 -75.0 1.46 FEM 0.35 1.04 292 -81.8 1.71

214

0.89 mm (a)

0.84 mm (b)

0.78 mm (c)

0.22 mm

Fig. 4.58 - Crack patterns and crack widths for slabs RS3-a and RS3-b at age 122 days: (a) FEM results; (b) test results for RS3-a; (c) test results for RS3-b.

Table 4.23 - Comparison of experimental and FEM results for slab RS3-a and RS3-b at age 122 days. Description Average crack width (mm) Sum of all crack widths (mm) Steel stress at crack (MPa) Steel stress away from crack (MPa) Concrete stress away from crack (MPa) Slab RS3-a 0.84 0.84 532 -19.2 1.45 Slab RS3-b 0.50 1.00 467 -33.4 1.31 FEM 0.89 0.89 491 -61.4 1.52

215

0.18 mm 0.16 mm

0.38 mm (a)

0.15 mm

0.18 mm

0.28 mm 0.29 mm (b)

0.18 mm

0.26 mm

0.32 mm (c)

0.16 mm

Fig. 4.59 - Crack patterns and crack widths for slabs RS4-a and RS4-b at age 122 days: (a) FEM results; (b) test results for RS4-a; (c) test results for RS4-b.

Table 4.24 - Comparison of experimental and FEM results for slab RS4-a and RS4-b at age 122 days. Description Average crack width (mm) Sum of all crack widths (mm) Steel stress at crack (MPa) Steel stress away from crack (MPa) Concrete stress away from crack (MPa) Slab RS4-a 0.23 0.93 270 -45.4 1.64 Slab RS4-b 0.25 0.74 276 -54.1 1.71 FEM 0.22 0.87 168 -58.2 1.40

216

4.5.4

Discussion

Numerical aspects of the simulation of the restrained deformation tests are here mentioned. Recalling the convergence tests discussed in Chapter 3, Section 3.7.6, a force or a displacement convergence criterion can be used in a non-linear solution procedure. Since the restrained deformation specimens were not subjected to externally applied loads, a displacement convergence criterion must be employed. Furthermore, the rate of development of both creep and shrinkage of concrete reduces with age of concrete, the time-dependent deformation of a concrete structure will also reduce with time. In a time-dependent finite element analysis, time is discretized such that the time steps are small at early ages and are gradually increased as the rate of deformation of the structure reduces at an older age. Other than ensuring an effective usage of the computer resources, the time discretization method also maintains a stable numerical procedure. It ensures that the increase in deformation of a structure within a particular time step is not too small to reach the prescribed tolerance for displacement convergence. In the analyses of the restrained deformation specimens, small time steps were used at early ages and hence the age of first cracking could be easily traced. As time increases, the time steps advance with larger time intervals, which causes difficulties in tracing the age of formation for each crack. For example, taking the simulation of slab RS1-a or RS1-b (see Figure 4.53), a 10-day time step was used between age 40 days and age 50 days and three cracks were computed within this time step. An attempt was made to trace the formation of each individual crack. However, convergence was difficult due to the use of small time steps. Despite so, the numerical results have a good agreement with the experimental results. In the authors view, even if the model were able to trace one crack at a time, no further cracking would have been computed after the formation of the second crack, since the shrinkage-induced tension in the specimen is greatly relieved and the development of subsequent shrinkage in the aging concrete is too low to induce further cracking within the duration of the test. This invoked further investigation into the details of the experimental results. By analysing the concrete surface strains measured from the experiment, it is believed that, one of the factors causing restrained shrinkage cracking, which was unaccounted for in most previous studies, is the coalescence of
217

microcracks induced by the internal restraints within the concrete member during the entire shrinkage process. To illustrate this phenomenon, the concrete surface strain measurements with time of two typical specimens are referred to, as shown in Figures 4.60a and 4.60b. The corresponding locations of the measurements are shown in Figure 4.60c. The curves with marked indicators in Figures 4.60a and 4.60b are the strain measurements of the locations containing the cracks of the specimens where the tensile strains are relatively high. After first cracking, the concrete away from the crack began to shrink, which is indicated as negative strain measurements in Figures 4.60a and 4.60b. By examining the test results carefully, it can be noticed that the compressive strains in the concrete are not uniform throughout the slab length. For concrete adjacent to the crack, this is because of the low restraint imposed by the reinforcement before the bond was fully

1500

1750

Concrete surface strain ( )

1250 1000 750 500 250 0 -250 -500 0 20 40 60 80 100 120 140

Concrete surface strain ( )

1 2 3 4 5 6 7 8 9 10 11

1500 1250 1000 750 500 250 0 -250 -500 -750 0 20 40 60 80 100 120 140

1 2 3 4 5 6 7 8 9 10 11

160

160

Age (days)

Age (days)

(a)

(b)

11

10

(c)

DEMEC surface strain target

Fig. 4.60 - Experimental measurements of concrete surface strain: (a) slab RS1-a; (b) slab RS2-b; (c) locations of the DEMEC surface strain targets.

218

developed. For concrete elsewhere, in the authors view the non-uniform compressive concrete strain is caused by the shrinkage-induced microcracking due to the internal restraints. It is well recognized that the presence of restraints is the major cause of shrinkage cracking in reinforced concrete structures. Restraints on shrinkage of concrete can occur at various scales. The finite element simulation of the restrained deformation tests has taken into consideration both the external and internal restraints to shrinkage. These include the external restraint imposed at the end supports and the internal restraint imposed by the reinforcement. However, the inherent internal restraints on the mezzo level within concrete, which consist of aggregates restraint and self-restraint caused by differential shrinkage, were not accounted for or, more precisely, out of the capability of the present macroscopic modelling method. These internal restraints are said to be part of the causes of microcracking in concrete, which had been investigated extensively by, for example, Baant and Ralfshol (1982) and Bisschop (2002). In most studies, restrained deformation cracking of reinforced concrete members is treated in a similar manner to that in a load-induced direct tension member. The only difference is that, instead of applying external loads, tension in a restrained member develops through the development of drying shrinkage. For a restrained concrete member, the first crack forms at the weakest section as shrinkage develops and the tensile stress in concrete away from the crack is relieved by the formation of the crack. The tensile stress in the concrete increases as further shrinkage takes place. A new crack will form if the tensile stress in concrete violates the tensile strength criterion and the tensile force in the member is once again relieved. This process takes place until the shrinkage-induced tension becomes insufficient to cause further cracking. This theory assumes restrained shrinkage cracking as a progressive process where one crack can only form at a time. However, the theory above cannot explain the observations in the restrained deformation tests. For example, the second to the fourth crack of slab RS1-a (see Figure 4.60a) occurred simultaneously at age about 30 days, which obviously disagrees with the aforementioned theory which assumes crack to form individually at a section at a time instance. For slab RS2-b (see Figure 4.60b), although the second and third crack

219

did not appear on the same day, the compressive strains over the cracks were constantly low compared to the strain measurements at other locations. This indicates high tension was induced simultaneously at the two different locations. From the experimental observations, it can be postulated that the concrete must have undergone some sort of deteriorating mechanism and the author believes that shrinkage microcracking due to internal restraints is the main factor causing this phenomenon.

4.6 Other Numerical Examples


4.6.1 Continuous Beams Subjected to Long-term Sustained Load

Two simply supported beams and two continuous beams subjected to long-term sustained loads were tested by Bakoss et al. (1982) to compare the experimental measurements with the results calculated using empirical design approaches. The simply supported beams were similar to those tested by Gilbert and Nejadi (2004) for which the loading points were at the third points of the beams, thus finite element analysis was only carried out on the continuous beams. The two continuous beams were identical in both setting and cross-section. The cross-sections of the beams were 100 mm wide and 150 mm deep with two 12 mm diameter bars at an effective depth of 130 mm on the tension sides of the beams, as indicated in Figure 4.61a. A 6 kN point load was applied at the midspan of each of the two equal 6 m spans at 23 days after casting. In addition to the beam tests, Bakoss et al. also tested companion specimens cast from the same batch of concrete in order to measure the creep and shrinkage properties of the concrete. The localized cracking crack band model was employed to model the continuous beam. One half of the beam was modelled with symmetry and the finite element mesh is shown in Figure 4.61b. The mesh consists of 948 nodes, 714 concrete elements, 119 steel truss elements, 119 bond-slip interface elements and 3 stiff elastic support

220

100 212 bars 150 130 20 212 bars

Section A-A

Section B-B (a) 6 kN

Section C-C

A 6000 mm (b)

Fig. 4.61 - Details of Bakoss et al.s (1982) continuous beam: (a) cross-sections of the beam; (b) finite element mesh.

elements. The steel truss elements were linked to the concrete elements via bond-slip interface elements. The reinforcing steel at the left end of the beam in Figure 4.61b was assumed to have full anchorage where no slip is permitted. In the experiment, the continuous beams were loaded horizontally with the beams suspended sideways therefore the self-weight of the beams was omitted in the analysis. Bakoss et al. reported the creep property as creep coefficients with increasing time for the concrete first loaded at age 23 days. The creep coefficients were converted into a compliance function using the relationship J (t ,23) = [1 + (t ,23)] E c.23 . The concrete parameters used in the analysis were: E c.23 = 27.3 GPa ,

E 0 = 1.6 E c.28 = 49.9 GPa , f cm = 39 MPa , f ct.28 = 2.5 MPa , hc = 35 mm, = 0.2, Ash = 850 , Bsh = 150 days , q 2 = 125 MPa , q3 = 0.01 MPa ,

q 4 = 26 MPa and A0 = 35.5 MPa 1 . The approximated time-dependent shrinkage strain curve and compliance curve are shown in Figures 4.62a and 4.62b, respectively. The elastic moduli and retardation times of the Kelvin chain units for the creep model

221

are presented in Table 4.25. By using the equation recommended by AS 3600 (2001), as given by Eq. 2.1, the mean concrete tensile strength was calculated from the experimental concrete compressive strength at different ages. The time-dependent growth of the mean concrete tensile strength was approximated by A f ct = 3.1 MPa and B fct = 10 days with the concrete tensile strength for each element assigned randomly at 10% of the mean value. The concrete fracture energy G f was taken as 70 N mm in accordance with CEB-FIP Model Code 1990 (1993). The reinforcing steel was modelled as elastic-perfectly plastic material with a yield strength of 400 MPa and an

800

Shrinkage strain ( )

Ash = 850 Bsh = 150 days J(t,t') (10-6 / MPa)

200 Experimental data (t'=23 days) 150 Model (t'=23 days)

600

400

100

200

Experimental data Model

50

0 0 200 400 600 800

0 0.1 1 10 100 1000

Time since commencement of drying (days)

Time under load, t-t' (days)

(a)

(b)

Fig. 4.62 - Creep and shrinkage measurements of Bakoss et al. (1982) compared with models: (a) shrinkage strain since commencement of drying; (b) compliance curve.

Table 4.25 - Kelvin chain properties for solidification creep model.


-th unit 1 2 3 4 5 6 7 8
E (MPa) 0.12584 0.10705 0.09214 0.08030 0.07091 0.06345 0.05753 0.05283

(days) 0.0001 0.001 0.01 0.1 1 10 100 1000

222

elastic modulus of 200 GPa. The parameters for the CEB-FIP 1990 (1993) bond-slip model were: s1 = s2 = 0.6 mm , s3 = 1.0 mm , max = 12.5 MPa , f = 1.9 MPa and an unloading modulus of 100 MPa mm was assumed. The midspan deflection calculated by the finite element model is compared with the experimental average deflection for the two beams in Figure 4.63. The model calculated a slightly lower time-dependent deflection than the experimental measurement but with a difference not more than 10%, for which the correlation is considered reasonable. No further experimental data was reported by Bakoss et al. except the midspan deflection with time. To demonstrate the formation of crack at discrete location, the crack patterns computed by the model at instantaneous loading, at 50 days and at 400 days under sustained load are shown respectively in Figures 4.64a, 4.64b and 4.64c. The model computed an average crack spacing of 70 mm. Without the available experimental data, a comparison is made with the average crack spacings calculated using the tension chord model and the recommendation of CEB-FIP Model Code 1990 (1993). For stabilized cracking, the average crack spacing is given by (CEB-FIP, 1993) s rm = 2 s rm0 3 with s rm 0 = 3.6 eff (4.1)

where s rm 0 is the maximum crack spacing, is the bar diameter and eff is the effective reinforcement ratio defined as the ratio of the area of reinforcing steel and the effective area of concrete in tension (see Section 3.3.1). The comparison of the results is shown in Table 4.26. It is seen that the finite element model computed the largest average crack spacing while the tension chord model gives the lowest prediction.

223

16 Midspan deflection (mm) 14 12 10 8 6 4 2 0 0 100 200 Time (days) 300 400


Experimental FEM

Fig. 4.63 - Comparison of FEM and experimental midspan deflections versus time.

(a)

(b)

(c) Fig. 4.64 - Crack patterns computed by FEM (principal cracking strain plot): (a) instantaneous loading; (b) 50 days after loading; (c) 400 days after loading.

224

Table 4.26 - Comparison of average crack spacing calculated by different models. Model description FEM Tension chord model CEB-FIP Model (1993) Average crack spacing 70 mm 42 mm 51 mm

In the analysis, the beam was cracked at instantaneous loading. The cracks of the beam widened with time due to the effects of creep and shrinkage, as can be seen in Figure 4.64. A summary of the crack opening over 400 days is given in Table 4.27. The average crack widths were calculated by averaging the widths of the cracks within the region of moment greater than 80% of the maximum moments. The maximum crack opening was found to occur at the maximum negative moment region above the middle support where the bending moment is highest in the beam.

Table 4.27 - Crack widths calculated by FEM at various times t after first loading. Region Instantaneous Positive moment region Avg. Max. Negative moment region Avg. Max. 0.025 0.026 0.036 0.048 Crack width (mm) t = 100 days 0.045 0.055 0.062 0.0.82 t = 400 days 0.062 0.078 0.083 0.106

4.6.2

Time-dependent Forces Induced by Supports Settlement of Continuous Beams

A series of tests was performed by Ghali et al. (1969) to investigate the time-dependent reaction forces induced by the differential settlement of supports of continuous reinforced concrete beams. The test consisted of 4 pairs of two-span continuous beams. Each pair of the beams was tested in a vertical position so as to disregard the bending caused by the self-weight. Rollers were inserted at the end of the beams to separate the beam set and a mid-support settlement (or deflection) was introduced by means of
225

two threaded bars tying the beams. The applied settlement was monitored by two dial gauges and the induced forces were recorded throughout the duration of the test. The setup and the details of the beams are shown in Figure 4.65. Each pair of the beams was subjected to the same final deflection of 1.65 mm but with different deflection increments applied at different ages. The beams were tested up to a period of 300 days. The details of the age of application of deflections for each beam set are indicated in Table 4.28. In addition to the support settlement tests, a beam with the same layout and cross-section was subjected a sustained load of 17.3 kN at the mid-length of the beam in the absence of the middle support. The beam was also tested in a vertical position therefore the bending effect due to its self-weight was eliminated.

Dial gauge

914.4

A 2133.6

Calibrated rod

Threaded bar

Section A-A (b)

914.4

Dial gauge

212.7 25.4

212.7
25.4

Stirrups
6.35 at 152.4
209.6

(c) (a) Fig. 4.65 - Details of Ghali et al.s (1969) continuous beams: (a) longitudinal layout of a beam set; (b) section of the test; (c) cross-section of the beam.

226

Table 4.28 - Details of application of deflections.


Test No. 1 2 3 4 Age (days) of application of deflections for increment number 1 9 12 12 11 2 12 13 15 3 12 15 27 4 13 20 41 5 14 26 72 Deflection increment (mm) 1.65 0.33 0.33 0.33 Duration for each deflection increment (minute) 30 10 10 10

Localized cracking is not the major interest in this investigation and in addition, the beams contained both longitudinal and transverse reinforcements, it is more appropriate to adopt a distributed cracking model, in this study it is the cracked membrane model. By exploiting the symmetric nature of the beam, only one half of the beam is required for the simulation. The mesh of the beams is shown Figure 4.66 with the material regions indicated. The mesh is made up of 374 nodes and contains 320 reinforced concrete elements and 3 stiff elastic support elements. Table 4.29 shows the reinforcement properties of the reinforced concrete zones indicated in Figure 4.66.

RC zone 1

RC zone 2

Stiff elastic elements

Fig. 4.66 - Finite element mesh for Ghali et al.s (1969) beams.

Table 4.29 - Reinforcement properties for Ghali et als (1969) beams.


RC zone 1 2 Reinforcement ratio

x 0 0.049090

y 0.004091 0.004091

227

The time-dependent properties of the concrete were not reported by Ghali et al. (1969). To facilitate the numerical analysis, the compliance function and shrinkage function recommended by CEB-FIP Model Code 1990 (1993) were adopted and were incorporated into the finite element model by using the appropriate creep and shrinkage parameters. A relative humidity of 65% was used in the calculation of the compliance function and shrinkage function. The CEP-FIP 1990 creep and shrinkage models are presented in Appendix C. Ghali et al. tested a number of concrete cylinders to obtain the compressive strengths at ages range from 7 days to 190 days and approximated the growth of concrete compressive strength by a function of age t given by f c' = 37.92 7 t + 0.75 (4.2)

The growth of concrete tensile strength can be calculated using the recommendation in AS 3600 (2001) (Eq. 2.1, Chapter 2) based on the compressive strengths at various ages. The concrete parameters used in the analysis were: E0 = 1.6 Ec.28 = 46.3 GPa , f cm = 38 MPa , f ct.28 = 2.5 MPa , A fct = 2.8 MPa B fct = 3.5 days ,

= 0.2 ,

Ash = 450 ,

Bsh = 60 days ,

q 2 = 142.3 MPa ,

q3 = 6.4 MPa , q 4 = 16.3 MPa and A0 = 40.4 MPa 1 . The elastic moduli and corresponding retardations times of the Kelvin chain units are tabulated in Table 4.30. The concrete fracture energy G f was taken as 75 N mm . For the reinforcing steel, an elastic-perfectly plastic stress-strain relationship with an elastic modulus of 200 GPa and a yield strength of 400 MPa was employed. Figures 4.67a to 4.67d show the comparisons of the numerical and experimental results for the reactions at mid-supports with time induced by the applied deflections. It is seen that the model calculated a more rapid change in reaction than the test results. This is attributed to the effects of creep and shrinkage of concrete. A better correlation could be obtained if the experimental creep and shrinkage data were available. Nevertheless, the results calculated by the model are overall in a good agreement with the test data. In Figure 4.68, the calculated midspan deflection for the beam subjected to a sustained load is compared with the test results with a good correlation obtained.

228

Table 4.30 - Kelvin chain properties for solidification creep model.


-th unit 1 2 3 4 5 6 7 8
E (MPa) 0.11054 0.09404 0.08094 0.07054 0.06229 0.05574 0.05054 0.04641

(days) 0.0001 0.001 0.01 0.1 1 10 100 1000

Reaction at mid-support (kN)

Reaction at mid-support (kN)

20.0 18.0 16.0 14.0 12.0 10.0 8.0 6.0 4.0 2.0 0.0 0 50 100 150 Age (days)

18.0 16.0 14.0 12.0 10.0 8.0 6.0 4.0 2.0 0.0 0 50 100 150 Age (days) 200 250
Experimental FEM

Experimental FEM

200

250

(a)
Reaction at mid-support (kN)
16.0 14.0 12.0 10.0 8.0 6.0 4.0 2.0 0.0 0 50 100 150 Age (days) 200 250
Experimental FEM

(b)
Reaction at mid-support (kN)
16.0 14.0 12.0 10.0 8.0 6.0 4.0 2.0 0.0 0 50 100 150 Age (days) 200 250
Experimental FEM

18.0

(c)

(d)

Fig. 4.67 - Comparison of FEM and experimental time-dependent reactions at midsupports of Ghali et al.s (1969) controlled deflection specimens: (a) to (d) test 1 to test 4, respectively.

229

4.0 Midspan deflection (mm) 3.5 3.0 2.5 2.0 1.5 1.0 0.5 0.0 0 50 100 150 Time (days) 200 250
Experimental FEM

Fig. 4.68 - Comparison of FEM and experimental time-dependent midspan deflection for the beam subjected to a sustained load of Ghali et al. (1969).

4.6.3

Slender Columns Subjected to Long-term Eccentric Axial Loads

This section deals with a slightly different type of structural problem where nonlinearity does not arise solely from the material properties of reinforced concrete but also from the excessive deformation of the structure due to the effects of creep. This geometric non-linearity may lead to creep buckling and is common in slender reinforced concrete columns subjected to sustained eccentric axial compression. Bradford (2005) tested a series of eccentrically loaded slender reinforced concrete columns and used the test results to investigate the behaviour of slender columns analytically. Five identical columns were tested with various degree of end eccentricity. All columns were 5 m long and the eccentric compressive loads were applied at the two ends of the columns. The test arrangement and the cross-section details of the specimens are shown in Figure 4.69. Three dial gauges were placed on the side face of the column, as shown in Figure 4.69a, in order to measure the sideways deflections. One dial gauge was placed at the mid-height whilst the other two were positioned at the quarter points. The details of the compressive loads and the end eccentricities eT and eB as shown in Figure 4.69a are given in Table 4.31.

230

Steel channel section

eT
Strong wall

Eccentric loading Tensioning cable

Dial gauge

Dial gauge

1250 A A 1250 5000

Dial gauge

1250

Test column

1250

eB
(a) Stirrups 10 at 150 2N12

Hydraulic jack I-section loading arm Load cell

Clear cover 15 mm 2N12

150
Section A-A (b) Fig. 4.69 - Details of Bradfords (2005) slender columns: (a) layout of the test; (b) cross-section of the column.

231

Table 4.31 - Details of the loading conditions for Bradfords (2005) columns. Specimen eT (mm) eB (mm) Load (kN) C1 50 50 70.0 C2 50 25 70.0 C3 50 0 80.0 C4 50 -25 80.0 C5 50 -50 85.0

The distributed cracking cracked membrane model was used to simulate these tests. The finite element mesh for the columns is shown in Figure 4.70 and the reinforcement details for the mesh are given in Table 4.32. The 200 rows by 6 columns mesh consists of 1407 nodes and 1200 reinforced concrete elements. The properties of concrete were taken as those reported by Bradford. The tensile strength of concrete, which was not reported by Bradford, was taken as that calculated using the equation given in AS 3600 (2001). The parameters for creep and shrinkage were determined by fitting the test data obtained from the creep and shrinkage tests conducted in conjunction with the slender column tests. Figures 4.71a and 4.71b show

50

50

50

50

50

(200 rows of elements)

5000

50

(a)

25

(b) RC zone 1

(c)

(d)

25

(e)

50

RC zone 2

Fig. 4.70 - Finite element mesh for Bradfords (2005) columns: (a) column C1; (b) column C2; (c) column C3; (d) column C4; (e) column C5.

232

Table 4.32 - Reinforcement properties for Bradfords (2005) columns. RC zone 1 2 Reinforcement ratio

x 0.00711 0.00711

y 0.03492 0

the approximated curves for shrinkage strain and for compliance data, respectively. Since the age of loading for the columns were not reported explicitly by Bradford, it was taken as 12 days in this study and it is the loading age for the creep test. The concrete parameters used in the simulation were: E0 = 1.6 Ec.28 = 35.4 GPa , f cm = 29.3 MPa , Ash = 420 , f ct.28 = 2.2 MPa , Bsh = 90 days , A f ct = 2.2 MPa B f ct = 12 days , = 0.2 , q3 = 2.5 MPa ,

q 2 = 80.2 MPa ,

q 4 = 38.5 MPa and A0 = 22.8 MPa 1 . The elastic moduli and retardations times for the Kelvin chain units are shown in Table 4.33. Similar to the previous examples, the concrete fracture energy G f was taken as 75 N mm and the stress-strain relationship for reinforcing steel was taken to be elastic-perfectly plastic with an elastic modulus of 200 GPa and a yield strength of 500 MPa.

400

Shrinkage strain ( )

Ash = 420 Bsh = 90 days J(t,t') (10-6 / MPa)

250 Experimental data (t'=12 days) 200 150 100 50 0 Model (t'=12 days)

300

200

100

Experimental data Model

0 0 50 100 150 200 250 300 350

0.1

10

100

1000

Time since commencement of drying (days)

Time under load, t-t' (days)

(a)

(b)

Fig. 4.71 - Comparison of experimental shrinkage and creep measurements (Bradford, 2005) with approximated models: (a) shrinkage strain; (b) compliance curve.

233

Table 4.33 - Kelvin chain properties for solidification creep model.


-th unit 1 2 3 4 5 6 7 8
E (MPa) 0.19612 0.16685 0.14361 0.12516 0.11051 0.09889 0.08966 0.08234

(days) 0.0001 0.001 0.01 0.1 1 10 100 1000

The deflected shapes calculated by the finite element model are shown in Figure 4.72. It is seen that the column with the largest equal end eccentricities (specimen C1) has the largest deflection, whilst column C5, which was subjected to the largest unequal end eccentricities, has the smallest deflection. The calculated deflections at mid-height and top and bottom quarter points with time are compared with the experimental results in Figure 4.73. Although the calculated results for column C1 are slightly lower than that observed in the test, the overall correlation between the model results and the test data is reasonable.

(a)

(b)

(c)

(d)

(e)

Fig. 4.72 - Deflected shapes computed by the model (scale 30): (a) column C1; (b) column C2; (c) column C3; (d) column C4; (e) column C5.

234

50.0 45.0 Deflections (mm) 40.0 35.0 30.0 25.0 20.0 15.0 10.0 5.0 0.0 0

Column C1
Deflections (mm)

30.0 25.0 20.0 15.0 10.0 5.0 0.0

Column C2

Exp (top) Exp (mid) Exp (bot)

FEM (top) FEM (mid) FEM (bot)

Exp (top) Exp (mid) Exp (bot)

FEM (top) FEM (mid) FEM (bot)

100 200 Time since loading (days)

300

(a)

100 200 Time since loading (days)

300

(b)

25.0 20.0 15.0 10.0 5.0 0.0 0

Column C3
Deflections (mm)

12.0 10.0 8.0 6.0 4.0 2.0 0.0

Column C4

Deflections (mm)

Exp (top) Exp (mid) Exp (bot)

FEM (top) FEM (mid) FEM (bot)

Exp (top) Exp (mid) Exp (bot)

FEM (top) FEM (mid) FEM (bot)

100 200 Time since loading (days)

300

(c)

100 200 Time since loading (days)

300

(d)

5.0 4.0 3.0 2.0 1.0 0.0 -1.0 -2.0 -3.0 -4.0 -5.0 -6.0 0

Column C5

Deflections (mm)

Exp (top) Exp (mid) Exp (bot)

FEM (top) FEM (mid) FEM (bot)

100 200 Time since loading (days)

300

(e)

Fig. 4.73 - Comparison of FEM and experimental deflections versus time: (a) column C1; (b) column C2; (c) column C3; (d) column C4; (e) column C5.

Comparing the calculated central deflection for column C5 with the test data, the model seems to have computed the column to deflect to the wrong side since the calculated and experimental central deflections are of different signs as seen in Figure 4.73e. A close examination reveals that the computation was theoretically sensible.

235

Although column C5 was subjected to end eccentricities of 50 mm, the magnitudes of the end moments were not identical. The column had a larger moment at the bottom end due to the self-weight of the column. This can be seen in the test results for the quarter points shown in Figure 4.73e, the deflection of the bottom quarter point is slightly larger than that of the top quarter point. Due to the larger moment at the bottom end, the central deflection should also deflect to the same direction as the bottom quarter point, as computed by the model. The author believes that the difference between the results is due to vagaries of experiments or more undesirably, errors arising when measuring the test data.

236

CHAPTER 5 NUMERICAL EXPERIMENTS

5.1 Introduction
Numerous studies were conducted in the 1950s to 1970s to gain better understanding of the cracking mechanisms of reinforced concrete structures (for example, Clark, 1956, Broms., 1965, Base et al., 1966, Ferry-Borges, 1966, Gergely et al., 1968 and Nawy et al., 1970). However, these studies are limited to instantaneous cracking. Cracking of reinforced concrete structures under long-term loads has received relatively less attention. Although there are studies dedicated to the investigation of the timedependent behaviour of reinforced concrete, their objectives are usually concerned with the variation of deflection with time. Relatively little attention has been given to the study of the time-dependent change in crack widths and crack spacings in reinforced concrete structures. The reason for this is that time-dependent tests are not only time consuming but also require a large amount of laboratory resources. Cracking of reinforced concrete is well known for its semi-random nature. A large number of specimens are needed for experimentation in order to obtain statistically representative results. Therefore, the need for a large laboratory area over a lengthy period is unavoidable and expensive. In this chapter, a number of series of numerical experiments were devised and analysed using the numerical model described in Chapter 3 in order to investigate timedependent cracking of reinforced concrete structural elements. The primary aims of the investigation are to identify the major parameters influencing time-dependent flexural cracking and to acquire a better qualitative understanding of the interactions of the parameters. As a result, the localized cracking Crack Band Model with the use of bondslip interface elements (see Section 3.4 for formulation and Chapter 4 for evaluation) is better suited for this parametric investigation compared to the localized cracking non-

237

local model due to its capability of modelling the concrete-steel stress transfer via bond and the subsequent formation of cracks in a reinforced concrete structure, while remaining more computationally effective. This investigation essentially focuses on the following aspects of cracking in reinforced concrete structures: (1) Short-term and long-term crack spacings; (2) Short-term and long-term crack widths; (3) The ratio of long-term to short-term crack width; The factors affecting the time-dependent behaviour of reinforced concrete structures can be classified primarily into three categories: (i) factors associated with material properties; (ii) factors related to environmental conditions (such as relative humidity); and (iii) factors related to the structural type, including geometry and boundary conditions (Gilbert, 1979). In this study the following parameters have been selected as variables to be investigated: (A) Bottom concrete cover cb ; (B) Diameter of tensile reinforcing steel st ; (C) Quantity of tensile reinforcement Ast ; (D) Quantity of compressive reinforcement Asc ; (E) Magnitude of mean tensile strength of concrete f ct ; (F) Bond strength between steel and concrete max ; (G) Random fluctuation limit of concrete tensile strength l fct ; (H) Magnitude of creep; (I) Magnitude of shrinkage; (J) Bond creep; (K) Quantity of shear reinforcement Asv ;

238

(L) Yield stress of reinforcing steel f sy ; (M) Load histories; (N) Geometry of sections; (O) Boundary conditions. The numerical experiments were designed so that a comprehensive parametric investigation of cracking could be undertaken in reinforced concrete beams and slabs under service loads. In each series of tests, one parameter is selected as the variable while the others are kept constant or otherwise manipulated in order that their effects are not significant on the crack width and crack spacing. In this way the effect of the selected variable can be thoroughly examined.

5.2 Description of Numerical Experiments


Beam and slab specimens were used in the numerical experiments with different section geometries. Both the beam and slab specimens were subjected to two types of boundary conditions so as to examine the effects of boundary conditions on timedependent cracking. A simply supported case and a continuous case, as shown in Figure 5.1, were considered in this study. All specimens were first loaded at age 14 days and, unless stated otherwise, the benchmark material properties for each specimen were taken as follows: f cm = 25 MPa , f ct = 2 MPa , E 0 = 1.6 E c.28 = 40 GPa , = 0.2 and G f = 70 N/m . A 10% random fluctuation of the mean concrete tensile strength was assigned to the concrete elements. Shrinkage was assumed to commence at age 14 days and the shrinkage parameters were taken as Ash = 600 and Bsh = 45 days , which give a
* final shrinkage strain sh = 600 . The solidification creep parameters were obtained

by

fitting

the

creep

curves

of

CEB-FIP

Model

Code

1990

and

were

q 2 = 121.1 MPa , q3 = 6.7 MPa , q 4 = 20.8 MPa , which correspondingly

239

L/3

L/3 L = 5000 (a) P P

L/3

Tensile steel Compressive steel

L/3

L/3 L = 5000 (b)

L/3

Fig. 5.1 - Boundary conditions for beam and slab specimens: (a) simply supported; (b) continuous.

give a final creep coefficient * = 2.2 . The negative infinity area of the continuous retardation spectrum was A0 = 34.3 MPa-1. Eight Kelvin chain units were used to store viscoelastic strain history and the corresponding elastic modulus E and retardation time for each Kelvin chain unit are shown in Table 5.1. Reinforcing steel was taken as an elastic-perfectly plastic material with a yield strength f sy = 500 MPa and an elastic modulus E s = 200 GPa . The parameters for the bond model were:

s1 = s2 = 0.6 mm , s3 = 1.0 mm , max = 10.0 MPa , f = 1.5 MPa and the unloading modulus was taken as k u = 100 MPa mm .

240

Table 5.1 - Kelvin chain properties for solidification creep model.


-th unit 1 2 3 4 5 6 7 8
E (MPa) 0.12989 0.11050 0.09511 0.08289 0.07319 0.06549 0.05938 0.05453

(days) 0.0001 0.001 0.01 0.1 1 10 100 1000

5.2.1

Beam Specimens

The details of a typical beam section are shown in Figure 5.2. The beams were 300 mm wide and the effective depth to the tensile reinforcement was consistently 450 mm in both the positive and negative moment regions. The top concrete cover was kept the same as the bottom concrete cover for all specimens. A continuous beam can withstand much higher loads than a simply supported beam of the same clear span due to the additional moment resistance at the supports. The capability of taking high loads inevitably leads to a high shear force in the shear span of the specimens (distance from support to the position of the load). Since the major purpose of this study is to investigate time-dependent flexural cracking, it is undesirable for the specimens to fail prematurely in shear. According to the design
300

Asc
b

450

Ast Fig. 5.2 - Typical cross-section for beam specimens.

241

procedures of AS 3600 (2001), calculation shows that some specimens will experience shear failure rather than flexural failure if shear reinforcement is not provided. Therefore, for the sake of preventing premature shear failure in the numerical beam experiments, unless stated otherwise, all continuous beams are reinforced with 2-legged N10 stirrups at 120 mm spacing over the shear span. The concrete cover cb (Figure 5.2) is measured from the top or bottom edge of the beam to the nearest side of the longitudinal reinforcement (and not to the stirrup). Although in real structures cover is usually measured to the stirrup, it does not matter in the current two-dimensional modelling since each stirrup acts essentially as a vertical tie in the beam to prevent excessive shear cracking. The finite element meshes for the simply supported and continuous beam specimens are shown in Figures 5.3 and 5.4, respectively. The tensile reinforcement is shown as solid lines and the compressive reinforcement is shown as dashed lines in each figure. All meshes were generated such that the crack band width is hc = 40 mm.

(a) Asc

Ast

Ast

(b) Tensile reinforcement Compressive reinforcement

Fig. 5.3 - Finite element meshes for simply supported beam specimens: (a) beam with tensile reinforcement; (b) beam with both tensile and compressive reinforcement.

242

The longitudinal steel elements were connected to the concrete elements via bond-slip interface elements while the transverse steel elements were overlaid directly onto the concrete elements. For the singly reinforced simply supported beams (see Figure 5.3), the mesh consists of 941 nodes and is made up of 792 concrete elements, 66 steel elements, 66 bond-slip elements and 4 stiff elastic support elements. The simply supported beam with compressive reinforcement has an additional 65 nodes, 66 steel elements and 66 bond-slip interface elements. The nodes of the steel elements at the outer edge of the beam were rigidly connected to the nodes of the concrete elements in order to simulate full anchorage of the reinforcement.

Ast-

Ast

(a)
Nodal enslavement in horizontal direction

Ast-

Asc

Asc-

(b)

Ast

Legend for longitudinal reinforcement: Tensile reinforcement Compressive reinforcement Fig. 5.4 - Finite element meshes for continuous beam specimens: (a) beam with tensile reinforcement; (b) beam with both tensile and compressive reinforcement.

243

The mesh of the continuous beam containing tensile reinforcement only is made up of 962 nodes, 792 concrete elements, 91 longitudinal steel elements, 91 interface elements and 204 transverse steel elements. The detailing of the longitudinal reinforcing steel was done in accordance with the guidelines in AS 3600 (2001), so that the tensile reinforcement, both top and bottom, was extended a distance equal to the total depth of the section past the point of inflection. The major aim of this study is to investigate time-dependent flexural cracking of reinforced concrete members. Instead of providing roller supports at the vertical row of nodes at the right boundary, the mesh was enslaved in the horizontal direction so as to maintain a zero slope at the midspan and to prevent boundary-induced membrane actions due to arching effect.

5.2.2

Slab Specimens

The slab section used in the numerical experiments was 1000 mm wide, with an effective depth of 200 mm. Like the beam specimens, the top and bottom reinforcement had the same concrete cover. Figure 5.5 shows the typical cross-section of a slab specimen containing compressive reinforcement. Both simply supported and continuous slabs were analysed in this study and the arrangement of these two types of slabs has been shown previously in Figure 5.1.

200

Asc Ast

Fig. 5.5 - Typical cross-section for slab specimens.

The same crack band width ( hc = 40 mm) as for the beam meshes was used to generate the finite element meshes for the slab specimens. Figures 5.6 and 5.7 show the finite element meshes for the simply supported and continuous slabs, respectively. The
244

mesh for the singly reinforced simply supported slab (Figure 5.6a) contains 539 nodes, 396 concrete elements, 66 steel elements, 66 bond-slip interface elements and 4 stiff elastic support elements. The bond-slip interface elements were placed between the concrete and steel elements for transfer of stress via bond. The overlapping nodes of the concrete and steel elements at the left edge of the mesh were rigidly connected so as to simulate full anchorage of the reinforcement. For the doubly reinforced simply supported slab (Figure 5.6b), there is an extra row of steel and bond-slip interface elements in the compression zone, which leads to an additional 65 nodes, 66 steel elements and 66 bond-slip interface elements.

(a) Asc

Ast

Ast

(b) Tensile reinforcement Compressive reinforcement

Fig. 5.6 - Finite element meshes for simply supported slab specimens: (a) slab with tensile reinforcement; (b) slab with both tensile and compressive reinforcement.

The slab specimen has a much larger cross-sectional area than the beam specimen and hence has a higher shear strength. A check in accordance with AS 3600 showed that no shear reinforcement was required for the continuous slab specimens. Furthermore, an additional development length equal to the total depth of the section was provided for both the top and bottom tensile reinforcements at the point of inflection. The finite element mesh for the continuous slab specimens containing tensile reinforcement only (Figure 5.7a) is made up of 548 nodes, 396 concrete elements, 79 steel elements and 79 bond-slip interface elements and the mesh for continuous slabs
245

with both tensile and compressive reinforcement (Figure 5.7b) has 599 nodes, 396 concrete elements, 132 steel elements and 132 bond-slip interface elements. As for the previous finite element meshes, the steel element nodes were connected to the concrete nodes via bond-slip interface elements. For the same reasons as discussed for the continuous beam specimens, the nodes at the right edge of the continuous slab specimens were enslaved in the horizontal direction.

Ast-

Ast

(a)
Nodal enslavement in horizontal direction

Ast-

Asc

Asc-

(b)

Ast

Legend for longitudinal reinforcement: Tensile reinforcement Compressive reinforcement Fig. 5.7 - Finite element meshes for continuous slab specimens: (a) slab with tensile reinforcement; (b) slab with tensile and compressive reinforcements.

5.2.3

Testing Method

This section describes the series of tests (designated test series A to M). In each series a parameter was selected as the experimental variable, while other parameters were held constant. The number and designation of the tests on each type of specimen are presented in Table 5.2.

246

Table 5.2 - Tests conducted on specimens. No. 1 2 3 4 Specimen Simply supported beam specimens Continuous beam specimens Simply supported slab specimens Continuous slab specimens Test Series A to J, L and M A to K A to J, L and M A to J Total No. of tests 12 tests 11 tests 12 tests 10 tests

5.2.3.1

Test Series A to J: Material and Environmental Parameters

The details of the combinations of variables are given in Tables 5.3 and 5.4 for beam and slab specimens, respectively. The convention for the test numbering is such that the first letter denotes the member type (B for beam or S for slab), the second letter denotes the test series (A to M) followed by the test number in that particular series. The test designated B-A0 and S-A0 are control specimens. The shaded columns in Tables 5.3 and 5.4 represent the experimental variables for the test series. It should be noted that the reinforcement details shown in both Tables 5.3 and 5.4 are for positive moment regions only and the reinforcement details in negative moment regions for continuous beams and slabs are shown in Tables 5.5 and 5.6, respectively. All specimens in test series A to J were subjected to a sustained load at age 14 days that produced 50% of the moment capacity M u at the critical section. The details of the loads P (see Figure 5.1) required to achieve the desired moments are given in Tables 5.11 to 5.14. In test series E, the tensile strength of concrete is selected as the parametric variable. The concrete tensile strength was varied with the tension softening parameters (see Section 3.6.1 of Chapter 3) held constant and the bond strengths of the specimens were adjusted according to the tensile strength using a relationship obtained by combining Eq. 2.1 of Chapter 2 and the definition of max given in Table 3.1 of Chapter 3, that is, max = 5 f ct .

247

Table 5.3 - Combinations of variables for beam specimens (simply supported and continuous specimens).
Test No. B-A0 B-A1 B-A2 B-A3 B-B1 B-B2 B-C1 B-C2 B-D1 B-D2 B-E1 B-E2 B-E3 B-E4 B-F1 B-F2 B-F3 B-G1 B-G2 B-G3 B-G4 B-H1 B-H2 B-H3 B-H4 B-I1 B-I2 B-I3 B-J1 B-J2 B-J3

cb (mm) 20 35 50 70 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20

st (mm) 2N20 2N20 2N20 2N20 2N16 2N24 3N20 4N20 2N20 2N20 2N20 2N20 2N20 2N20 2N20 2N20 2N20 2N20 2N20 2N20 2N20 2N20 3N20 2N20 3N20 2N20 2N20 2N20 2N20 3N20 4N20

Ast (mm2) 620 620 620 620 400 900 930 1240 620 620 620 620 620 620 620 620 620 620 620 620 620 620 620 620 620 620 620 620 620 930 1240

Asc (mm2) 0 0 0 0 0 0 0 0 330 (3N12) 620 (2N20) 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

fct (MPa) 2 2 2 2 2 2 2 2 2 2 1 1.5 2.5 3 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2

max (MPa)
10 10 10 10 10 10 10 10 10 10 5 7.5 12.5 15 3 5 15 10 10 10 10 10 10 10 10 10 10 10 10 10 10

lfct (%) 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 1 5 15 20 10 10 10 10 10 10 10 10 10 10

*
2.2 2.2 2.2 2.2 2.2 2.2 2.2 2.2 2.2 2.2 2.2 2.2 2.2 2.2 2.2 2.2 2.2 2.2 2.2 2.2 2.2 0.3 3.9 0.3 3.9 2.2 2.2 2.2 2.2 2.2 2.2

sh* () 600 600 600 600 600 600 600 600 600 600 600 600 600 600 600 600 600 600 600 600 600 0 0 600 600 0 300 900 600 600 600

Bond creep? Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes No No No

Reinforcement details for positive moment regions only.

* = 0.3 and * = 3.9 were achieved by setting q 4 as 0 and 40 /MPa, respectively.

248

Table 5.4 - Combinations of variables for slab specimens (simply supported and continuous specimens).
Test No. S-A0 S-A1 S-A2 S-A3 S-B1 S-B2 S-C1 S-C2 S-D1 S-D2 S-E1 S-E2 S-E3 S-E4 S-F1 S-F2 S-F3 S-G1 S-G2 S-G3 S-G4 S-H1 S-H2 S-H3 S-H4 S-I1 S-I2 S-I3 S-J1 S-J2 S-J3

cb (mm) 20 35 50 70 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20 20

st (mm) 5N16 5N16 5N16 5N16 5N12 5N20 3N16 8N16 5N16 5N16 5N16 5N16 5N16 5N16 5N16 5N16 5N16 5N16 5N16 5N16 5N16 5N16 5N16 5N16 5N16 5N16 5N16 5N16 3N16 5N16 8N16

Ast (mm2) 1000 1000 1000 1000 550 1550 600 1600 1000 1000 1000 1000 1000 1000 1000 1000 1000 1000 1000 1000 1000 1000 1000 1000 1000 1000 1000 1000 600 1000 1600

Asc (mm2) 0 0 0 0 0 0 0 0 600 (3N16) 1000 (5N16) 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

fct (MPa) 2 2 2 2 2 2 2 2 2 2 1 1.5 2.5 3 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2 2

max (MPa)
10 10 10 10 10 10 10 10 10 10 5 7.5 12.5 15 3 5 15 10 10 10 10 10 10 10 10 10 10 10 10 10 10

lfct (%) 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 1 5 15 20 10 10 10 10 10 10 10 10 10 10

*
2.2 2.2 2.2 2.2 2.2 2.2 2.2 2.2 2.2 2.2 2.2 2.2 2.2 2.2 2.2 2.2 2.2 2.2 2.2 2.2 2.2 0.3 3.9 0.3 3.9 2.2 2.2 2.2 2.2 2.2 2.2

sh* () 600 600 600 600 600 600 600 600 600 600 600 600 600 600 600 600 600 600 600 600 600 0 0 600 600 0 300 900 600 600 600

Bond creep? Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes Yes No No No

Reinforcement details for positive moment regions only.

* = 0.3 and * = 3.9 were achieved by setting q 4 as 0 and 40 /MPa, respectively.

249

Table 5.5 - Reinforcement details for continuous beam specimens.


Test No. B-A0 to A3; B-E1 to E4; B-F1 to F3; B-G1 to G4; B-H1 to H4; B-I1 to I3; B-J1; B-K1 to K3. B-B1 B-B2 B-C1; B-J2. B-C2; B-J3. B-D1 B-D2 Positive moment region Ast (mm2) Asc (mm2) 620 (2N20) 0 Negative moment region Ast- (mm2) Asc- (mm2) 1350 (3N24) 0

400 (2N16) 900 (2N24) 930 (3N20) 1240 (4N20) 620 (2N20) 620 (2N20)

0 0 0 0 330 (3N12) 620 (2N20)

800 2040 2250 3060 1350 1350

(4N16) (2N36) (5N24) (3N36) (3N24) (3N24)

0 0 0 0 620 (2N20) 1350 (3N24)

Table 5.6 - Reinforcement details for continuous slab specimens.


Test No. S-A0 to A3; S-E1 to E4; S-F1 to F3; S-G1 to G4; S-H1 to H4; S-I1 to I3; S-J2. S-B1 S-B2 S-C1; S-J1. S-C2: S-J3. S-D1 S-D2 Positive moment region Ast (mm2) Asc (mm2) 1000 (5N16) 0 Negative moment region Ast- (mm2) Asc- (mm2) 2200 (11N16) 0

550 (5N12) 1550 (5N20) 600 (3N16) 1600 (8N16) 1000 (5N16) 1000 (5N16)

0 0 0 0 600 (3N16) 1000 (5N16)

1100 3720 1200 3720 2200 2200

(10N12) (12N20) (6N16) (12N20) (11N16) (11N16)

0 0 0 0 1000 (5N16) 2200 (11N16)

5.2.3.2

Test Series K: Amount of Shear Reinforcement

This series of tests aims to investigate the influence of the amount of shear reinforcement on time-dependent cracking in reinforced concrete beams. The tests were conducted on the continuous beam specimens only, which are subjected to high shear stresses over the shear span. The continuous beam of test No. B-A0 (see Tables 5.3 and 5.5) was selected as the base specimen and the specimens were analysed with various amounts of shear reinforcement. The specimens were loaded to 50% of their moment capacity M u at the critical section at age 14 days. Thereafter the load was kept constant throughout the test (refer to Table 5.13 for magnitude of load P). The details

250

of the shear reinforcement are shown in Table 5.7 and the following parameters were kept unchanged: cb = 20 mm, f ct = 2 MPa, max = 10 MPa, l fct = 10% and bond
* creep, creep ( * = 2.2) and shrinkage ( sh = 600 ) were accounted for.

Table 5.7 - Details of shear reinforcement for test series K (continuous beams).
Test No. B-K1 B-K2 B-K3 Stirrup at spacing s (mm) No stirrup 2-legged N10 at 240 2-legged N10 at 120 Asv/s (mm) 0 0.6667 1.3333

5.2.3.3

Test Series L: Impact of 500 MPa Steel Reinforcement

To investigate the impact of the introduction of the 500 MPa high yield reinforcing steel to replace the previously used 400 MPa reinforcing steel, the specimens were designed as under-reinforced members using both the old and the new types of steel. The beam specimens were designed to resist ultimate moments of 130 kNm and 250 kNm and the slabs specimens to resist ultimates moments of 115 kNm and 145 kNm. Only simply supported specimens were tested in this series of test. The specimens tested were subjected to a load history identical to that for test series A to K, that is, a moment of 50% of the moment capacity M u of the critical sections was applied at age 14 days and held constant throughout the test. The loads P to produce these moments are given in Tables 5.11 and 5.12. The details of this test series are shown in Tables 5.8 and 5.9 for beam and slab specimens, respectively, and the following parameters were held constant: cb = 20 mm, Asc = 0, f ct = 2 MPa, max = 10 MPa and l fct = 10%. Bond creep, creep ( * = 2.2)
* and shrinkage ( sh = 600 ) were included.

Furthermore, the same bar size was used in each specimen.

251

Table 5.8 - Combinations of variables for test series L (simply supported beams).
Test No. B-L1 B-L2 B-L3 B-L4 Mu (kNm) 130 130 250 250 fsy (MPa) 400 500 400 500 st (mm) 4N16 3N16 5N20 4N20 Ast (mm2) 800 600 1550 1240

Table 5.9 - Combinations of variables for test series L (simply supported slabs).
Test No. S-L1 S-L2 S-L3 S-L4 Mu (kNm) 115 115 145 145 fsy (MPa) 400 500 400 500 st (mm) 5N20 4N20 10N16 8N16 Ast (mm2) 1550 1240 2000 1600

5.2.3.4

Test Series M: Load Histories

Five types of load histories as shown in Figures 5.8a to 5.8e were adopted in the numerical experiments. Each of the loading histories was applied up to age 1000 days. Load history LH-1 as presented in Figure 5.8a was applied to test series A to K as described in the previous sections. The remaining load histories were applied only to simply supported beam and slab specimens with the following material parameters held constant: cb = 20 mm , Asc = 0, f ct = 2 MPa, max = 10 MPa , l fct = 10% and
* inclusive of bond creep, creep ( * = 2.2) and shrinkage ( sh = 600 ) . The

reinforcement details of these specimens are as shown in Table 5.10. Load history LH-2 is similar to LH-1 but with a lower sustained moment equal to 35% of the moment capacity M u of the specimens. Load histories LH-3 and LH-4 are designed to examine the effects of unloading and age of unloading on time-dependent cracking. The specimens were loaded up to 0.5M u at age 14 days and then unloaded (with specimens only subjected to moments due to their self-weights M sw ) at age 50 days and 200 days for load histories LH-3 and LH-4, respectively. For load history

252

Moment
LH-1

Moment
LH-2

0.5Mu 0.35Mu

14 Moment 0.5Mu

Time (days)

1000

14 Moment 0.5Mu

(a)

Time (days)

1000

(b)

LH-3

LH-4

Msw 14 50 1000 Time (days)

Msw 14

(c)

200 1000 Time (days)

(d)

Moment
LH-5

0.5Mu 0.35Mu

14

Time (days)

1000

(e)

Fig. 5.8 - Load histories for simply supported beam and slab specimens: (a) LH-1; (b) LH-2; (c) LH-3; (d) LH-4; (e) LH-5. LH-5, the specimens were first loaded up to 0.5M u and then immediately unloaded to 0.35M u . The load was then held constant up to age 1000 days. The magnitudes of the load P to achieve the desired moments are given in Tables 5.11 and 5.12.

253

Table 5.10 - Reinforcement details for test series M (simply supported specimens).
Test No. BEAM specimens B-M1 B-M2 B-M3 SLAB specimens S-M1 S-M2 S-M3 st (mm) 2N20 3N20 4N20 3N16 5N16 8N16 Ast (mm2) 620 930 1240 600 1000 1600

Table 5.11 - Magnitudes of external loads for simply supported beams.


Test No. SIMPLY SUPPORTED BEAMS B-A0 to A3; B-D1 to D2; B-E1 to E4; B-F1 to F3; B-G1 to G4; B-H1 to H4; B-I1 to I3; B-J1. B-B1 B-B2 B-C1; B-J2. B-C2; B-J3. B-L1 B-L2 B-L3 B-L4 B-M1 B-M2 B-M3

Reinforcement area Ast (mm2) 620 (2N20)

Moment M (kNm) 0.50Mu

Load P (kN) 32.9

400 (2N16) 900 (2N24) 930 (3N20) 1240 (4N20) 800 (4N16) 600 (3N16) 1550 (5N20) 1240 (4N20) 620 (2N20) 930 (3N20) 1240 (4N20)

0.50Mu 0.50Mu 0.50Mu 0.50Mu 0.50Mu 0.50Mu 0.50Mu 0.50Mu 0.50Mu 0.35Mu 0.50Mu 0.35Mu 0.50Mu 0.35Mu

19.4 49.3 51.0 67.9 32.0 32.0 67.9 67.9 32.9 21.1 51.0 33.7 67.9 45.5

Reinforcement details for positive moment regions only.

254

Table 5.12 - Magnitudes of external loads for simply supported slabs.


Test No. SIMPLY SUPPORTED SLABS S-A0 to A3; S-D1 to D2; S-E1 to E4; S-F1 to F3; S-G1 to G4; S-H1 to H4; S-I1 to I3; S-J2. S-B1 S-B2 S-C1; S-J1. S-C2; S-J3. S-L1 S-L2 S-L3 S-L4 S-M1 S-M2 S-M3

Reinforcement area Ast (mm2) 1000 (5N16)

Moment M (kNm) 0.50Mu

Load P (kN) 18.1

550 (5N12) 1550 (5N20) 600 (3N16) 1600 (8N16) 1550 (5N20) 1240 (4N20) 2000(10N16) 1600 (8N16) 600 (3N16) 1000 (5N16) 1600 (8N16)

0.50Mu 0.50Mu 0.50Mu 0.50Mu 0.50Mu 0.50Mu 0.50Mu 0.50Mu 0.50Mu 0.35Mu 0.50Mu 0.35Mu 0.50Mu 0.35Mu

6.5 31.1 7.3 33.3 23.4 23.4 32.4 32.4 7.3 2.1 18.1 9.7 33.3 20.3

Reinforcement details for positive moment regions only.

Table 5.13 - Magnitudes of external loads for continuous beams.


Test No. CONTINUOUS BEAMS B-A0 to A3; B-D1 to D2; B-E1 to E4; B-F1 to F3; B-G1 to G4; B-H1 to H4; B-I1 to I3; B-J1; B-K1 to K3. B-B1 B-B2 B-C1; B-J2. B-C2; B-J3.

Reinforcement area Ast (mm2) 620 (2N20)

Moment M (kNm) 0.5Mu

Load P (kN) 112.0

400 (2N16) 900 (2N24) 930 (3N20) 1240 (4N20)

0.5Mu 0.5Mu 0.5Mu 0.5Mu

71.5 161.0 166.1 216.8

Reinforcement details for positive moment regions only.

255

Table 5.14 - Magnitudes of external loads for continuous slabs.


Test No. CONTINUOUS SLABS S-A0 to A3; S-D1 to D2; S-E1 to E4; S-F1 to F3; S-G1 to G4; S-H1 to H4; S-I1 to I3; S-J2. S-B1 S-B2 S-C1; S-J1. S-C2; S-J3.

Reinforcement area Ast (mm2) 1000 (5N16)

Moment M (kNm) 0.5Mu

Load P (kN) 73.6

550 (5N12) 1550 (5N20) 600 (3N16) 1600 (8N16)

0.5Mu 0.5Mu 0.5Mu 0.5Mu

36.8 115.5 41.0 119.1

Reinforcement details for positive moment regions only.

5.3 Presentation and Discussion of Results


The presentation of the results for the numerical experiments is organized in the same sequence as the test series. The numerical test results for all specimens, including the simply supported beams, the simply supported slabs, the continuous beams and the continuous slabs, are shown in groups so that a direct comparison of the results for all types of specimens can be made conveniently. The numerical test results are shown in terms of widths and spacings of cracks within the constant moment region (region between the two point loads, see Figure 5.1). Average crack widths were calculated by averaging the widths of the primary cracks (i.e. crack widths of cracked elements) within the constant moment region. Crack widths are determined at the soffit of the specimen. In the presentation of the results, only average and maximum crack widths at age 1000 days (final crack widths), are shown. In addition, the ratios of final crack width to crack width at instantaneous loading (initial crack width) are presented to show the change in width of cracks with time.

256

5.3.1

Test Series A Bottom Concrete Cover

Figure 5.9 shows the final crack widths and the crack spacings for the specimens analysed taking bottom concrete cover as the variable parameter. It is evident that the thickness of the bottom cover has a pronounced effect on crack spacing and crack width irrespective of the type of structure and the boundary conditions. The cracks are wider as the thickness of the bottom cover increases. It is seen that the beam and slab specimens have very similar trends for both final crack width and crack spacing at various bottom covers. The boundary conditions of the specimens do not seem to have a prominent effect on crack width and crack spacing. In a flexural member, the strain is linearly distributed across the section. For the specimens analysed in this series, the effective depths were held constant both in the positive and negative (if any) moment regions. Consequently, the sectional strain distribution between the extreme compressive fibre and the tensile steel level must be the same for all specimens of the same type. The strain at the extreme tensile fibre depends on the bottom concrete cover of the specimens and so, clearly, the thicker the bottom cover is, the larger the strain at the extreme tensile fibre. Therefore the crack width for sections with thicker bottom cover is always larger since crack width is directly proportional to the strain in the crack opening direction (see Eq. 3.10 of Chapter 3). In addition, the large crack width for sections with thick bottom cover is because the crack spacings are larger. As seen in Eq. 3.10, it is clear that crack width is proportional to crack spacing. It is, however, useful to examine the relations between concrete cover and crack spacing. Taking a reinforced concrete tension chord similar to that discussed in Section 3.3.1, the rate of stress transfer between concrete and steel between two cracks depends on the bond stress, the amount of reinforcement and the area of concrete. From Eq. 3.2 of Chapter 3, which is derived from equilibrium across the tension chord, it can be seen that the stress transfer rate between concrete and steel is lower if the concrete area of the tension chord is large. In other words, a tension chord with low reinforcement ratio has a larger development length of bond (or weaker bond), which means a longer distance is required to form the next crack and hence a

257

0.70

Final crack width (mm)

Crack spacing (mm)

0.60 0.50 0.40 0.30 0.20 0.10 0.00 0

Simply supported beams

300 250 200 150 100 50 0

Simply supported beams

Avg Max

Instantaneous t=1000 days

20 40 60 80 Bottom concrete cover, cb (mm)

(a)

20 40 60 80 Bottom concrete cover, cb (mm)

(b)

0.70

Final crack width (mm)

Crack spacing (mm)

0.60 0.50 0.40 0.30 0.20 0.10 0.00 0

Continuous beams

300 250 200 150 100 50 0

Continuous beams

Avg Max

Instantaneous t=1000 days

20

40

60

80

Bottom concrete cover, cb (mm)

(c)

20 40 60 80 Bottom concrete cover, cb (mm)

(d)

0.70

Final crack width (mm)

Crack spacing (mm)

0.60 0.50 0.40 0.30 0.20 0.10 0.00 0

Simply supported slabs

300 250 200 150 100 50 0

Simply supported slabs

Avg Max

Instantaneous t=1000 days

20 40 60 80 Bottom concrete cover, cb (mm)

(e)

20 40 60 80 Bottom concrete cover, cb (mm)

(f)

0.70

Final crack width (mm)

Crack spacing (mm)

0.60 0.50 0.40 0.30 0.20 0.10 0.00 0

Continuous slabs

300 250 200 150 100 50 0

Continuous slabs

Avg Max

Instantaneous t=1000 days

(h) (g) Fig. 5.9 - Final crack widths and crack spacings for test series A: (a) and (b) simply supported beams; (c) and (d) continuous beams; (e) and (f) simply supported slabs; (g) and (h) continuous slabs.
258

20 40 60 80 Bottom concrete cover, cb (mm)

20 40 60 80 Bottom concrete cover, cb (mm)

larger crack spacing. Moreover, weaker bond also indicates larger slips in the vicinity of the cracks which causes the cracks to open more widely. In the case of a flexural member, the tension chord analogy can be applied to the tensile fibre containing the tensile reinforcement. For a fixed quantity of tensile reinforcement, the spacing of cracks is larger for a bigger effective area of concrete in tension surrounding the reinforcing steel since the effective reinforcement ratio (area of tensile steel divided by effective area of concrete in tension) in the tension region is low. In both the provisions of CEB-FIP Model Code 1990 (1993) and Eurocode 2 (1992), the effective tensile area of concrete is taken as the width of the section times a depth equal to 2.5 times the distance from the centroid of the reinforcement to the tensile face of the section. Therefore, the effective tensile area of concrete is larger with thicker bottom cover and as a result, both crack spacing and crack width will be larger. In addition, the crack spacings and crack widths appear to increase at a decreasing rate as the bottom cover increases. It seems that the crack spacing is approaching an upper bound limit no matter how thick the concrete cover is. After a detailed investigation of the finite element results, it is found that the bottom cover ceases to affect the effective area of concrete in tension after a threshold bottom cover thickness is reached. This can be explained by examining the stress distribution of concrete between any two adjacent cracks in the constant moment region, as shown in Figure 5.10. Concrete stress developed around the surface of the reinforcing steel (at 20 mm

Distance from bottom (mm)

cb = 20 mm

Distance from bottom (mm)

500 400 300 200 100 0 -8 -6 -4 -2 Effective depth of concrete in tension


Instantaneous t=1000 days

500 400 300 200 100 0 -8 -6 -4 -2 Effective depth of concrete in tension


Instantaneous t=1000 days

cb = 70 mm

Concrete stress betw cracks (MPa)

(a)

Concrete stress betw cracks (MPa)

(b)

Fig. 5.10 - Typical stress distribution of section between cracks for the simply supported beam specimen with: (a) cb = 20 mm ; (b) cb = 70 mm .
259

and at 70 mm above the soffit of the specimens) through stress transfer via bond. It can be seen in Figure 5.10a that the bottom cover has limited the potential development of stress in concrete, which basically results in a smaller effective area of concrete in tension. For the case of cb = 70 mm as shown in Figure 5.10b, the bottom cover is sufficiently thick to allow the full development of tensile stress in concrete. This means the bottom cover has reached a value where further increase will no longer affect the effective tensile area of concrete and consequently the crack spacing. The ratios of final to initial crack width are plotted against bottom cover thickness for all specimens and are shown in Figure 5.11. It is seen that the time-dependent percentage increase in crack width is not affected by the thickness of bottom cover. In addition, the continuous specimens have a slightly larger percentage increase in crack width with time compared to the simply supported specimens.

Crack width ratio, wcr.f/wcr.i

Simply supported beams

Crack width ratio, wcr.f/wcr.i

2.50 2.00 1.50 1.00 0.50 0.00 0

6.00 5.00 4.00 3.00 2.00 1.00 0.00 0

Continuous beams

Avg Max

Avg Max

20 40 60 80 Bottom concrete cover, cb (mm)

(a)

20 40 60 80 Bottom concrete cover, cb (mm)

(b)

Crack width ratio, wcr.f/wcr.i

Crack width ratio, wcr.f/wcr.i

3.00 2.50 2.00 1.50 1.00 0.50 0.00 0

Simply supported slabs

3.00 2.50 2.00 1.50 1.00 0.50 0.00 0

Continuous slabs

Avg Max

Avg Max

20 40 60 80 Bottom concrete cover, cb (mm)

(c)

20 40 60 80 Bottom concrete cover, cb (mm)

(d)

Fig. 5.11 - Ratios of final to initial crack width for test series A: (a) simply supported beams; (b) continuous beams; (c) simply supported slabs; (d) continuous slabs.

260

In Figure 5.11b, there is a data point for the widest crack that is particularly higher than the rest of the data points. This is because the crack that has the largest width at age 1000 days was not the widest crack at first loading. With time, the crack opened extensively and became the widest crack.

5.3.2

Test Series B Diameter of Tensile Reinforcing Steel

In this series of test, the effects of the diameter of tensile reinforcing bars are investigated. All specimens had the same bottom cover, the same spacing between bars and the same effective depth. Although the spacing between reinforcing bars cannot be modelled by a two-dimensional finite element model, it can be indirectly simulated by using the same number of bars for each specimen and assume that the bars are spaced sufficiently far to develop full bonding between steel and concrete. In addition, the specimens were loaded such that the reinforcing steels are resisting the same level of stress. The results for final crack width and crack spacing are plotted against bar diameter and are shown in Figure 5.12. Crack widths are smaller if reinforcing bars of larger size are used and this applies to all specimen types and boundary conditions. The crack spacings are also in the same trend as the crack widths. At first glance, one may be surprised by this result. It seems at odds with the detailing rules for crack control and conventional wisdom, which is to avoid the use of a small number of large diameter bars and instead, use a larger number of small diameters bars. In this test, however, the same number of bars has been used in each specimen. Examining the results closely, it can be noticed that the total contact surface of the bars and concrete is in fact the primary factor affecting the distribution of cracks and crack widths. Better bonding can be achieved by providing larger contact surface area of concrete and steel. With good bonding, the load transfer rate between concrete and steel will be higher and therefore cracks form at a closer spacing and are smaller in width. Since the number of bars were the same for each specimen, the bar diameter becomes the only variable

261

0.50 0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00

Final crack width (mm)

Crack spacing (mm)

Simply supported beams

300 250 200 150 100 50 0

Simply supported beams

Avg Max

Instantaneous t=1000 days

10

15 20 25 Bar diameter, st (mm)

30

10

(a)

15 20 25 Bar diameter, st (mm)

30

(b)

0.50 0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00

Final crack width (mm)

Crack spacing (mm)

Continuous beams

300 250 200 150 100 50 0

Continuous beams

Avg Max

Instantaneous t=1000 days

10

15

20

25

30

10

Bar diameter, st (mm)

(c)

15 20 25 Bar diameter, st (mm)

30

(d)

0.50 0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00

Final crack width (mm)

Crack spacing (mm)

Simply supported slabs

300 250 200 150 100 50 0

Simply supported slabs

Avg Max

Instantaneous t=1000 days

10

15 20 Bar diameter, st (mm)

25

10

(e)

15 20 Bar diameter, st (mm)

25

(f)

0.50 0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00

Final crack width (mm)

Crack spacing (mm)

Continuous slabs

300 250 200 150 100 50 0

Continuous slabs

Avg Max

Instantaneous t=1000 days

10

15

20

25

10

(g) (h) Fig. 5.12 - Final crack widths and crack spacings for test series B: (a) and (b) simply supported beams; (c) and (d) continuous beams; (e) and (f) simply supported slabs; (g) and (h) continuous slabs.
262

Bar diameter, st (mm)

15 20 Bar diameter, st (mm)

25

affecting the concrete-steel contact surface area. Consequently, the bond between concrete and steel is better for reinforcing steel with larger diameter. This clearly explains why the spacing and the width of cracks are smaller for specimens containing larger bars. Some design codes stipulate a maximum permissible spacing of tensile reinforcement to control cracking in reinforced concrete structures. As a consequence, it is commonly thought that the spacing of reinforcing steel is one of the factors affecting crack spacing and crack width. However, in the authors view the concretesteel contact surface area is the primary determinant. As long as the reinforcing steels are spaced sufficiently far apart for the development of optimal bond, the bar spacing is important only in terms of limiting the number of large tensile bars detailed in a section. By limiting the maximum bar spacing, structural designers must resort to using reinforcing steels of smaller sizes in order to provide the required amount of reinforcement. This indirectly provides more concrete-steel contact surface area and results in better bonding to distribute cracks. The ratio of final to initial crack width is shown in Figure 5.13. The bar diameter appears to have no influence on the crack width ratio. The amount of time-dependent crack opening is generally larger for the continuous specimens than for the simply supported specimens. However, the difference is only slight. For the simply supported slab specimens, it is seen that the crack width ratio calculated from average crack widths for the specimens containing 5N12 reinforcement (test S-B1) is significantly higher than other data points (see Figure 5.13c). This is because this slab specimen was quite lightly reinforced (reinforcement ratio of 0.275%) and the 0.5M u external moment applied to the specimen only just exceeded the cracking moment. At first loading the cracks were quite fine. The cracks opened extensively with time and consequently a high crack width ratio.

263

Crack width ratio, wcr.f/wcr.i

Crack width ratio, wcr.f/wcr.i

2.50 2.00 1.50 1.00 0.50 0.00

Simply supported beams

2.50 2.00 1.50 1.00 0.50 0.00

Continuous beams

Avg Max

Avg Max

10

15 20 25 Bar diameter, st (mm)

30

10

(a)

15 20 25 Bar diameter, st (mm)

30

(b)

Crack width ratio, wcr.f/wcr.i

Crack width ratio, wcr.f/wcr.i

4.00 3.50 3.00 2.50 2.00 1.50 1.00 0.50 0.00

Simply supported slabs

4.00 3.50 3.00 2.50 2.00 1.50 1.00 0.50 0.00

Continuous slabs

Avg Max

Avg Max

10

15

20

25

10

Bar diameter, st (mm)

(c)

15 20 Bar diameter, st (mm)

25

(d)

Fig. 5.13 - Ratios of final to initial crack width for test series B: (a) simply supported beams; (b) continuous beams; (c) simply supported slabs; (d) continuous slabs.

5.3.3

Test Series C Quantity of Tensile Reinforcement

The effects of the amount of tensile reinforcement are investigated in this series of tests. Only one single bar size was used for each type of specimen and various reinforcement ratios were achieved by varying the number of reinforcing bars. The results for final crack width and crack spacing are plotted against reinforcement ratio and are shown in Figure 5.14. It is observed that both final crack width and crack spacing decrease with increasing reinforcement ratio. The results for the simply supported and continuous beam specimens have very similar trends. The reduction in the crack width and crack spacing with increasing reinforcement ratio is in fact, again, due to the increasing contact surface area between the concrete and the reinforcing steel. Since the bar size was held constant, the number of bars

264

0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00

Final crack width (mm)

Crack spacing (mm)

Simply supported beams

250 200 150 100 50

Simply supported beams

Avg Max

Instantaneous t=1000 days

0.002

0.004 0.006 0.008 Reinforcement ratio,

0.01

0 0.002

(a)

0.004 0.006 0.008 Reinforcement ratio,

0.01

(b)

0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00

Final crack width (mm)

Crack spacing (mm)

Continuous beams

250 200 150 100 50

Continuous beams

Avg Max

Instantaneous t=1000 days

0.002

0.004 0.006 0.008 Reinforcement ratio,

0.01

0 0.002

(c)

0.004 0.006 0.008 Reinforcement ratio,

0.01

(d)

0.60

Final crack width (mm)

0.50 0.40 0.30 0.20 0.10 0.00 0.002 0.004 0.006 0.008 Reinforcement ratio, 0.01
Avg Max

Crack spacing (mm)

Simply supported slabs

400 350 300 250 200 150 100 50

Simply supported slabs

Instantaneous t=1000 days

0 0.002

(e)

0.004 0.006 0.008 Reinforcement ratio,

0.01

(f)

0.60

Final crack width (mm)

0.50 0.40 0.30 0.20 0.10 0.00 0.002 0.004 0.006 0.008 Reinforcement ratio, 0.01
Avg Max

Crack spacing (mm)

Continuous slabs

400 350 300 250 200 150 100 50

Continuous slabs

Instantaneous t=1000 days

0 0.002

(g) (h) Fig. 5.14 - Final crack widths and crack spacings for test series C: (a) and (b) simply supported beams; (c) and (d) continuous beams; (e) and (f) simply supported slabs; (g) and (h) continuous slabs.
265

0.004 0.006 0.008 Reinforcement ratio,

0.01

is the main factor that governs the total contact surface area between concrete and steel. As discussed earlier in Section 5.3.2, cracks are better distributed with larger concretesteel contact surface area. When the reinforcement ratio increases, with the same bar size, the total concrete-steel contact surface area also increases proportionally, and therefore crack widths and crack spacings are smaller. Figure 5.15 shows the ratios of final to initial crack width for all specimens of test series C. The percentage increase in crack width with time is in the range of 40% to 90% of the instantaneous crack width and it is seen that the quantity of reinforcement does not significantly affect this crack width ratio.

2.00 1.80 1.60 1.40 1.20 1.00 0.80 0.60 0.40 0.20 0.00

Simply supported beams

Avg Max

0.002

0.004 0.006 0.008 Reinforcement ratio,

0.01

(a)

2.00 1.80 Continuous beams 1.60 1.40 1.20 1.00 0.80 0.60 Avg 0.40 Max 0.20 0.00 0.002 0.004 0.006 0.008 Reinforcement ratio,

Crack width ratio, wcr.f/wcr.i

Crack width ratio, wcr.f/wcr.i

0.01

(b)

Crack width ratio, wcr.f/wcr.i

Simply supported slabs

Crack width ratio, wcr.f/wcr.i

2.50 2.00 1.50 1.00 0.50 0.00

2.50 2.00 1.50 1.00 0.50

Continuous slabs

Avg Max

Avg Max

0.002

0.004 0.006 0.008 Reinforcement ratio,

0.01

0.00 0.002

(c)

0.004 0.006 0.008 Reinforcement ratio,

0.01

(d)

Fig. 5.15 - Ratios of final to initial crack width for test series C: (a) simply supported beams; (b) continuous beams; (c) simply supported slabs; (d) continuous slabs.

266

5.3.4

Test Series D Quantity of Compressive Reinforcement

The results for final crack width and crack spacing of this test series are shown in Figure 5.16. The amount of compressive reinforcement is presented as ratios of compressive reinforcement area to tensile reinforcement area. From many previous studies on time-dependent behaviour of reinforced concrete structures (for example, Gilbert, 1979), it is well known that the use of compressive reinforcement is important in controlling long-term deflection in flexural members. However, it is seen from the results that the quantity of compressive reinforcement has no pronounced effect on crack width and crack spacing. For a reinforced concrete structure subjected to long-term loads, the compressive reinforcement restrains the concrete from undergoing excessive creep deformation in the compression zone of the section. This limits the increase of curvature with time and consequently reduces the long-term deflection. In addition, the presence of compressive reinforcement enables a further relief of the compressive stress in the concrete as creep develops with time. This is because a significant portion of the compressive force is gradually transferred from the concrete to the compressive reinforcement. Therefore after a period of long-term loading, the depth of the neutral axis for a cracked member containing compressive reinforcement is smaller than that without or with less compressive reinforcement as illustrated in Figure 5.17. Despite the reduction in creep deformation in the compression zone as well as the smaller final depth of neutral axis, the resultant influence to the strain in the tensile reinforcement is insignificant. This is also shown schematically in Figure 5.17. As a result, although the presence of compressive reinforcement is important to the longterm deflection of reinforced concrete structures, it is not important with regard to the spacing and the opening of cracks. Figure 5.18 shows the ratios of the final to initial crack width of the specimens. The ratios are in the same range as those of the previously presented test series. Moreover, the percentage increase of crack width under sustained loads is not sensitive to the quantity of compressive reinforcement.

267

0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0

Final crack width (mm)

Crack spacing (mm)

Simply supported beams

300 250 200 150 100 50 0

Simply supported beams

Avg Max

Instantaneous t=1000 days

0.5 1 1.5 Compression steel quantity, Asc/Ast

(a)

0.5 1 1.5 Compression steel quantity, Asc/Ast

(b)

0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0

Final crack width (mm)

Crack spacing (mm)

Continuous beams

300 250 200 150 100 50 0

Continuous beams

Avg Max

Instantaneous t=1000 days

0.5

1.5

Compression steel quantity, Asc/Ast

(c)

0.5 1 1.5 Compression steel quantity, Asc/Ast

(d)

0.50 0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0

Final crack width (mm)

Crack spacing (mm)

Simply supported slabs

300 250 200 150 100 50 0

Simply supported slabs

Avg Max

Instantaneous t=1000 days

0.5 1 1.5 Compression steel quantity, Asc/Ast

(e)

0.5 1 1.5 Compression steel quantity, Asc/Ast

(f)

0.50 0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0

Final crack width (mm)

Crack spacing (mm)

Continuous slabs

300 250 200 150 100 50 0

Continuous slabs

Avg Max

Instantaneous t=1000 days

0.5

1.5

(g) (h) Fig. 5.16 - Final crack widths and crack spacings for test series D: (a) and (b) simply supported beams; (c) and (d) continuous beams; (e) and (f) simply supported slabs; (g) and (h) continuous slabs.
268

Compression steel quantity, Asc/Ast

0.5 1 1.5 Compression steel quantity, Asc/Ast

sc

sc

dNA

dNA

st

st

(a)
Instantaneous strain Final strain

(b)
Instantaneous steel strain Steel strain increment over time

Fig. 5.17 - Schematic

description

of

the

long-term

effects

of

compressive

reinforcement on tensile steel strain: (a) section without compressive reinforcement; (b) section with compressive reinforcement.

Crack width ratio, wcr.f/wcr.i

Simply supported beams

Crack width ratio, wcr.f/wcr.i

2.50 2.00 1.50 1.00 0.50 0.00 0

2.50 2.00 1.50 1.00 0.50 0.00 0

Continuous beams

Avg Max

Avg Max

0.5

1.5

Compression steel quantity, Asc/Ast

(a)

0.5 1 1.5 Compression steel quantity, Asc/Ast

(b)

Crack width ratio, wcr.f/wcr.i

Simply supported slabs

Crack width ratio, wcr.f/wcr.i

2.50 2.00 1.50 1.00 0.50 0.00 0

2.50 2.00 1.50 1.00 0.50 0.00 0

Continuous slabs

Avg Max

Avg Max

0.5

1.5

Compression steel quantity, Asc/Ast

(c)

0.5 1 1.5 Compression steel quantity, Asc/Ast

(d)

Fig. 5.18 - Ratios of final to initial crack width for test series D: (a) simply supported beams; (b) continuous beams; (c) simply supported slabs; (d) continuous slabs.
269

5.3.5

Test Series E Tensile Strength of Concrete

The bond strength between concrete and steel is closely related to the properties of concrete, such as compressive strength and tensile strength. This is evident in many bond models, which were formulated as a function of either the compressive strength or the tensile strength of concrete (for instance, Drr, 1980 and Ciampi et al., 1981). For the same type of deformed bar the bond strength increases with increasing concrete tensile strength (or concrete compressive strength). To achieve a more realistic study of the influence of concrete tensile strength on crack width and crack spacing, the bond strength of each specimen was adjusted accordingly with the concrete tensile strengths using the method described in Section 5.2.3.1. The final crack width and crack spacing versus concrete tensile strength diagrams are shown in Figure 5.19. It should be pointed out that the absence of results for the continuous slab specimens in Figures 5.19g and 5.19h at f ct = 1 MPa is due to the fact that the concrete tensile strength is too low and results in a large shear crack occurring in the negative moment region, which leads to a premature shear failure. In Figure 5.19, the crack spacings at first loading tend to increase with increasing concrete tensile strength. At age 1000 days the crack spacings reduce to a constant final value for concrete tensile strengths between 2.0 MPa and 3.0 MPa. The influence of concrete tensile strength on final crack width is, however, less significant than its influence on crack spacing except for the continuous slab specimens (Figure 5.19g). The crack widths have a slight ascending trend for concrete tensile strengths between 1.0 MPa and 1.5 MPa and remain roughly constant or decrease slightly with further increase in tensile strength. The continuous slab specimens have the most obvious descending final crack widths at high concrete tensile strengths. On the whole, the effects of concrete tensile strength on final crack spacings and final crack widths are not very significant. Both concrete tensile strength and bond strength are important factors affecting the formation of cracks in a reinforced concrete structure. Bond strength is a governing factor of the rate of stress transfer between concrete and steel. With higher bond

270

strength, the stress transfer rate is higher and therefore the distance required to reach the concrete tensile strength is shorter, which means a smaller crack spacing. On the other hand, for a given bond characteristic (or stress transfer rate), higher concrete tensile strength results in larger crack spacing since a relatively long distance is required for the concrete stress to develop to the tensile strength from an existing crack. In this series of tests, bond strength is taken to be directly proportional to the concrete tensile strength of the specimens. Consequently, the effect of bond strength to a large extent cancels out the effect of concrete tensile strength and results in an almost constant final crack spacing especially at tensile strengths between 1.5 MPa and 3.0 MPa. As a consequence, the crack widths are also nearly constant for various concrete tensile strengths except for those in the continuous slab specimens. The decrease of crack widths with increasing concrete tensile strengths as observed in Figure 5.19 is attributed to the fact that the concrete was too strong to crack extensively at instantaneous loading. As the concrete tensile strength increases, a point is reached that the specimens only just crack at the service load levels. Although cracking was successfully initiated in these specimens, the initial crack width was small, but widened significant with time. This can also be seen in Figure 5.20 where the ratios of final to initial crack width are almost constant at low concrete tensile strengths but increase significantly after the concrete tensile strength exceeds about 2.5 MPa. This indicates that the initial crack widths are small compared to the final crack widths. Although cracks in the specimens with high concrete tensile strength opened with time, they did not open as much as the cracks in those specimens with lower tensile strengths which had cracked more extensively at instantaneous loading. Therefore, the final crack width decreases with increasing tensile strengths. The results presented herein for this series of tests were obtained using a particular relationship between the bond strength and the concrete tensile strength. It should be pointed out that if a different relationship is used the results may vary accordingly. However, due to the combined effects of bond strength and concrete tensile strength, the author believes that the final outcome would be similar to that obtained in this investigation, for which the overall effects of variations in the concrete

271

0.50 0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0

Final crack width (mm)

Crack spacing (mm)

Simply supported beams

300 250 200 150 100 50 0

Simply supported beams

Avg Max

Instantaneous t=1000 days

1 2 3 4 Concrete tensile strength, fct (MPa)

(a)

1 2 3 4 Concrete tensile strength, fct (MPa)

(b)

0.50 0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0

Final crack width (mm)

Crack spacing (mm)

Continuous beams

300 250 200 150 100 50 0

Continuous beams

Avg Max

Instantaneous t=1000 days

1 2 3 4 Concrete tensile strength, fct (MPa)

(c)

1 2 3 4 Concrete tensile strength, fct (MPa)

(d)

0.50 0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0

Final crack width (mm)

Crack spacing (mm)

Simply supported slabs

300 250 200 150 100 50 0

Simply supported slabs

Avg Max

Instantaneous t=1000 days

1 2 3 4 Concrete tensile strength, fct (MPa)

(e)

1 2 3 4 Concrete tensile strength, fct (MPa)

(f)

0.50 0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0

Final crack width (mm)

Crack spacing (mm)

Continuous slabs

300 250 200 150 100 50 0

Continuous slabs

Avg Max

Instantaneous t=1000 days

(g) (h) Fig. 5.19 - Final crack widths and crack spacings for test series E: (a) and (b) simply supported beams; (c) and (d) continuous beams; (e) and (f) simply supported slabs; (g) and (h) continuous slabs.
272

1 2 3 4 Concrete tensile strength, fct (MPa)

1 2 3 4 Concrete tensile strength, fct (MPa)

Crack width ratio, wcr.f/wcr.i

Simply supported beams


Avg Max

Crack width ratio, wcr.f/wcr.i

2.50 2.00 1.50 1.00 0.50 0.00 0

2.50 2.00 1.50 1.00 0.50 0.00 0

Continuous beams
Avg Max

1 2 3 4 Concrete tensile strength, fct (MPa)

(a)

1 2 3 4 Concrete tensile strength, fct (MPa)

(b)

4.50 4.00 3.50 3.00 2.50 2.00 1.50 1.00 0.50 0.00 0

Crack width ratio, wcr.f/wcr.i

Simply supported slabs


Avg Max

Crack width ratio, wcr.f/wcr.i

4.00 3.50 3.00 2.50 2.00 1.50 1.00 0.50 0.00 0

Continuous slabs
Avg Max

1 2 3 4 Concrete tensile strength, fct (MPa)

(c)

1 2 3 4 Concrete tensile strength, fct (MPa)

(d)

Fig. 5.20 - Ratios of final to initial crack width for test series E: (a) simply supported beams; (b) continuous beams; (c) simply supported slabs; (d) continuous slabs.

tensile strength are not significant to crack spacings and crack widths. This is evident in the crack spacing equations proposed in many previous studies, for example CEB-FIP Model Code 1990 (1993) and Eurocode 2 (1992), where concrete tensile strength is not taken as a critical factor affecting the spacing of cracks. In addition, although the tension chord model crack spacing equation used in this study (see Eq. 3.4 of Chapter 3) has included concrete tensile strength as one of the governing parameters, the use of a direct proportionality between bond stress and tensile strength in fact cancels out the influence of concrete tensile strength on crack spacings, and hence crack widths.

273

5.3.6

Test Series F Bond Strength between Steel and Concrete

Figure 5.21 shows the final crack widths and crack spacings plotting against the maximum bond capacity of the specimens. The importance of bond between concrete and steel to the formation of cracks has been extensively emphasized throughout this thesis. Therefore, bond strength is well expected to have a significant effect on crack width and crack spacing. Although it is apparent in Figure 5.21 that crack width and crack spacing reduce with increasing maximum bond capacity, it should be noted that it is not the maximum bond capacity that affects the formation of cracks but the characteristics of the of the bond model. The bond-slip model of CEB-FIP Model Code 1990 (1993) has an infinite initial modulus and the gradient of the bond-slip curve reduces with increasing slip. Upon reaching the bond strength, the bond stress descends to a residual frictional bond stress corresponding to a pull-out failure. Under service load conditions, bond stresses are well below the maximum bond capacity. Figure 5.22 shows the bond stress versus slip curves for the CEB-FIP bond models used in this series of tests. It can be seen that bond is stiffer (higher bond stress at the same slip) for the curves with higher bond strength. A stiff bond model has better stress transfer capability and consequently reduces the spacing and width of cracks. Another interesting observation is the difference between maximum and average crack widths at various bond strengths. The difference between maximum and average crack widths is large at low bond strength but reduces as bond strength increases. For specimens with low bond strength, the widest crack has a noticeably larger width compared to other cracks of the same specimens and it tends to occur below the third point loads where the specimens undergo a sudden change in curvature. This is due to the poor distribution of cracking resulting from low bond strength. The large crack spacing causes the cracks to open more widely at locations with a sudden change in curvature such as the third points of the specimens. On the other hand, with high bond strength the cracks are spaced more closely and the opening of cracks due to change in curvature can be easily distributed between the nearby cracks. Consequently, the difference between maximum and average crack width is much smaller.

274

1.20

Final crack width (mm)

1.00 0.80 0.60 0.40 0.20 0.00 0

Max

Crack spacing (mm)

Simply supported beams

350
Avg

300 250 200 150 100 50 0 0

Simply supported beams

Instantaneous t=1000 days

5 10 15 20 Maximum bond capacity, max (MPa)

(a)

5 10 15 20 Maximum bond capacity, max (MPa)

(b)

1.20

Final crack width (mm)

1.00 0.80 0.60 0.40 0.20 0.00 0 5 10 15

Max

Crack spacing (mm)

Continuous beams

350
Avg

300 250 200 150 100 50 0 0

Continuous beams

Instantaneous t=1000 days

20

Maximum bond capacity, max (MPa)

(c)

5 10 15 20 Maximum bond capacity, max (MPa)

(d)

1.20

Final crack width (mm)

1.00 0.80 0.60 0.40 0.20 0.00

Max

Crack spacing (mm)

Simply supported slabs

400
Avg

350 300 250 200 150 100 50 0 0

Simply supported slabs

Instantaneous t=1000 days

0 5 10 15 20 Maximum bond capacity, max (MPa)

(e)

5 10 15 20 Maximum bond capacity, max (MPa)

(f)

1.20

Final crack width (mm)

1.00 0.80 0.60 0.40 0.20 0.00

Max

Crack spacing (mm)

Continuous slabs

400
Avg

350 300 250 200 150 100 50 0 0

Continuous slabs

Instantaneous t=1000 days

(g) (h) Fig. 5.21 - Final crack widths and crack spacings for test series F: (a) and (b) simply supported beams; (c) and (d) continuous beams; (e) and (f) simply supported slabs; (g) and (h) continuous slabs.
275

0 5 10 15 20 Maximum bond capacity, max (MPa)

5 10 15 20 Maximum bond capacity, max (MPa)

16

Bond stress (MPa)

14 12 10 8 6 4 2 0 0 0.5

max = 15 MPa max = 10 MPa max = 5 MPa max = 3 MPa

1.5

Slip (mm)

Fig. 5.22 - CEB-FIP bond models used in test series F.

Figure 5.23 shows the ratios of final to initial crack width plotted against bond strength. Bond strength appears to have little influence on the percentage increase of crack width with time. The slightly higher crack width ratios for the continuous beam

Crack width ratio, wcr.f/wcr.i

3.00 2.50 2.00 1.50 1.00 0.50 0.00

Simply supported beams

Crack width ratio, wcr.f/wcr.i

3.50

3.50 3.00 2.50 2.00 1.50 1.00 0.50 0.00

Continuous beams

Avg Max

Avg Max

0 5 10 15 20 Maximum bond capacity, max (MPa)

(a)

0 5 10 15 20 Maximum bond capacity, max (MPa)

(b)

Crack width ratio, wcr.f/wcr.i

Simply supported slabs

Crack width ratio, wcr.f/wcr.i

3.00 2.50 2.00 1.50 1.00 0.50 0.00 0

3.00 2.50 2.00 1.50 1.00 0.50 0.00 0

Continuous slabs

Avg Max

Avg Max

5 10 15 20 Maximum bond capacity, max (MPa)

(c)

5 10 15 20 Maximum bond capacity, max (MPa)

(d)

Fig. 5.23 - Ratios of final to initial crack width for test series F: (a) simply supported beams; (b) continuous beams; (c) simply supported slabs; (d) continuous slabs.
276

specimens with max = 3 MPa and max = 5 MPa are due to the switch of the largest cracks during the sustained loading period, which is of the same reason as discussed in the last paragraph of Section 5.3.1 for test series A.

5.3.7

Test Series G Concrete Tensile Strength Fluctuation Limit

The randomly generated concrete tensile strengths were incorporated in the finite element model for two reasons. The first reason is to model the stochastic nature of concrete tensile strength. The second is to avoid bifurcation of the numerical solutions. The interests of this test series are twofold: to investigate the effects of the variability of concrete tensile strength for a given mean value and, to examine the effects of the fluctuation limit on the numerical outcomes. The results for final crack width and crack spacing are plotted versus concrete tensile strength fluctuation limit in Figure 5.24. It is seen that the fluctuation limit does not have significant influence on crack width or crack spacing. At low fluctuation limits, the crack widths and crack spacings tend to be slightly smaller than those at high fluctuation limits. This is probably because of the tensile stresses just prior to cracking in the concrete elements at the soffits were so close that more cracks tended to form at the same instance. Similarly for the case using a single concrete tensile strength (l fct = 0%) , one would expect the concrete elements at the soffit to crack at the same time. In addition, it is seen that the effects of fluctuation of concrete tensile strength vanishes at fluctuation limits beyond 10%. The ratios of final to initial crack width are presented in Figure 5.25. The ratios for this test series fall between 1.5 and 2, which are similar to those for the earlier test series. Furthermore, the percentage increase in crack width with time is unaffected by the concrete tensile strength fluctuation limit.

277

0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0

Final crack width (mm)

Crack spacing (mm)

Simply supported beams

250 200 150 100 50 0

Simply supported beams

Avg Max

Instantaneous t=1000 days

5 10 15 20 25 Concrete tensile strength limit, lfct (%)

(a)

5 10 15 20 25 Concrete tensile strength limit, lfct (%)

(b)

0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00

Final crack width (mm)

Crack spacing (mm)

Continuous beams

250 200 150 100 50 0

Continuous beams

Avg Max

Instantaneous t=1000 days

0 5 10 15 20 25 Concrete tensile strength limit, lfct (%)

(c)

0 5 10 15 20 25 Concrete tensile strength limit, lfct (%)

(d)

0.50 0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00

Final crack width (mm)

Crack spacing (mm)

Simply supported slabs

250 200 150 100 50 0

Simply supported slabs

Avg Max

Instantaneous t=1000 days

0 5 10 15 20 25 Concrete tensile strength limit, lfct (%)

(e)

0 5 10 15 20 25 Concrete tensile strength limit, lfct (%)

(f)

0.50 0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0

Final crack width (mm)

Crack spacing (mm)

Continuous slabs

250 200 150 100 50 0

Continuous slabs

Avg Max

Instantaneous t=1000 days

10

15

20

25

(g) (h) Fig. 5.24 - Final crack widths and crack spacings for test series G: (a) and (b) simply supported beams; (c) and (d) continuous beams; (e) and (f) simply supported slabs; (g) and (h) continuous slabs.
278

Concrete tensile strength limit, lfct (%)

5 10 15 20 25 Concrete tensile strength limit, lfct (%)

Crack width ratio, wcr.f/wcr.i

Simply supported beams

Crack width ratio, wcr.f/wcr.i

2.50 2.00 1.50 1.00 0.50 0.00 0

2.50 2.00 1.50 1.00 0.50 0.00 0

Continuous beams

Avg Max

Avg Max

5 10 15 20 25 Concrete tensile strength limit, lfct (%)

(a)

5 10 15 20 25 Concrete tensile strength limit, lfct (%)

(b)

Crack width ratio, wcr.f/wcr.i

Simply supported slabs

Crack width ratio, wcr.f/wcr.i

2.50 2.00 1.50 1.00 0.50 0.00

2.50 2.00 1.50 1.00 0.50 0.00

Continuous slabs

Avg Max

Avg Max

0 5 10 15 20 25 Concrete tensile strength limit, lfct (%)

(c)

0 5 10 15 20 25 Concrete tensile strength limit, lfct (%)

(d)

Fig. 5.25 - Ratios of final to initial crack width for test series G: (a) simply supported beams; (b) continuous beams; (c) simply supported slabs; (d) continuous slabs.

5.3.8

Test Series H Magnitude of Creep

The final crack width and crack spacing of the specimens are plotted against final creep coefficient in Figure 5.26. The ratios of final to initial crack width are shown in Figure 5.27. The results show that cracks are slightly larger in width with increasing creep but the crack spacings are not sensitive to the amount of creep. The final crack widths of the specimens with a final shrinkage strain of 600 have a similar trend as those without shrinkage. It should be pointed out that the crack patterns within the constant moment regions of the specimens in this test series have fully developed at instantaneous loading and therefore crack spacings do not reduce further under the effects of shrinkage. For this reason, the crack spacings of the specimens with
* * shrinkage ( sh = 600 ) and without shrinkage ( sh = 0 ) are not dissimilar as

279

0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0

Final crack width (mm)

Crack spacing (mm)

Simply supported beams

250 200 150 100 50 0 5 0

Simply supported beams

Avg (sh*=0 ) Max (sh*=0 ) Avg (sh*=600 ) Max (sh*=600 )

Instantaneous t=1000 days

1 2 3 4 Final creep coefficient, *

(a)

1 2 3 4 Final creep coefficient, *

(b)

0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0

Final crack width (mm)

Crack spacing (mm)

Continuous beams

250 200 150 100 50 0 5 0

Continuous beams

Avg (sh*=0 ) Max (sh*=0 ) Avg (sh*=600 ) Max (sh*=600 )

Instantaneous t=1000 days

1 2 3 4 Final creep coefficient, *

(c)

1 2 3 4 Final creep coefficient, *

(d)

0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0

Final crack width (mm)

Crack spacing (mm)

Simply supported slabs

300 250 200 150 100 50 0 5 0

Simply supported slabs

Avg (sh*=0 ) Max (sh*=0 ) Avg (sh*=600 ) Max (sh*=600 )

Instantaneous t=1000 days

1 2 3 4 Final creep coefficient, *

(e)

1 2 3 4 Final creep coefficient, *

(f)

0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0

Final crack width (mm)

Crack spacing (mm)

Continuous slabs

300 250 200 150 100 50 0 5 0

Continuous slabs

Avg (sh*=0 ) Max (sh*=0 ) Avg (sh*=600 ) Max (sh*=600 )

Instantaneous t=1000 days

(g) (h) Fig. 5.26 - Final crack widths and crack spacings for test series H: (a) and (b) simply supported beams; (c) and (d) continuous beams; (e) and (f) simply supported slabs; (g) and (h) continuous slabs.
280

1 2 3 4 Final creep coefficient, *

1 2 3 4 Final creep coefficient, *

2.00 1.80 1.60 1.40 1.20 1.00 0.80 0.60 0.40 0.20 0.00 0

Simply supported beams

Avg (sh*=0 ) Max (sh*=0 ) Avg (sh*=600 ) Max (sh*=600 )

2.00 1.80 1.60 1.40 1.20 1.00 0.80 0.60 0.40 0.20 0.00 0

Crack width ratio, wcr.f/wcr.i

Crack width ratio, wcr.f/wcr.i

Continuous beams

Avg (sh*=0 ) Max (sh*=0 ) Avg (sh*=600 ) Max (sh*=600 )

1 2 3 4 Final creep coefficient, *

(a)

1 2 3 4 Final creep coefficient, *

(b)

Crack width ratio, wcr.f/wcr.i

Simply supported slabs

Crack width ratio, wcr.f/wcr.i

2.50 2.00 1.50 1.00 0.50 0.00 0

2.50 2.00 1.50 1.00 0.50 0.00 0

Continuous slabs

Avg (sh*=0 ) Max (sh*=0 ) Avg (sh*=600 ) Max (sh*=600 )

Avg (sh*=0 ) Max (sh*=0 ) Avg (sh*=600 ) Max (sh*=600 )

Final creep coefficient, *

(c)

1 2 3 4 Final creep coefficient, *

(d)

Fig. 5.27 - Ratios of final to initial crack width for test series H: (a) simply supported beams; (b) continuous beams; (c) simply supported slabs; (d) continuous slabs. shown in Figures 5.26b, 5.26d, 5.26f and 5.26h. In addition, the crack width ratios also have the same trend as the crack widths, which tend to be higher at larger final creep coefficients. Initially, the author thought that creep of concrete might be beneficial in reducing crack width of the specimens. The reason for this is that, the concrete between cracks undergoes tension stiffening and it expands longitudinally in the direction of reinforcing steel towards the cracks due to elasticity and creep. Higher creep increases the time-dependent expansion of the concrete between cracks and causes the cracks to
* close. To illustrate this, the results of the continuous beam specimen with sh = 0

for * = 0.3 and * = 3.9 are presented. The typical stress and strain distributions at a section between two adjacent cracks within the constant moment region are shown in Figures 5.28. Comparing Figures 5.28a and 5.28c, the concrete strain surrounding the

281

reinforcing steel (indicated by dashed circle) for * = 3.9 is clearly larger than that for

* = 0.3 , although the concrete tension stiffening stresses for both specimens are of the
same magnitude (about 1.6 MPa as shown in Figures 5.28b and 5.28d). Nevertheless, the test results are not as expected and crack widths tend to be larger for concrete with higher final creep coefficients. After a detailed investigation, it was found that creep in fact has a noticeable effect on the time-dependent change in steel stress. Figures 5.29a and 5.29b compare the time-dependent increase in steel stress between * = 3.9 and * = 0.3 for the continuous beam specimens without the effect of shrinkage. It is seen that concrete with larger creep actually causes larger increase in steel stress and results in widening of cracks with time. The increase in crack width caused by the increase in steel stress is

Distance from bottom (mm)

400 300 200 100 0 -1500 larger expansion in concrete betw cracks
Instantaneous t=1000 days ( sh*=0 )

* = 3.9

Distance from bottom (mm)

500

500 400 300 200 100 0 -10 -5 0 5 Concrete stress betw cracks (MPa)
Instantaneous t=1000 days ( sh*=0 )

* = 3.9

-1000

-500

500

Concrete strain betw cracks ( )

(a)

(b)

400 300 200 100 0 -1500 smaller expansion in concrete betw cracks
Instantaneous t=1000 days ( sh*=0 )

* = 0.3

Distance from bottom (mm)

Distance from bottom (mm)

500

500 400 300 200 100 0 -10 -5 0 5 Concrete stress betw cracks (MPa)
Instantaneous t=1000 days ( sh*=0 )

* = 0.3

-1000

-500

500

Concrete strain betw cracks ( )

(c)

(d)

Fig. 5.28 - Typical stress and strain distributions between cracks for continuous beam specimens: (a) strain diagram for * = 3.9 ; (b) stress diagram for * = 3.9 ; (c) strain diagram for * = 0.3 ; (d) stress diagram for * = 0.3 .

282

300 Steel stress (MPa) 200 100 0 -100 0

t=1000 days ( sh*=0 )

Steel stress (MPa)

Instantaneous

300 200 100 0 -100

Instantaneous t=1000 days ( sh*=0 )

* = 3.9
1000 2000 3000

* = 0.3
0 1000 2000 3000

Distance from left edge (mm)

(a)

Distance from left edge (mm)

(b)

Fig. 5.29 - Longitudinal distribution of stress for bottom reinforcing steel of continuous beam specimens for: (a) * = 3.9 ; (b) * = 0.3 . larger than the crack closing effect resulting from creep induced expansion of concrete between cracks. Consequently, the final outcome is the increase in crack width with increasing final creep. Another important observation from this test series is the extent of cracking with time in the presence of shrinkage. Figures 5.30 and 5.31 show the crack patterns for beam and slab specimens (all specimens shown have a final shrinkage strain
* sh = 600 ), respectively, and compare the effects for * = 0.3 and * = 3.9 . The

crack patterns at instantaneous loading are the same for both * = 0.3 and * = 3.9 since creep under short-term loading is insignificant. It can be seen that the simply supported specimens with very low creep (Figures 5.30c and 5.31c) underwent much more severe time-dependent cracking in the shear span than those with higher creep (Figures 5.30e and 5.31e). This is due to the fact that creep relieves the shrinkageinduced tension in concrete caused by the restraint produced by the reinforcing steel. In absence of creep (or with very low creep), the shrinkage-induced tension in concrete at the level of the reinforcing steel develops and eventually causes further cracking in the shear span. However, this phenomenon is not conspicuous in the continuous specimens due to the nature of the moment distribution in which there is a transition of sagging to hogging moments. Although higher creep causes larger time-dependent opening of cracks, its presence is of crucial importance in preventing extensive time-dependent cracking in reinforced concrete structures.
283

(a)

(b)

(c)

(d)

(e)

(f)

Fig. 5.30 - Crack patterns (cracking strains) for beams: (a) and (b) instantaneous crack patterns for simply supported and continuous beams, respectively; (c) and (d) final crack patterns calculated using * = 0.3 for simply supported and continuous beams, respectively; (e) and (f) final crack patterns calculated using * = 3.9 for simply supported and continuous beams, respectively.

(a)

(b)

(c)

(d)

(e)

(f)

Fig. 5.31 - Crack patterns (cracking strains) for slabs: (a) and (b) instantaneous crack patterns for simply supported and continuous slabs, respectively; (c) and (d) final crack patterns calculated using * = 0.3 for simply supported and continuous slabs, respectively; (e) and (f) final crack patterns calculated using * = 3.9 for simply supported and continuous slabs, respectively.

284

5.3.9

Test Series I Magnitude of Shrinkage

It is well known that shrinkage is the most important factor affecting the timedependent growth of cracks in reinforced concrete structures. This is further reassured in this series of tests. Figure 5.32 shows the final crack widths and crack spacings for the specimens plotted against final shrinkage strain. The effects of shrinkage on crack width are apparent. Since the crack patterns of the specimens have fully developed at instantaneous loading, the crack spacings are not very sensitive to the amount of shrinkage except for the specimens with the largest final shrinkage strain
* ( sh = 900 ) , where a reduction in crack spacing is observed at age 1000 days. It is

also seen that the width of cracks increases linearly with the final shrinkage strains. The mechanisms of the widening of cracks with time by shrinkage are straightforward. The specimens of this test series were pre-cracked at instantaneous loading. As the concrete between each adjacent crack shrinks the crack widths increase. The amount of time-dependent widening of cracks depends on the magnitude and rate of shrinkage. The ratios of final to initial crack width versus final shrinkage strain are shown in Figure 5.33. The percentage increase in crack width with time is larger if the concrete has a higher final shrinkage strain and the relationship also appears to be linear.

5.3.10 Test Series J Bond Creep The effects of bond creep were investigated herein with three different reinforcement ratios for each specimen. The specimens were analysed with and without bond creep. The results for analyses with bond creep are identical to those shown in Figure 5.14 of Section 5.33. The final crack widths and crack spacings for the specimens both inclusive and exclusive of bond creep are plotted versus reinforcement ratio in Figure 5.34. The results for the specimens with bond creep are indicated as control specimens in the figures.

285

0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0

Final crack width (mm)

Crack spacing (mm)

Simply supported beams

300 250 200 150 100 50 0

Simply supported beams

Avg Max

Instantaneous t=1000 days

200 400 600 800 1000 Final shrinkage strain, sh* ( )

(a)

200 400 600 800 1000 Final shrinkage strain, sh* ( )

(b)

0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0

Final crack width (mm)

Crack spacing (mm)

Continuous beams

300 250 200 150 100 50 0

Continuous beams

Avg Max

Instantaneous t=1000 days

200 400 600 800 1000 Final shrinkage strain, sh* ( )

(c)

200 400 600 800 1000 Final shrinkage strain, sh* ( )

(d)

0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0

Final crack width (mm)

Crack spacing (mm)

Simply supported slabs

350 300 250 200 150 100 50 0 0

Simply supported slabs

Avg Max

Instantaneous t=1000 days

200 400 600 800 1000 Final shrinkage strain, sh* ( )

(e)

200 400 600 800 1000 Final shrinkage strain, sh* ( )

(f)

0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0

Final crack width (mm)

Crack spacing (mm)

Continuous slabs

350 300 250 200 150 100 50 0 0

Continuous slabs

Avg Max

Instantaneous t=1000 days

(g) (h) Fig. 5.32 - Final crack widths and crack spacings for test series I: (a) and (b) simply supported beams; (c) and (d) continuous beams; (e) and (f) simply supported slabs; (g) and (h) continuous slabs.
286

200 400 600 800 1000 Final shrinkage strain, sh* ( )

200 400 600 800 1000 Final shrinkage strain, sh* ( )

2.00 1.80 1.60 1.40 1.20 1.00 0.80 0.60 0.40 0.20 0.00 0

Simply supported beams

Avg Max

2.00 1.80 1.60 1.40 1.20 1.00 0.80 0.60 0.40 0.20 0.00 0

Crack widths ratio, wcr.f/wcr.i

Crack widths ratio, wcr.f/wcr.i

Continuous beams

Avg Max

200 400 600 800 1000 Final shrinkage strain, sh* ( )

(a)

200 400 600 800 1000 Final shrinkage strain, sh* ( )

(b)

Crack width ratio, wcr.f/wcr.i

Simply supported slabs

Crack width ratio, wcr.f/wcr.i

2.50 2.00 1.50 1.00 0.50 0.00 0

2.50 2.00 1.50 1.00 0.50 0.00 0

Continuous slabs

Avg Max

Avg Max

(c) (d) Fig. 5.33 - Ratios of final to initial crack width for test series I: (a) simply supported beams; (b) continuous beams; (c) simply supported slabs; (d) continuous slabs.

200 400 600 800 1000 Final shrinkage strain, sh* ( )

200 400 600 800 1000 Final shrinkage strain, sh* ( )

The final crack widths for the specimens without bond creep in Figure 5.34 have the same trends as those with bond creep, but with a somewhat reduced magnitude. This indicates that inclusion of bond creep increases the crack widths. The crack spacings for the specimens without bond creep are generally lower compared to those with bond creep. This is due to the fact that bond without taking account of bond creep is much stiffer and has better stress transfer between concrete and steel, which consequently results in smaller crack spacing. Figure 5.35 shows the ratios of final to initial crack width of the test specimens. The crack width ratios for specimens with bond creep are slightly higher than those without bond creep, which indicates bond creep causes a further increase in crack opening with time.

287

0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00

Final crack width (mm)

Crack spacing (mm)

Simply supported beams

250 200 150 100 50

Simply supported beams

Avg (no bondcrp) Max (no bondcrp) Avg (control) Max (control)

0.002

0.004 0.006 0.008 Reinforcement ratio,

0.01

0 0.002

Inst. (no bondcrp) t=1000days (no bondcrp) Inst. (control) t=1000days (control)

(a)

0.004 0.006 0.008 Reinforcement ratio,

0.01

(b)

0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00

Final crack width (mm)

Crack spacing (mm)

Continuous beams

250 200 150 100 50

Continuous beams

Avg (no bondcrp) Max (no bondcrp) Avg (control) Max (control)

0.002

0.004 0.006 0.008 Reinforcement ratio,

0.01

0 0.002

Inst. (no bondcrp) t=1000days (no bondcrp) Inst. (control) t=1000days (control)

(c)

0.004 0.006 0.008 Reinforcement ratio,

0.01

(d)

0.60

Final crack width (mm)

0.50 0.40 0.30 0.20 0.10 0.00 0.002


Avg (no bondcrp) Max (no bondcrp) Avg (control) Max (control)

Crack spacing (mm)

Simply supported slabs

400 350 300 250 200 150 100 50

Simply supported slabs

0.004 0.006 0.008 Reinforcement ratio,

0.01

0 0.002

Inst. (no bondcrp) t=1000days (no bondcrp) Inst. (control) t=1000days (control)

(e)

0.004 0.006 0.008 Reinforcement ratio,

0.01

(f)

0.60

Final crack width (mm)

0.50 0.40 0.30 0.20 0.10 0.00 0.002


Avg (no bondcrp) Max (no bondcrp) Avg (control) Max (control)

Crack spacing (mm)

Continuous slabs

400 350 300 250 200 150 100 50

Continuous slabs

(g) (h) Fig. 5.34 - Final crack widths and crack spacings for test series J: (a) and (b) simply supported beams; (c) and (d) continuous beams; (e) and (f) simply supported slabs; (g) and (h) continuous slabs.
288

0.004 0.006 0.008 Reinforcement ratio,

0.01

0 0.002

Inst. (no bondcrp) t=1000days (no bondcrp) Inst. (control) t=1000days (control)

0.004 0.006 0.008 Reinforcement ratio,

0.01

Crack width ratio, wcr.f/wcr.i

Crack width ratio, wcr.f/wcr.i

2.50 2.00 1.50 1.00 0.50 0.00

Simply supported beams

2.50 2.00 1.50 1.00 0.50

Continuous beams

Avg (no bondcrp) Max (no bondcrp) Avg (control) Max (control)

0.002

0.004 0.006 0.008 Reinforcement ratio,

0.01

0.00 0.002

Avg (no bondcrp) Max (no bondcrp) Avg (control) Max (control)

(a)

0.004 0.006 0.008 Reinforcement ratio,

0.01

(b)

Crack width ratio, wcr.f/wcr.i

Simply supported slabs

Crack width ratio, wcr.f/wcr.i

2.50 2.00 1.50 1.00 0.50 0.00

2.50 2.00 1.50 1.00 0.50

Continuous slabs

Avg (no bondcrp) Max (no bondcrp) Avg (control) Max (control)

0.002

0.004 0.006 0.008 Reinforcement ratio,

0.01

0.00 0.002

Avg (no bondcrp) Max (no bondcrp) Avg (control) Max (control)

(c)

0.004 0.006 0.008 Reinforcement ratio,

0.01

(d)

Fig. 5.35 - Ratios of final to initial crack width for test series J: (a) simply supported beams; (b) continuous beams; (c) simply supported slabs; (d) continuous slabs.

5.3.11 Test Series K Quantity of Shear Reinforcement As mentioned previously in Section 5.2.1, the shear stresses in the shear span of the continuous beam specimens are high and therefore shear reinforcement must be provided to prevent premature shear failure in the specimens. The idealisation of reinforcement as one-dimensional truss elements in the finite element model does not account for the stress concentration effects caused by the geometry of stirrups. However, the primary interest of this study is to investigate time-dependent cracking related to longitudinal reinforcement and the inclusion of shear reinforcement is merely for the sake of preventing unwanted shear failure in the specimens. Therefore, instead of attempting to model the precise physical influence of stirrups in a real beam, the actual purpose of this test series is to examine the influence of the amount of shear resisting truss elements on the results of the numerical model.

289

The results of final crack width, crack spacing and final to initial crack width ratio are plotted against quantity of shear reinforcement are shown in Figures 5.36a, 5.36b and 5.36c, respectively. The amount of shear reinforcement has no effect on the development of flexural cracks, either in terms of width, spacing or increase of width with time. The outcome of this test series does not agree with the experimental observation of Divakar and Dilger (1987), from which they concluded that cracking tends to occur at the location of stirrups. However, for the purpose of this study, it may be concluded that the inclusion of shear reinforcement does not induce any unanticipated effects on the numerical results. This is important in the sense of facilitating an objective comparison of results between the simply support beam and continuous beam specimens.

0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00 0

Avg Max

200 180 160 140 120 100 80 60 40 20 0 0

Final crack width (mm)

Crack spacing (mm)

Instantaneous t=1000 days

0.5 1 1.5 Shear reinf. quantity, Asv /s (mm)

(a)

0.5 1 1.5 Shear reinf. quantity, Asv /s (mm)

(b)

2.00 1.80 1.60 1.40 1.20 1.00 0.80 0.60 0.40 0.20 0.00 0

Crack width ratio, wcr.f/wcr.i

Avg Max

(c) Fig. 5.36 - Results of continuous beam specimens for test series K: (a) and (b) final crack width and cracking spacing, respectively; (c) ratio of final to initial crack width.

0.5 1 1.5 Shear reinf. quantity, Asv /s (mm)

290

5.3.12 Test Series L Impact of 500 MPa Steel Reinforcement Figure 5.37 shows the results for final crack width, crack spacing and final to initial crack width ratio plotted against moment capacity of the specimens. It is seen that the final crack widths for specimens using the reinforcement with f sy = 500 MPa are generally larger than those containing reinforcement with f sy = 400 MPa . For under-reinforced flexural members, the reinforcing steel reaches the yield stress before the concrete fails in crushing. Consequently, the moment capacity of a section is proportional to the strength of the reinforcement. Comparing the two types of reinforcement, the use of 500 MPa steel inevitably reduces the quantity of steel required for a desired strength. Using the same bar size, the required number of bars in a specimen containing 500 MPa steel must be fewer than that using 400 MPa steel. The consequence of this is a smaller concrete-steel contact surface area, which means weaker bond. Weak bond results in large crack spacing and wider cracks. This clearly explains why the use of 500 MPa steel leads to larger crack widths and crack spacing as observed in Figure 5.37. In addition, under the same loads, smaller quantities of steel also result in higher tensile steel stresses. This is exactly the case for the specimens containing 500 MPa steel. The higher steel stress in the 500 MPa steel inevitably results in wider cracks. This is evidenced in Figures 5.37a and 5.37b for the beam specimens with M u = 130 kNm . The crack spacings are the same for both beams containing 400 MPa and 500 MPa steel, however, crack widths are significantly different. This is because of the significantly higher tensile stress in the 500 MPa steel than in the 400 MPa steel at service loads. In Figures 5.37e and 5.37f, it is seen that the percentage increase in crack width for the specimens with 400 MPa steel are in general higher than those with 500 MPa steel. This is, again, due to the difference of the tensile steel stresses in these two types of steel reinforcement. The lower tensile steel stress in the 400 MPa steel results in less

291

0.35

Final crack width (mm)

Crack spacing (mm)

0.30 0.25 0.20 0.15 0.10 0.05 0.00 0

Simply supported beams

160 140 120 100 80 60 40 20 0 200 300 0

Simply supported beams

Avg (fsy=500) Max (fsy=500) Avg (fsy=400) Max (fsy=400)

Inst. (fsy=500) t=1000days (fsy=500) Inst. (fsy=400) t=1000days (fsy=400)

100

Design moment capacity, Mu (kNm)

(a)

100 200 300 Design moment capacity, Mu (kNm)

(b)

0.40

Final crack width (mm)

0.30 0.25 0.20 0.15 0.10 0.05 0.00 100


Avg (fsy=500) Max (fsy=500) Avg (fsy=400) Max (fsy=400)

110 120 130 140 150 Design moment capacity, Mu (kNm)

Crack spacing (mm)

0.35

Simply supported slabs

200 180 Simply supported slabs 160 140 120 100 80 60 Inst. (fsy=500) t=1000days (fsy=500) 40 Inst. (fsy=400) 20 t=1000days (fsy=400) 0 100 110 120 130

(c)

140 150 Design moment capacity, Mu (kNm)

(d)

Crack width ratio, wcr.f/wcr.i

Simply supported beams

Crack width ratio, wcr.f/wcr.i

2.50 2.00 1.50 1.00 0.50 0.00 0

2.50 2.00 1.50 1.00 0.50

Simply supported slabs

Avg (fsy=500) Max (fsy=500) Avg (fsy=400) Max (fsy=400)

100

200

300

0.00 100

Avg (fsy=500) Max (fsy=500) Avg (fsy=400) Max (fsy=400)

Design moment capacity, Mu (kNm)

(e)

110 120 130 140 150 Design moment capacity, Mu (kNm)

(f)

Fig. 5.37 - Results for test series L: (a) and (c) final crack widths for simply supported beams and slabs, respectively; (b) and (d) crack spacings for simply supported beams and slabs, respectively; (e) and (f) ratios of final to initial crack width for simply supported beams and slabs, respectively. expansion in the concrete between two adjacent cracks due to creep and elasticity (also see the discussion in Section 5.3.8 for test series H). With less expansion, the cracks of the specimens with 400 MPa steel opened more widely with time due to shrinkage. It is evident that the use of 500 MPa steel reinforcement changes the way reinforced concrete structures behave at service load conditions. The consequent
292

reduction of tensile reinforcement adversely increases the tensile steel stress. In addition, the reduction of steel quantity can easily lead to reduced bond between concrete and steel, particularly if the detailing of reinforcement is not done with care. This eventually results in larger crack spacings. The consequence of these is wider cracks and an increased likelihood of serviceability problem.

5.3.13 Test Series M Load Histories

5.3.13.1 Comparisons between LH-2 and LH-1 In this part of the test series, the effects of the sustained load levels are investigated. The final crack width, crack spacing and ratios of final to initial crack width for two different sustained load levels are presented in Figure 5.38. The specimens were subjected to either load histories LH-2 (sustained moment 0.35M u ) or LH-1 (sustained moment 0.5M u ). It should be noted in Figures 5.38c, 5.38d and 5.38f that the data points for the lightly reinforced slab containing 3N16 reinforcing bars ( = 0.003) subjected LH-2 are absent because the 0.35M u applied moment was too low to initiate cracking in the slab specimen. As expected, the final crack widths are, on the whole, smaller for the specimens subjected to lower sustained loads. However, it is surprising to see that the final crack widths caused by a sustained moment of 0.35M u are not very much smaller than those caused by 0.5M u . The specimens subjected to the lower sustained moments generally have a larger crack spacing, due to the fact that a stabilized crack pattern does not form initially under the smaller loads. The large crack spacings cancel out the effects of low tensile stress in the steel due to the smaller sustained moment 0.35M u . Therefore the crack widths are not much smaller than those caused by 0.5M u .

293

0.40

Final crack width (mm)

Crack spacing (mm)

0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00

Simply supported beams

300 250 200 150 100 50

Simply supported beams

Avg (LH-2) Max (LH-2) Avg (LH-1) Max (LH-1)

0.002

0.004 0.006 0.008 Reinforcement ratio,

0.01

0 0.002

Inst. (LH-2) t=1000days (LH-2) Inst. (LH-1) t=1000days (LH-1)

(a)

0.004 0.006 0.008 Reinforcement ratio,

0.01

(b)

0.60

Final crack width (mm)

0.50 0.40 0.30 0.20 0.10 0.00 0.002


Avg (LH-2) Max (LH-2) Avg (LH-1) Max (LH-1)

Crack spacing (mm)

Simply supported slabs

400 350 300 250 200 150 100 50

Simply supported slabs

0.004 0.006 0.008 Reinforcement ratio,

0.01

0 0.002

Inst. (LH-2) t=1000days (LH-2) Inst. (LH-1) t=1000days (LH-1)

(c)

0.004 0.006 0.008 Reinforcement ratio,

0.01

(d)

Crack width ratio, wcr.f/wcr.i

Simply supported beams

Crack width ratio, wcr.f/wcr.i

3.00 2.50 2.00 1.50 1.00 0.50 0.00

3.00 2.50 2.00 1.50 1.00 0.50

Simply supported slabs

Avg (LH-2) Max (LH-2) Avg (LH-1) Max (LH-1)

0.002

(e) (f) Fig. 5.38 - Test series M comparisons of results for load histories LH-2 (sustained moment 0.35Mu) and LH-1 (sustained moment 0.5Mu): (a) and (c) final crack widths for simply supported beams and slabs, respectively; (b) and (d) crack spacings for simply supported beams and slabs, respectively; (e) and (f) ratios of final to initial crack width for simply supported beams and slabs, respectively.

0.004 0.006 0.008 Reinforcement ratio,

0.01

0.00 0.002

Avg (LH-2) Max (LH-2) Avg (LH-1) Max (LH-1)

0.004 0.006 0.008 Reinforcement ratio,

0.01

An interesting feature is observed in Figures 5.38e and 5.38f, which show the final to initial crack width ratios for the specimens. The percentage increase in crack width for the specimens subjected to 0.35M u are noticeably higher than that for
294

specimens subjected to 0.5M u . This indicates that the crack widths for the specimens loaded at 0.35M u were initially small at instantaneous loading but increased significantly with time. Figures 5.39a and 5.39b show the typical stress distributions on the sections containing cracks resulting from load histories LH-1 and LH-2, respectively, for the simply supported beams. For the beam subjected to sustained moment 0.5M u , the concrete stress at the crack (indicated by a dashed circle in Figure 5.39a) has completely reduced to zero as expected. However, for the beam subjected to lower sustained moment it is seen that some residual concrete tensile stresses are still present at the crack (dashed circle in Figure 5.39b). These residual stresses are the cohesive stresses in the concrete at the crack front (fracture process zone), which indicates that a portion of the crack is still undergoing the process of fracture (tension softening). The implication of this is that, the beam specimen subjected to sustained moment 0.35M u is yet to be thoroughly cracked. In other words, the crack faces were not completely separated. With time, the cohesive stresses started to dissipate as the cracks widened due to shrinkage of concrete. This can be seen in Figure 5.39b where the crack is completely stress-free at age 1000 days. The phenomenon described above can explain why the lightly loaded specimens have a higher ratio of final to initial crack width. In the lightly loaded specimens, although cracking has been initiated upon loading, the cracks were unable to open freely due to the presence of residual stress in the cracks. Shrinkage of concrete overcame the residual stress and widened the cracks as they developed with time. This observation is not merely academic but actually happens to reinforced concrete structures in service. This observation has important implications to long-term serviceability of reinforced concrete structures. Fine cracks in a lightly loaded structure that seem to be acceptable under short-term loading can turn out to be large unsightly cracks, or more seriously, they can affect the long-term durability of the structure.

295

Distance from bottom (mm)

Distance from bottom (mm)

500 400 300 200 100 0 -10 -8 -6 -4 -2 No residual tension softening stress at instantaneous loading
Instantaneous t=1000 days

LH-1

500 400 300 200 100 0 -10 -8 -6 -4 -2 Residual tension softening stress at instantaneous loading
Instantaneous t=1000 days

LH-2

Concrete stress (MPa)

(a)

Concrete stress (MPa)

(b)

Fig. 5.39 - Typical stress distributions of section containing cracks for the simply supported beam specimen subjected to: (a) LH-1 (sustained moment 0.5M u ); (b) LH-2 (sustained moment 0.35M u ).

5.3.13.2 Comparisons between LH-5 and LH-2 This part of the test is complementary to the previous part that has been presented in Section 5.3.13.1. The specimens were tested under load history LH-5, for which the specimens were loaded to 0.5M u and followed by an immediate unloading to 0.35M u . The results for this loading history are compared with the results for load history LH-2 (sustained moment 0.35M u ) in Figure 5.40, which are presented in terms of final crack width, crack spacing and final to initial crack width ratio. It should be noted that crack width ratio was calculated by dividing the crack width at age 1000 days by the crack width right after unloading to 0.35M u . Although the magnitudes of the long-term moment are the same for both load histories LH-5 and LH-2, the final crack widths and crack spacings for the specimens subjected to LH-5 are smaller than those subjected to LH-2. The difference between the two load histories is that, under load history LH-5 the specimens were fully cracked (cracks are stress-free) at instantaneous loading and the stabilized crack patterns were established. These also explain why the cracks of the specimens subjected to LH-5 were more closely spaced and have smaller widths. Although the instantaneous crack

296

0.35

Final crack width (mm)

Crack spacing (mm)

0.30 0.25 0.20 0.15 0.10 0.05 0.00

Simply supported beams

300 250 200 150 100 50

Simply supported beams

Avg (LH-5) Max (LH-5) Avg (LH-2) Max (LH-2)

0.002

0.004 0.006 0.008 Reinforcement ratio,

0.01

0 0.002

Inst. (LH-5) t=1000days (LH-5) Inst. (LH-2) t=1000days (LH-2)

(a)

0.004 0.006 0.008 Reinforcement ratio,

0.01

(b)

0.45 0.40 0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00

Final crack width (mm)

Crack spacing (mm)

Simply supported slabs

400 350 300 250 200 150 100 50

Simply supported slabs

Avg (LH-5) Max (LH-5) Avg (LH-2) Max (LH-2)

0.002

0.004 0.006 0.008 Reinforcement ratio,

0.01

0 0.002

Inst. (LH-5) t=1000days (LH-5) Inst. (LH-2) t=1000days (LH-2)

(c)

0.004 0.006 0.008 Reinforcement ratio,

0.01

(d)

Crack width ratio, wcr.f/wcr.i

Simply supported beams

Crack width ratio, wcr.f/wcr.i

3.00 2.50 2.00 1.50 1.00 0.50 0.00

3.00 2.50 2.00 1.50 1.00 0.50

Simply supported slabs

Avg (LH-5) Max (LH-5) Avg (LH-2) Max (LH-2)

0.002

(e) (f) Fig. 5.40 - Test series M comparisons of results for load histories LH-5 (initial moment 0.5Mu then to 0.35Mu) and LH-2 (sustained moment 0.35Mu): (a) and (c) final crack widths for simply supported beams and slabs, respectively; (b) and (d) crack spacings for simply supported beams and slabs, respectively; (e) and (f) ratios of final to initial crack width for simply supported beams and slabs, respectively. widths may not be as small as those for specimens subjected to LH-2, the long-term crack widths are in fact smaller and are more acceptable from a serviceability point of view.

0.004 0.006 0.008 Reinforcement ratio,

0.01

0.00 0.002

Avg (LH-5) Max (LH-5) Avg (LH-2) Max (LH-2)

0.004 0.006 0.008 Reinforcement ratio,

0.01

297

In addition, the lower final to initial crack width ratio for the specimens under load history LH-5 is less deceptive than that due to LH-2. This is more advantageous in practical situations in that the users of the structures frequently become unduly concerned if cracks gradually appear with time and subsequently continue to widen.

5.3.13.3 Comparisons between LH-3 and LH-4 In this part of the test series, the effects of the age of unloading were investigated. All specimens were initially loaded to 0.5M u and held constant with time. After a certain period, the external loads were completely removed. To achieve a more realistic simulation as in real structures, the self-weight of the specimens remained in place after the external loads were removed. The specimens were unloaded either at age 50 days or at age 200 days (designated as load histories LH-3 and LH-4, respectively). Figure 5.41 shows the final crack width, the crack spacing and the ratio of final to initial crack width, for both load histories LH-3 and LH-4. Since the loads are the same for the both load histories at instantaneous loading, the crack spacings are also not different. Moreover, no further change of crack patterns is observed after the loads were removed. In Figures 5.41a and 5.41c, it is seen that the final crack widths for the specimens subjected to LH-3 are generally smaller than those subjected to LH-4, but the difference is only slight. In most sustained loading cases, the time-dependent opening of cracks is largely dependent on the ability of a concrete to shrink. However, when the loads vary with time, the tensile stress in the reinforcing steel changes and this affects the width of cracks. Since all specimens were assumed to have the same shrinkage properties, with the same crack spacing the cracks should open by the same amount under the effects of shrinkage. The larger crack widths observed in Figures 5.41a and 5.41c for the specimens unloaded at age 200 days are due to the fact that aging concrete is less able to undergo creep recovery. In addition, the specimens unloaded at age 200 were also subjected to an additional 150 days (from age 50 days to age 200 days) of sustained

298

0.18 0.16 0.14 0.12 0.10 0.08 0.06 0.04 0.02 0.00

Final crack width (mm)

Crack spacing (mm)

Simply supported beams

250 200 150 100 50

Simply supported beams

Avg (LH-3) Max (LH-3) Avg (LH-4) Max (LH-4)

0.002

0.004 0.006 0.008 Reinforcement ratio,

0.01

0 0.002

Inst. (LH-3) t=1000days (LH-3) Inst. (LH-4) t=1000days (LH-4)

(a)

0.004 0.006 0.008 Reinforcement ratio,

0.01

(b)

0.40

Final crack width (mm)

Crack spacing (mm)

0.35 0.30 0.25 0.20 0.15 0.10 0.05 0.00

Simply supported slabs

300 250 200 150 100 50

Simply supported slabs

Avg (LH-3) Max (LH-3) Avg (LH-4) Max (LH-4)

0.002

0.004 0.006 0.008 Reinforcement ratio,

0.01

0 0.002

Inst. (LH-3) t=1000days (LH-3) Inst. (LH-4) t=1000days (LH-4)

(c)

0.004 0.006 0.008 Reinforcement ratio,

0.01

(d)

Crack width ratio, wcr.f/wcr.i

0.70 0.60 0.50 0.40 0.30 0.20 0.10

Simply supported beams

0.00 0.002

Avg (LH-3) Max (LH-3) Avg (LH-4) Max (LH-4)

(e) (f) Fig. 5.41 - Test series M comparisons of results for load histories LH-3 (initial moment 0.5Mu and unloaded at age 50 days) and LH-4 (initial moment 0.5Mu and unloaded at age 200 days): (a) and (c) final crack widths for simply supported beams and slabs, respectively; (b) and (d) crack spacings for simply supported beams and slabs, respectively; (e) and (f) ratios of final to initial crack width for simply supported beams and slabs, respectively. loads which inevitably causes a higher tensile steel stress with time than those subjected to a shorter period of loading. Figure 5.42 compares the average tensile steel

0.004 0.006 0.008 Reinforcement ratio,

0.01

Crack width ratio, wcr.f/wcr.i

0.80

1.80 Simply supported slabs 1.60 1.40 1.20 1.00 0.80 0.60 Avg (LH-3) 0.40 Max (LH-3) Avg (LH-4) 0.20 Max (LH-4) 0.00 0.002 0.004 0.006 0.008 Reinforcement ratio,

0.01

299

stresses of the beam specimens containing 3N20 steel bars ( = 0.0069) subjected to load histories LH-3 and LH-4. The additional increase in steel stress due to a longer period of sustained loading and the reduced ability of concrete to undergo creep recovery at an older age result in higher final stress in steel and therefore wider cracks. In Figures 5.41e and 5.41f, it is seen that the ratios of final to initial crack width for the specimens subjected to unloading at age 50 days are slightly lower compared to those subjected to unloading at age 200 days. This is to be expected since the specimens containing the same quantity of steel have the same instantaneous crack widths but the final crack widths of the specimens unloaded at age 50 days are smaller, which results in a smaller crack width ratio. Another observation in Figures 5.41e and 5.41f is the reduction of crack width ratio with increasing reinforcement ratio. The results may seem intriguing at first glance since the crack width ratios for specimens subjected to sustained moments in the previous test series are consistently insensitive to the quantity of reinforcement. In fact, this feature originates from the different magnitudes of unloading for each specimen. For example, the most heavily reinforced specimens were subjected to the largest initial moment equivalent to 0.5M u . Having the same self-weights as other specimens, the most heavily reinforced specimens undergo the largest unloading which subsequently causes the crack widths to have the largest decrease, too.

Average steel stress (MPa)

300 250 200 150 100 50 0 0 200 400 600 Age (days) 800 1000
LH-3 LH-4

Fig. 5.42 - Comparison of steel stress of beam specimens containing 3N20 steel reinforcement subjected to load histories LH-3 and LH-4.

300

For the same reason, the slab specimens have the same trend of crack width ratios as the beam specimens, but with a higher range. The ratio for the slabs lies between 0.7 and 1.4, while the crack width ratios for beam specimens are between 0.4 and 0.65. In addition, the higher crack width ratios are due to the larger self-weights of the slab specimens, which effectively retains a larger width of crack after unloading.

5.3.14 Section Geometry and Boundary Conditions Examining the results of all test series, it is observed that the type of specimen (beam or slab in this study) has no effect on crack spacing and crack width. The boundary conditions have a slight influence on the time-dependent growth of cracks. In general, the continuous specimens have slightly higher final to initial crack width ratios. This is probably due to the large bending moment in the negative moment region, which causes cracking to occur above the supports before cracking in the constant positive moment region. The delay of cracking in the constant positive moment region may have caused a relatively smaller instantaneous crack width than that in the simply supported case. With time, these cracks opened widely and achieved similar crack widths. Nevertheless, the difference of the crack width ratios between simply supported and continuous specimens is very small.

5.4 Summary
Various parameters have been investigated in this parametric study. The effects of each parameter were examined thoroughly and the test results have been evaluated and discussed. In addition to the investigation of the parameters, the versatility of the proposed finite element model has also been demonstrated. The finite element model has been shown to be able to produce a large amount of data for the study of timedependent cracking of reinforced concrete structures, and this can be used as an alternative to conventional laboratory experimentations. The parameters tested in this study are broadly divided into three groups. The first group includes the parameters which affect both the instantaneous crack spacing and
301

crack width and the time-dependent growth of cracks. The second group consists of the parameters that do not affect the formation of cracks at instantaneous loading but are important in the time-dependent widening of cracks. The final group are made up of the parameters that have no significant effect on either instantaneous or time-dependent cracking of reinforced concrete structures. Parameters such as bottom concrete cover, bar diameter, quantity of tensile reinforcement, concrete tensile strength and bond strength are in the first group. These parameters are important in the formation of cracks under short-term loads. A common feature of these parameters is that the ratios of final to initial crack widths are relatively insensitive to the variation of these parameters. In the broad sense, these parameters are interrelated in a way that they have some common interactions. They affect the formation of cracks by either changing the effective reinforcement ratio (area of tensile steel divided by effective area of concrete in tension) or the bond characteristic between steel and concrete. Bottom concrete cover and the quantity of tensile reinforcement are the factors affecting the effective reinforcement ratio which subsequently govern the crack spacing and crack width. A threshold thickness of bottom cover was identified beyond which the bottom cover ceases to affect the distribution of cracks in a structure. Bar diameter, quantity of tensile reinforcement and bond strength all influence the final bond characteristics of the structure. Bond quality can be improved by using stronger concrete which provides good confinement to the reinforcing steel and consequently improves local bond characteristics between concrete and steel. An alternative is to adjust the bar size or the number of bars, or both, so as to achieve a larger concrete-steel contact surface area. In addition, this investigation also reveals that spacing of reinforcing steel, which is often quoted in codes of practice for crack control, is not a critical factor that affects crack spacing and crack width. The more important factor is the concrete-steel contact surface area which directly affects the rate of stress transfer between steel and concrete. The concrete tensile strength parameter is a somewhat special factor which affects the local bond characteristics of the reinforcing steel and also limits the extent of the bond development length. The effects of concrete tensile strength in these two aspects
302

are inter-related and rather obscure. It is well known that bond strength increases with increasing concrete tensile strength, and this affects the stress transfer between concrete and steel. When these two aspects are considered simultaneously, the effects of concrete tensile strength become less important as demonstrated in Section 5.3.5. However, if a different relationship between bond and concrete tensile strength is used, the outcome of the tests may be different. Therefore, more research on the interaction of concrete tensile strength and bond is required before a final conclusion can be made. Creep, shrinkage, bond creep and load history are categorized into the second parameter group. These parameters vary constantly with time. Undoubtedly, shrinkage is the most important time-dependent factor that affects the opening of cracks with time, while creep is a less influential one. Nevertheless, the importance of creep cannot be completely overlooked, since creep has a prominent effect on the internal redistribution of stress in a structure which may subsequently lead to an increase in tensile stress in the steel. Although it is seen in the parametric study that the effects of creep are less significant on crack opening, care should still be taken when considering the effects of creep on time-dependent cracking of reinforced concrete structures. Load history has an important influence on crack width in the sense that the tensile steel stress changes according to the change in loads. Since the crack opening mechanisms depend largely on shrinkage (load-independent) and tensile stress in steel (due mainly to change in load), it is concluded that the final crack widths in a reinforced concrete structure are not particularly sensitive to the age of change of loads as demonstrated in Section 5.3.13.3. In addition, this investigation also shows the benefits of establishing a stabilized crack pattern at instantaneous loading. The instantaneous crack widths of a stabilized crack pattern are usually larger than those of a pre-stabilized crack pattern developed under a smaller load. However, in the course of time, the smaller crack spacing of the stabilized crack pattern is more beneficial in reducing the long-term width of cracks. Furthermore, the fine instantaneous cracks which result from a relatively small initial load may become surprisingly wide over time, and this can have devastating effects on the serviceability of the structure.

303

The third parameter group includes parameters such as compressive reinforcement, concrete tensile strength fluctuation limit, structure type and boundary conditions. The effects of these parameters are less important to crack spacing and crack width. Although concrete tensile strength fluctuation limit does not have significant effects on crack spacing and crack width, it is important in the initiation of cracking in concrete and the prevention of bifurcation of the numerical solution. The shear reinforcement parameter is left out of any of the parameter group since shear reinforcement was included only to prevent premature shear failure. Lastly, an investigation was conducted to examine the effects of increasing the steel yield stress from 400 MPa to 500 MPa. The results of the investigation show that the consequent reduction of the required quantity of steel in structural members results in higher tensile steel stress. As a result, the crack spacings are larger and the cracks are wider. This study demonstrates the adverse effects of the use of 500 MPa steel reinforcement and urges an appropriate review of the crack control procedures in the current Australian Standards, AS 3600 (2001).

304

CHAPTER 6 SUMMARY AND CONCLUSIONS

6.1 Summary
The analysis of a reinforced concrete structure is complicated by the non-linear behaviour arising from cracking of the concrete. A time-dependent analysis is further complicated by time effects resulting from the load history and creep and shrinkage of concrete. In this thesis, several two-dimensional numerical models for time-dependent analysis of reinforced concrete structures have been presented, with particular emphasis on the formation of cracks at service load conditions. Three numerical models, namely the distributed cracking Cracked Membrane Model, the localized cracking Crack Band Model and the localized cracking Non-local Smeared Crack Model, were developed to trace the gradual cracking process in a reinforced concrete structure subjected to the time-dependent effects of creep and shrinkage. In addition to the non-linearities resulting from cracking and time effects, non-linearity due to large deformation is also accounted for in the numerical models. In the first part of Chapter 2, the properties of concrete and steel reinforcement and the interaction between these two materials are described. An overview is given of the instantaneous and time-dependent properties of concrete including the well-known behaviour of uniaxially and biaxially loaded specimens, creep and shrinkage and the time-dependent fracture of concrete. The second part of Chapter 2 presents the state-ofthe-art of non-linear modelling of reinforced concrete structures. Various methods for modelling cracking in concrete are described and the pros and cons of each model are discussed. The commonly employed constitutive models for concrete are also presented. An introduction is given of issues related to strain localization in a standard smeared crack model and various regularization techniques developed to overcome these difficulties are described. In addition, the representation of reinforcing steel in the

305

context of a finite element formulation is presented. Lastly, the numerical treatment for a creep analysis based the principle of superposition is described. The formulations of the numerical models are presented in Chapter 3. The numerical models have been developed based on a smeared crack approach for which cracking is modelled as a reduction of material stiffness in the crack opening direction. Two modelling approaches have been adopted to simulate the behaviour of reinforced concrete structures. The first type is the distributed cracking approach. In this approach, steel reinforcement is smeared through the concrete elements and the effect of bondslip between steel and concrete is accounted for indirectly as tension stiffening in a cracked reinforced concrete member. One of the characteristics of the distributed cracking model is that cracking occurs in a smeared manner where no visible localized crack is computed. For an effective modelling of tension stiffening, the cracked membrane model of Kaufmann (1998) is employed. Based upon a stepped, rigid perfectly plastic bond stress-slip relationship as adopted in the tension chord model of Sigrist (1995), the crack spacing and the stress distribution in steel and in concrete for a cracked reinforced concrete member can be determined. In this way, tension stiffening is treated in a rational and tractable manner. The second approach described in Chapter 3 for modelling cracking is the localized cracking approach. In this approach, cracking in concrete is treated as localized fracture and reinforcing steel is modelled as discrete steel elements connected to concrete using bond-slip interface elements. The localized cracking approach facilitates a more refined simulation of the mechanism of cracking in a reinforced concrete structure. Tension stiffening is modelled as a result of stress transfer between the concrete and the steel via bond action. In this study two different fracture models were employed, namely the crack band model (Baant and Oh, 1983) and the non-local smeared crack model. For the non-local smeared crack model, a new formulation is proposed to avoid the unrealistic concrete stress in excess of the concrete tensile strength computed by Jirsek and Zimmermann (1998) in their model based on the nonlocal model with local strain formulation (see Section 3.4.2). In this study, plain concrete is treated as orthotropic material and the biaxial stresses are obtained using a modified form of the equivalent uniaxial strain concept of

306

Darwin and Pecknold (1977). Although the work in this thesis focuses on timedependent behaviour of reinforced concrete structures at service loads, the stress-strain relationships for biaxial loading to failure are incorporated in the numerical models so that a short-term failure analysis can also be performed. Creep is modelled using the solidification theory of Baant and Prasannan (1989a, b) and shrinkage is approximated by a growth function. Both these time effects are treated as inelastic pre-strains updated with time and applied to the structure as equivalent nodal forces. A time-dependent bond model is implemented with the bond-slip interface elements and is used in the localized cracking models and therefore the deterioration of bond due to creep under sustained loads is accounted for. The time-dependent analysis of a reinforced concrete structure is treated in a quasi-static manner such that the time domain is discretized into a number of finite time steps where the time-dependent cracking and deformation of the structure are determined through a step-by-step integration through the time domain. In addition, for large deformation problems such as creep buckling of slender columns, an updated Lagrangian formulation is adopted. Using this formulation the stress and strain of a structure are computed based on the last computed displaced configuration and therefore the engineering stress and strain notion can be retained. The numerical models developed in Chapter 3 are verified in Chapter 4. The major task is to evaluate the numerical models in computing crack spacing and crack width of reinforced concrete members. The long-term flexural cracking specimens and the restrained deformation slabs tested by Gilbert and Nejadi (2004) and Nejadi and Gilbert (2004) were simulated using the numerical models. Despite the random nature of cracking in the reinforced concrete members, the calculated results are in good agreement with the test data. In addition, a comparison of accuracy of the computed results is made among the numerical models and a discussion is given on the advantages and disadvantages of each model. Other types of structures such as continuous beams with long-term settling supports and creep buckling of slender columns have also been analysed. The good correlations between the calculated results and the test data demonstrate the ability of the numerical models in simulating timedependent behaviour of reinforced concrete structures under varying load histories and under the effect of geometric non-linearity.

307

In Chapter 5, a parametric study is conducted using the localized cracking Crack Band Model in order to investigate the qualitative interactions of the parameters affecting time-dependent cracking of reinforced concrete structures. A series of numerical experiments is devised based on various parameters associated with material properties, environmental conditions and structural type. The effects of the parameters are examined and discussed based on the observations from the numerical calculations. In addition, the impact of the introduction of 500 MPa steel to replace the previously used 400 MPa steel in Australia is also examined and evaluated using the numerical models.

6.2 Conclusions
The major objective of this thesis is to investigate time-dependent cracking in reinforced concrete structures with particular focus on the qualitative understanding of the time-dependent cracking process and the quantitative evaluation of crack spacing and crack width due to time effects. In Chapter 4, the numerical models developed in this study have been demonstrated to compute reasonably accurate quantitative results and, in Chapter 5, the models have been employed to investigate the process of timedependent cracking in reinforced concrete structures through a series of numerical experiments. Based upon the work presented in this thesis, the following conclusions have been reached. A realistic description of bond between concrete and steel is of vital importance for the accurate computation of crack spacing and crack width in a reinforced concrete structure. Although bond-slip is not modelled explicitly as a physical interaction between the concrete and the steel elements in the distributed cracking approach, the bond condition assumed in the tension chord model is crucial for the derivation of the crack spacing equation and hence the determination of crack width. For the localized cracking model, the influence of bond is even more apparent. As can be seen from the results of the parametric study, the characteristic of bond between concrete and steel has a marked effect on the crack spacing and crack width.

308

The distributed cracking approach using the Cracked Membrane Model provides a computationally economical tool that can be used to accurately model the timedependent behaviour of reinforced concrete structures since it requires relatively few elements in a numerical analysis. However, due to the strong dependence of the model on crack spacing for the calculation of crack width, the tension chord model, which is used to calculate crack spacing, needs to be further calibrated with more experimental data from various types of structures. For the modelling of a reinforced concrete structure using the localized cracking approach, a concrete fracture model must be complemented by the use of an appropriate bond element (for example the bond-slip interface element employed this study) that can simulate the slip between concrete and steel in the vicinity of a crack, thereby facilitating the opening of a crack. On the other hand, the use of a bond element without a reliable concrete fracture model (for example, cracking model with a sudden drop in stress to zero) can render the computation outcomes unobjective. The model may still be able to compute localized cracking without the use of a concrete fracture model. However, since the fracture energy dissipation in the fracture process zone of the concrete is not properly accounted for, the computed results will be highly sensitive to the finite element mesh configuration. The non-local model is widely known as one of the most effective tools for alleviating mesh sensitivity in a standard smeared crack model. However, most of the previous research relating to non-local models focuses mainly on fracture of plain concrete. The applicability of such models to reinforced concrete structures is often overlooked. This study has revealed the drawback of the non-local model in the context of reinforced concrete structures. The fracture zone of a non-local model stretches over several elements and the non-linear distribution of cracking strain within the fracture zone inevitably handicaps the computation of crack width. Moreover, the sensitivity of spatial averaging to high tensile stress regions in the model also prevents a crack from forming nicely into a single discontinuity. As a result the computed crack pattern in some cases can be rather dispersed and the localized cracks are difficult to identify. In contrast, the crack band model, which is less sophisticated than the non-local model, can better describe localized cracking in reinforced concrete structures. In

309

addition, the bond-slip interface element also works more effectively with the crack band model as shown in Section 4.4.4. An interesting finding for the crack band model is the stress-locking phenomenon observed in concrete fracture tests becomes insignificant when modelling a reinforced concrete structure. This indicates that stresslocking in the crack band model is only a plain concrete fracture problem and the problem is eliminated by the stabilizing effect of reinforcing steel. From the observation of the parametric investigation presented in Chapter 5, it is concluded that the contact surface area between the tensile steel and the concrete in a reinforced concrete member is an important factor affecting the overall bond characteristic in the tension zone of the member. Larger concrete-steel contact surface area can improve the bond characteristic of a member. The permissible maximum bar spacing stipulated in the simplified crack control procedures of many design codes is in fact not a critical factor for crack spacing and crack width. In the authors opinion, the objective of the limitation on tension bar spacing is to infer the use of more reinforcing bars of smaller size so as to provide a larger concrete-steel contact surface area and thus reduce crack spacing and crack width. Therefore, the author recommends the use of a more explicit requirement such as concrete-steel contact surface area per unit length for detailing of tensile reinforcing bars as a substitute for the currently stipulated maximum bar spacing in design codes of practice. The effective reinforcement ratio (i.e. the area of tensile steel divided by the effective area of concrete in tension) is one of the important factors that affects the stress transfer rate between the steel and the concrete in a cracked reinforced concrete member. This in turn has a marked effect on crack spacing. In addition to quantities of tensile reinforcement, bottom concrete cover is an important parameter in determining the effective concrete ratio. For a reinforced concrete member with a thicker concrete cover, the crack spacing is relatively larger since the effective reinforcement ratio is lower than a member with a smaller concrete cover. Shrinkage, creep, bond creep and load history are the main factors causing timedependent crack opening in reinforced concrete members. This study has confirmed the well-known influence of shrinkage on crack width. Deterioration of bond under sustained load results in increase in slip between concrete and steel with time. The

310

effect of creep is less prominent on crack width. However, due to its slight influence on the increase in tensile steel stress with time, the effect of creep cannot be totally overlooked since large tensile steel stress can also lead to unsightly wide cracks. On the other hand, the load history of a structure is important due to the fact that it changes the stress in the tensile steel, which effectively influences the crack width. This study has demonstrated that the replacement of 400 MPa steel reinforcement by the new 500 MPa steel reinforcement in AS 3600 (2001) can adversely lead to reduction in the required quantity of tensile reinforcement in a member, which results in undesirable increase in tensile steel stress and a possible tendency of using fewer bars of larger diameters. Consequently, a revision of the current crack control procedures in AS 3600 (2001) should be undertaken immediately.

6.3 Recommendations for Future Research


It has been demonstrated in this thesis that the numerical models developed in Chapter 3 and verified in Chapter 4 can accurately model time-dependent behaviour and trace time-dependent development of cracks of reinforced concrete structures. In Chapter 5, the model has been employed to investigate time-dependent cracking of reinforced concrete members and has provided an insight into the influence and interaction of the parameters affecting time-dependent cracking. However, this does not mean that the task is complete or the problem is solved, a great deal of work remains to be done. Some of the recommended future work is outlined below: Development of a three-dimensional finite element model for concrete, steel and bond in order to study the three-dimensional bond mechanism and the interaction of bond and effective tensile area of concrete in a cracked reinforced concrete member. This may provide a better understanding of the formation of cracks and the effects of bond. An intra-element discrete crack model to simulate cracking in reinforced concrete structures should be developed. The intra-element crack model would simulate a

311

crack as a complete stress-free discontinuity and would not require a remeshing technique to trace the crack trajectory. Modification of the algorithm for the generation of stochastic concrete tensile strengths based on the normal distribution in the localized cracking model. Extension of the finite element model for moisture migration analysis using the non-linear diffusion theory so as to obtain shrinkage at any point in a continuum taking into consideration boundary and size effects. A separate treatment for drying creep by employing a unified approach for longterm aging and drying of concrete such as the microprestress-solidification theory (Baant et al., 1997), which is an improved version of the solidification theory of creep adopted in this study. Further calibration of the crack spacing equation of the tension chord model with more experimental data from different types of structures. The use of different time-dependent bond models and investigation of the effect of each model on crack opening with time. Incorporation of an automatic time-stepping algorithm in the time-dependent finite element model so that the size of the time step can be adjusted automatically according to the state of deformation of the structure. Inclusion of thermal effects as a component in the time-dependent analysis of a reinforced concrete structure.

312

APPENDIX A: FE Implementation of Rate of Creep Method

The rate of creep method (RCM) has been used to model the time-dependent development of creep strain (Gilbert, 1988; Chong et al., 2004). The RCM requires as input a single creep coefficient versus time curve associated with the initial application of load. For all subsequent loadings, the creep coefficient versus time curve is assumed to be affine with that at first loading. That is, the rate of change of creep is assumed to be independent of the age at first loading. This, of course, is approximate and introduces some errors, particularly when the change of concrete stress with time is significant. However, the advantage is that the storage of the time-dependent stress history is not required. Similar to the solidification theory of creep presented in Chapter 3, two assumptions are made: (i) creep is linear with respect to stress; and (ii) the timedependent response in tension is identical to that in compression. For the implementation of the RCM into the finite element model, time is discretized into small intervals and loads are taken to remain constant during each time increment. The creep strain at the current time is obtained by summing the increments of creep obtained from the previous time intervals, that is

cp (t ) =

(t ) Ec (t 0 )

(A.1)

where is the change in creep coefficient during a particular time interval. The change of creep strain at time t is cp (t ) = ce (t ) (t ) (A.2)

where ce (t ) is the concrete elastic strain at time t. For the finite element implementation, Eq. A.2 is written as cp12 (t i ) = c12 (t i ) (t i ) (A.3)

313

where cp12 (t i ) is the change in creep strain in the principal strain directions and c12 (t i ) is the concrete elastic principal strain vector. Note that the change in creep coefficient (t i ) is a scalar. Before proceeding to the computation of creep strain of the current time step, the concrete global elastic strain vector ce (t i ) = [ cx (t i ) cy (t i ) cxy (t i )]T as

determined from the last time step is stored. The elastic strain vector in global directions is then transformed through an angle using the strain transformation matrix T to the elastic strain vector in principal directions giving c12 (t i ) = T ce (t i ) (A.4)

where c12 (t i ) = [ c1 (t i ) c 2 (t i ) 0]T is the principal concrete elastic strain. The change of creep strain in the current time step is calculated from Eq. A.3, in which the change of creep coefficient is given by (t i ) = (t i ) (t i 1 ) (A.5)

The change of creep strain cp12 (t i ) is transformed through to the global directions restoring strains to the global directions. The iterative procedures for the calculation of the change of creep strain over a time step is performed for each integration point of each element and is shown in Figure A.1. Lastly, the total creep strain is calculated by adding the change of creep strain cp (t i ) in the global directions to the stored total creep strain from the last time step at time t i 1 . That is cp (t i ) = cp (t i ) + cp (t i 1 ) The current material stress state is calculated from (t i ) = D c [(t i ) 0 (t i )]
314

(A.6)

(A.7)

in which the inelastic pre-strain vector 0 is the sum of the creep strain component calculated from Eq. A.6 and a shrinkage strain component (see Section 3.6.3 of Chapter 3). In the finite element implementation, the inelastic pre-strain vector is converted into a set of equivalent nodal forces and applied to the discretized structures. The details of the finite element implementation have been discussed in Section 3.7 of Chapter 3).

cy
y x

c2 cxy
Transform to directions
2

c1
1

cx Principal

cpy
y x

cp2 cpxy
Transform to directions
2

Calculate creep strain from instantaneous strain

cp1

cpx Global

Fig. A.1 - Computation of creep strain from elastic strain.

To verify the numerical treatment of creep using RCM, specimen B2-a tested by Gilbert and Nejadi (2004) was selected and analysed using the localized cracking cracked band model (see Section 3.4.1 of Chapter 3). The finite element mesh for the beam is shown in Figure A.2 and the mesh is made up of 665 nodes, 540 concrete elements, 54 steel elements and 54 bond-slip elements. The steel elements were connected to the concrete elements via bond-slip interface elements and the free edge node of the steel bar was rigidly connected to the concrete element node to simulate anchorage of the bar. A 10% random fluctuation of the mean concrete tensile strength

315

Fig. A.2 - Finite element mesh for beam B2-a.

was assigned to the concrete elements. The material properties for the concrete were: f cm = 25 MPa , Ec = 25 GPa , f ct = 2.0 MPa ( 10% ) , G f = 75 N/m , hc = 35 mm ,

= 0.2 . For the time-dependent properties of concrete, the development of shrinkage


strain and creep coefficient were approximated using Eq. 3.46 in Chapter 3 giving Ash = 950 , Bsh = 45 days , Acp = 1.5 and Bcp = 24 days , and the comparison between the approximated models with test data is shown in Figure A.3. For the reinforcing steel, a nominal yield strength of 500 MPa and elastic modulus of 200 GPa were employed. The bond-slip parameters were: s1 = s2 = 0.6 mm , s3 = 1.0 mm ,

max = 10.0 MPa , f = 1.5 MPa and k u = 100 MPa mm .

1.6 1.4

Creep coeficient

1.2 1.0 0.8 0.6 0.4 0.2 0.0 0 50 100 150 200 Experimental data Model

Shrinkage strain ( )

Acp = 1.5 Bcp = 24 days

1000 800 600 400 200 0 0

Ash = 950 Bsh = 45 days

Experimental data Model

Time since first loading (days)

50 100 150 200 Time since commencement of drying (days)

(a)

(b)

Fig. A.3 - Creep and shrinkage measurements of Gilbert and Nejadi (2004) compared with approximated models: (a) creep coefficient versus time since first loading; (b) shrinkage strain versus time since commencement of drying.

316

The calculated mid-span deflection versus time curve is compared with the experimental results in Figure A.4a with a good correlation. In the constant moment region a comparison is made in Figure A.4b between the calculated and measured crack widths at the soffit of the beam with increasing time. The comparison is presented for the variation of crack width with time for the widest crack and for the average of crack width as observed in the test and as predicted by the model. The agreement is reasonable, but the model calculated a slightly lower maximum crack width and a higher average crack width compared to the observed results. Figures A.5a and A.5b show the computed crack patterns at instantaneous loading and at 380 days, respectively. It is seen that the computed crack spacing agrees well with the test results as shown in Figure 4.28 of Chapter 4.

14 Midspan deflection (mm) 12 10 8 6 4 2 0 0

Beam B2-a

0.40 0.35

Beam B2-a

maximum

Crack width (mm)

0.30 0.25 0.20 0.15 0.10 0.05 0.00


average Experimental FEM

Experimental FEM 100 200 Time (days) 300 400

100

200

300

400

(a)

Time (days)

(b)

Fig. A.4 - Comparison of calculated time-dependent behaviour and test data: (a) midspan deflection with time; (b) crack widths with time.

(a)

(b)

Fig. A.5 - Crack pattern of beam B2-a: (a) and (b) crack patterns (cracking strain plot) at instantaneous loading and at 380 days, respectively.

317

APPENDIX B: Illustration of Treatment for Inelastic Prestrain by Simple Hand Calculation

The behaviour of a material is often characterized by the stress-strain relationship of the material, in which the state of stress is related through a relationship to the state of strain of a structure. However, in a time-dependent analysis, when the effects of creep and shrinkage are prominent, the state of stress in a structure cannot be related directly by a stress-strain law to the state of strain as the strain now contains the creep and shrinkage components. This section aims at describing the algorithm for including the time-dependent components, namely creep and shrinkage of concrete, into a finite element model by presenting a simple illustrative example which can be analysed by hand calculation. To facilitate this task, the finite element implementation flowchart presented in Figure 3.17 of Chapter 3 is simplified and is shown in Figure B.1. The iterative procedure for a finite element stiffness analysis is reduced to an iterative stress-strain analysis where the computation of stiffness matrix is not required. For simplicity, no attempt is made to include the material non-linearity arising from non-linear stress-strain relationship of concrete and this assumption is justified for a concrete structure under service load conditions. In addition, the rate of creep method presented in Appendix A is employed for the computation of creep strain in this section. In Figure B.1 the calculation begins with a short-term analysis and time is set to t = 0 . In the first iterative loop, the creep strain cp , shrinkage strain sh and the total strain are all zero. Therefore, the instantaneous strain and the internal stress int are also zero and the out-of-balance stress out in step 6 of the flowchart is directly equal to the applied external stress ext . Having known out , the strain increment can be calculated from a stress-strain relationship and the total strain is determined by adding to the total strain obtained from the previous iterative loop. The next step is to check convergence for the out-of-balance stress and to decide whether more iteration is required or to proceed to the next time step. If the prescribed convergence

318

tolerance is not satisfied, the process is transferred back to step 3 of the flowchart. The total pre-strain 0 is calculated by summing creep strain cp and shrinkage strain sh obtained based on the current time and the current state of stress. The instantaneous strain inst is computed by subtracting the pre-strain from the total strain and the internal stress of the structure is calculated using the stress-strain relationship as shown in step 5. Out-of-balance stress out is recalculated and the iterative process from step 3 to step 9 continues until the prescribed convergence tolerance is reached. After the equilibrium state is sought, calculation can proceed to the next time step or it can be terminated as indicated in step 10.

Apply external stress ext and initialise short-term analysis t = 0; t = 0

Increase time instance t = t + t

Proceed to next time step or STOP calculation

10

Determine pre-strains due to creep and shrinkage 0 = cp + sh

Calculate out-of-balance stress out = ext int


3

Determine instantaneous strain inst = 0

Determine strain increment = out / Ec


4

Tolerance not reached

Check convergence tolerance out / out.max < % tol. ?

Fig. B.1 - Calculation flowchart for time-dependent analysis.

319

Tolerance reached

Calculate internal stress int = inst Ec

Calculate total strain = +

To illustrate the algorithm, a plain concrete column subjected to sustained axial compression as shown in Figure B.2 is considered. The dimension of the column is 300 mm by 300 mm by 6000 mm long. The axial load and the properties of concrete are shown in Figure B.2 and in Table B.1. In this example the instantaneous analysis at t1 = 0 days is followed by two time steps t1 = 10 days and t 2 = 40 days giving t 2 = 10 days and t 3 = 50 days . The convergence tolerance is taken as 1%.

P = 1000 kN Ac Ac = 90000 mm2 Ec = 25000 MPa

300

300

Section A-A (b) (a) Fig. B.2 - Details of the plain concrete column subjected to sustained axial compression: (a) longitudinal loading configuration; (b) section A-A.

Table B.1 - Time-dependent concrete properties. Concrete parameter 0 days Creep coefficient, Shrinkage strain, sh (10 )
-6

Time, t 10 days 0.5 -100 50 days 1.5 -300 0 0

320

Initialise short-term analysis Step 1: Step 2:

ext =

1000 kN P = = 11.11 MPa Ac 90000 mm 2

t = 0 day iteration 1

Time t = 0 day Step 3: Step 4: Step 5: Step 6: Step 7: Step 8: Step 9:

0 = 0 inst = 0 int = 0 out = 11.11 MPa 0 = 11.11 MPa


= 11.11 MPa = 444.44 10 6 25000 MPa

= 0 + (444.44 10 6 ) = 444.44 10 6 out out. max = 100% > % tolerance (convergence not satisfied)
iteration 2

Time t = 0 day Step 3: Step 4: Step 5: Step 6: Step 7: Step 8: Step 9:

0 = 0 inst = 444.44 10 6 0 = 444.44 10 6 int = 444.44 10 6 25000 MPa = 11.11 MPa out = 11.11 MPa (11.11 MPa) = 0 MPa
= 0

= 444.44 10 6 + 0 = 444.44 10 6 out out. max = 0% < % tolerance (convergence satisfied)


Result for t = 0 day: = 444.44 10 6

Step 10: Step 2:

Proceed to the next time step t = t + t1 = 0 day + 10 days = 10 days iteration 1

Time t = 10 days Step 3:

0 =

11.11 MPa 0.5 + (100 10 6 ) = 322.22 10 6 25000


321

Step 4: Step 5: Step 6: Step 7: Step 8: Step 9:

inst = 444.44 10 6 (322.22 10 6 ) = 122.22 10 6 int = 122.22 10 6 25000 MPa = 3.06 MPa out = 11.11 MPa (3.06 MPa) = 8.06 MPa
= 8.06 MPa = 322.22 10 6 25000 MPa

= 444.44 10 6 + (322.22 10 6 ) = 766.67 10 6 out out. max = 100% > % tolerance (convergence not satisfied)
iteration 2

Time t = 10 days Step 3: Step 4: Step 5: Step 6: Step 7: Step 8: Step 9:

0 =

11.11 MPa 0.5 + (100 10 6 ) = 322.22 10 6 25000

inst = 766.67 10 6 (322.22 10 6 ) = 444.44 10 6 int = 444.44 10 6 25000 MPa = 11.11 MPa out = 11.11 MPa (11.11 MPa) = 0 MPa
= 0

= 766.67 10 6 + 0 = 766.67 10 6 out out. max = 0% < % tolerance (convergence satisfied)


Result for t = 10 days: = 766.67 10 6

Step 10: Step 2:

Proceed to the next time step t = t + t 2 = 10 days + 40 days = 50 days iteration 1

Time t = 50 days Step 3: Step 4: Step 5: Step 6: Step 7: Step 8:

0 =

11.11 MPa 1.5 + (300 10 6 ) = 966.67 10 6 25000

inst = 766.67 10 6 (966.67 10 6 ) = 200.00 10 6 int = 200.00 10 6 25000 MPa = 5.00 MPa out = 11.11 MPa 5.00 MPa = 16.11 MPa
= 16.11 MPa = 644.44 10 6 25000 MPa

= 766.67 10 6 + (644.44 10 6 ) = 1411.11 10 6


322

Step 9:

out out. max = 100% > % tolerance (convergence not satisfied)


iteration 2

Time t = 50 days Step 3: Step 4: Step 5: Step 6: Step 7: Step 8: Step 9:

0 =

11.11 MPa 1.5 + (300 10 6 ) = 966.67 10 6 25000

inst = 1411.11 10 6 (966.67 10 6 ) = 444.44 10 6 int = 444.44 10 6 25000 MPa = 11.11 MPa out = 11.11 MPa (11.11 MPa) = 0 MPa
= 0

= 1411.11 10 6 + 0 = 1411.11 10 6 out out. max = 0% < % tolerance (convergence satisfied)


Result for t = 50 days: = 1411.11 10 6

Step 10:

Stop calculation

The same example is analysed using a direct approach in which the total deformation is calculated as the sum of elastic strain, creep strain and shrinkage strain and the results are shown in Table B.2. It is seen that the approach presented in Figure B.1 and the direct approach both give the same answers. Although the approach shown Figure B.1 may seem cumbersome, the algorithm is useful for the analysis of large nonlinear problems using the finite element method. Furthermore, with the same algorithm, the analysis method of Figure B.1 or, a more complete version as shown in Figure 3.17 of Chapter 3, can be extended to include pre-strain in reinforcing steel and expansion of concrete due to heat effects.

Table B.2 - Calculation of deformation using the rate of creep method. Time Instantaneous strain t (days) inst 0 10 50 -444.44 10
-6

Creep coef.

Creep strain

Shrinkage strain

Total Strain

-444.44 10-6 -444.44 10-6

(t) 0 0.5 1.5

cp(t) 0 -222.22 10-6 -666.67 10-6

sh(t) 0.5 -100 10-6 -300 10-6

(t) = inst + cp(t) + sh(t) -444.44 10-6 -766.67 10-6 -1411.11 10-6

323

APPENDIX C: CEB-FIP Model Code 1990 Creep and Shrinkage Models

In the CEB-FIP Model Code 1990 (1993), the models for creep and shrinkage are provided for the prediction of the mean behaviour of a concrete cross-section. The models are valid for structural concrete with compressive strengths of 12 MPa to 80 MPa subjected to a compressive stress less than 40% of the compressive strength and exposed to mean relative humidity in the range of 40% to 100% at mean temperature of between 5C and 30C.

Creep Model The compliance function J (t , t ' ) , which is defined as the strain at time t produced by a unit stress applied at t ' and is given by J (t , t ' ) =

(t , t ' ) 1 + E c (t ' ) E c.28

(C.1)

where t ' is the variable for age of loading, (t , t ' ) is the creep coefficient, E c.28 is the elastic modulus of concrete at age 28 days in MPa and E c (t ' ) is the elastic modulus of concrete (MPa) at the age of loading which can be estimated by
0.5 28 E c (t ) = E c.28 exp s c s c t 0.5

(C.2)

where time t is in days and sc is a coefficient depending on the type of cement and is taken as sc = 0.2 for rapid hardening high strength cement, sc = 0.25 for normal and rapid hardening cements and sc = 0.38 for slow hardening cement. The creep coefficient (t , t ' ) at time t (days) for a concrete loaded at age t ' (days) in Eq. C.1 is given by

324

(t , t ' ) = RH 1 ( f cm ) 2 (t ' ) cp (t t ' )


in which

(C.3)

RH = 1 +

1 0.01RH 0.46 (0.01hs )1 3 5.3 (0.1 f cm ) 0.5 1

(C.4)

1 ( f cm ) =

(C.5)

2 (t ' ) =

0.1 + t ' 0.2


0.3

(C.6)

t t' cp (t t ' ) = H + (t t ' ) with

(C.7)

H = 1.5hs 1 + (0.012 RH )18 + 250 1500

(C.8)

where RH is the relative humidity of the ambient environment in percentage, f cm is the mean compressive strength of concrete in MPa and hs is a notational size of a member in mm which is defined as hs = 2 Ac u c (C.9)

where Ac is the cross-section area of the member and uc is the perimeter of the member exposed to the air.

325

Shrinkage Model The shrinkage strain at time t (days) for a concrete member commences drying at time t sh (days) can be calculated from

sh (t , t sh ) = shs ( f cm ) RH sh (t t sh )
where

(C.10)

RH = 1.55 1 + (0.01RH ) 3 RH = 0.25

for 40% RH < 99% for RH 99%


0.5

(C.11) (C.12)

t t sh sh (t t sh ) = 350 (0.01hs ) 2 + (t t sh )

(C.13)

shs ( f cm ) = { + 10 shc (9 0.1 f cm )} 10 6 160

(C.14)

where RH, f cm and hs are identical to those defined for the creep model presented previously and shc is a coefficient accounting for the type of cement and is taken as

shc = 4 for slow hardening cement, shc = 5 for normal or rapid hardening cement
and shc = 8 for rapid hardening high strength cement.

326

REFERENCES
ACI Committee 209 Report No. ACI 209 R-82. (1982). Prediction of creep shrinkage and temperature effects in concrete structures. ACI Special Publication SP-76, D. J. Carreira, Z. P. Baant, and D. E. Branson, eds., ACI, Detroit, Michigan, pp. 193-300. Al-Mahaidi. (1979). Nonlinear finite element analysis of reinforced concrete deep members. Research Report No. 79-1, Department of Structural Engineering, Cornell University. AS 3600. (2001). Australian Standard for Concrete Structures, Standards Australia, Sydney. AS/NZS 4671. (2001). Australian/New Zealand Standard for Steel Reinforcing Materials, Standards Australia, Sydney. ASCE Task Committee. (1982). State-of-the-Art Report on Finite Element Analysis of Reinforced Concrete, ASCE, New York. Attard, M. M., Nguyen, D. M., and Foster, S. J. (1996). Finite element analysis of out of plane buckling of reinforced concrete walls. Computers and Structures, 61(6), pp. 1037-1042. Bakoss, S. L., Gilbert, R. I., Faulkes, K. A., and Pulmano, V. A. (1982). Long-term tests on reinforced concrete beams. UNICIV Report No. R-204, School of Civil Engineering, The University of New South Wales, Australia. Balakrishnan, S., and Murray, D. W. (1988). Concrete constitutive model for NLFE analysis of structures. Journal of Structural Engineering, ASCE, 114(7), pp. 1449-1466. Balakrishnan, S., and Murray, D. W. (1986). Finite element prediction of reinforced concrete behaviour. Structural Engineering Report No. 138, University of Alberta, Edmonton, Alberta, Canada.

327

Bangash, M. Y. H. (2001). Manual of Numerical Methods in Concrete: Modelling and applications validated by experimental and site-monitoring data, Thomas Telford, London. Barpi, F., and Valente, S. (2003). Creep and failure in concrete: a fractional order rate approach. Engineering Fracture Mechanics, 70, pp. 611-623. Barpi, F., and Valente, S. (2001). Time-dependent fracture of concrete using fractional order rate laws. Fracture Mechanics of Concrete Structures, R. de Borst, J. Mazars, G. Pijaudier-Cabot, and J. G. M. Van Mier, eds., Swets & Zeitlinger, Lisse, pp. 153-159. Base, G. D., Read, J. B., Beeby, A. W., and Taylor, H. P. J. (1966). An investigation of the crack conctrol characteristics of various types of bar in reinforced concrete beams. Research Report No. 18, Part 1, Cement and Concrete Association, London. Batdorf, S. B., and Budianski, B. (1949). A mathematical theory of plasticity based on the concept of slip. Technical Note No. 1871, National Advisory Committee for Aeronautics, Washington D. C. Bathe, K. J. (1996). Finite Element Procedures, Prentice Hall, New Jersey. Baant, Z. P. (1971). Numerical stable algorithm with increasing time steps for integral-type ageing creep. The First International Conference on Structural Mechanics in Reactor Technology, West Berlin, Paper H2/3. Baant, Z. P. (1972). Numerical determination of long-range stress history from strain history in concrete. Materials and Structures, RILEM, 5(27), pp. 135-141. Baant, Z. P. (1976). Instability, ductility and size effect in strain softening concrete. Journal of the Engineering Mechanics Division, ASCE, 102, pp. 331-344. Baant, Z. P. (1979). Advanced topics in inelasticity and failure of concrete. CBI Special Publications, Swedish Cement and Concrete Research Institute, Stockholm.

328

Baant, Z. P. (1982). Mathematical models for creep and shrinkage of concrete. Symposium on Fundamental Research on Creep and Shrinkage of Concrete, Swiss Federal Institute of Technology, Lausane, pp. 163-256. Baant, Z. P. (1983). Comment on orthotropic models for concrete and geomaterials. Journal of Engineering Mechanics, ASCE, 109(3), pp. 849-865. Baant, Z. P. (1984). Imbricate continuum and its variational derivation. Journal of Engineering Mechanics, ASCE, 110, pp. 1693-1712. Baant, Z. P. (1986). Distributed cracking and nonlocal continuum. Finite Element Methods for Nonlinear Problems, Bergan, Bathe, and Wunderlich, eds., Springer, Berlin, pp. 77-102. Baant, Z. P. (1988). Material models for structural creep analysis. Mathematical Modeling of Creep and Shrinkage of Concrete, Z. P. Baant, ed., John Wiley & Sons, Inc., New York, pp. 99-215. Baant, Z. P. (1990). Recent advances in failure localization and nonlocal models. Micromechanics of Failure of Quasi-brittle Materials, S. P. Shah, S. E. Swartz, and M. L. Wang, eds., Elsevier, London, pp. 12-32. Baant, Z. P. (1993). Current status and advances in the theory of creep and interaction with fracture. Creep and Shrinkage of Concrete, Z. P. Baant and I. Carol, eds., E & FN Spon, London, pp. 291-307. Baant, Z. P., and Baweja, S. (1995a). Creep and shrinkage prediction model for analysis and design of concrete structures - model B3. Materials and Structures, RILEM, 28, pp. 357-365. Baant, Z. P., and Baweja, S. (1995b). Justification and refinements of Model B3 for concrete creep and shrinkage 1. statistics and sensitivity. Materials and Structures, RILEM, 28, pp. 415-430. Baant, Z. P., Belytschko, T. B., and Chang, T.-P. (1984). Continuum model for strain softening. Journal of Engineering Mechanics, ASCE, 110(12), pp. 1666-1692.

329

Baant, Z. P., Caner, F. C., Carol, I., Adley, M. D., and Akers, S. A. (2000). Microplane model M4 for concrete. (I: Formulation with work-conjugate deviatoric stress, II: Algorithm and calibration). Journal of Engineering Mechanics, ASCE, 126(9), pp. 944-961. Baant, Z. P., and Cedolin, L. (1979). Blunt crack band propagation in finite element analysis. Journal of the Engineering Mechanics Division, ASCE, 105(EM2), pp. 297-315. Baant, Z. P., and Chern, J. C. (1985). Concrete creep at variable humidity: constitutive law and mechanism. Materials and Structures, RILEM, 18, pp. 1-20. Baant, Z. P., and Gambarova, P. G. (1980). Rough cracks in reinforced concrete. Journal of the Structural Division, ASCE, 106(ST4), pp. 819-842. Baant, Z. P., and Gambarova, P. G. (1984). Crack shear in concrete: crack band microplane model. Journal of Structural Engineering, ASCE, 110(9), pp. 2015-2035. Baant, Z. P., and Gettu, R. (1992). Rate effects and load relaxation in static fracture of concrete. ACI Materials Journal, 89(5), pp. 456-468. Baant, Z. P., Hauggaard, A. B., Baweja, S., and Ulm, F.-J. (1997). Microprestresssolidification theory for concrete creep. (I: Aging and drying effects, II: Algorithm and verification). Journal of Engineering Mechanics, ASCE, 123(11), pp. 1188-1201. Baant, Z. P., and Jirsek, M. (2002). Nonlocal integral formulations of plasticity and damage: survey of progress. Journal of Engineering Mechanics, ASCE, 128(11), pp. 1119-1149. Baant, Z. P., and Kim, S. S. (1979). Approximate relaxation function for concrete. Journal of the Structural Division, ASCE, 105(ST12), pp. 2695-2705. Baant, Z. P., and Lin, F.-B. (1988a). Nonlocal smeared cracking model for concrete fracture. Journal of Structural Engineering, ASCE, 114(11), pp. 2493-2510.

330

Baant, Z. P., and Lin, F.-B. (1988b). Nonlocal yield-limit degradation. International Journal for Numerical Methods in Engineering, 26, pp. 1805-1823. Baant, Z. P., and Oh, B. H. (1983). Crack band theory for fracture of concrete. Materials and Structures, RILEM, 16(93), pp. 155-177. Baant, Z. P., and Obolt, J. (1990). Nonlocal microplane model for fracture, damage, and size effect in structures. Journal of Engineering Mechanics, ASCE, 116(11), pp. 2485-2505. Baant, Z. P., and Panula, L. (1978). Practical prediction of time-dependent deformations of concrete. Materials and Structures, Research and Testing, RILEM, 11(65), pp. 307-328. Baant, Z. P., and Panula, L. (1980). Creep and shrinkage characterization for analyzing prestressed concrete structures. Journal of the Prestressed Concrete Institute, 25(3), pp. 86-122. Baant, Z. P., and Panula, L. (1982). New model for practical prediction of creep and shrinkage. ACI Special Publication SP-76, ACI, Detroit, Michigan, pp. 7-23. Baant, Z. P., and Pijaudier-Cabot, G. (1988). Nonlocal continuum damage, localization instability and convergence. Journal of Applied Mechanics, 55, pp. 287-293. Baant, Z. P., and Pijaudier-Cabot, G. (1989). Measurement of characteristic length of nonlocal continuum. Journal of Engineering Mechanics, ASCE, 115(4), pp. 755-767. Baant, Z. P., and Prasannan, S. (1989a). Solidification theory of concrete creep. I: Formulation. Journal of Engineering Mechanics, ASCE, 115(8), pp. 1691-1703. Baant, Z. P., and Prasannan, S. (1989b). Solidification theory of concrete creep. II: Verification and application. Journal of Engineering Mechanics, ASCE, 115(8), pp. 1704-1725.

331

Baant, Z. P., and Prat, P. C. (1988). Microplane model for brittle - plastic material. (I: Theory, II: Verification). Journal of Engineering Mechanics, ASCE, 114(10), pp. 1672-1702. Baant, Z. P., and Ralfshol, W. J. (1982). Effect of cracking in drying and shrinkage specimens. Cement and Concrete Research, 12, pp. 209-226. Baant, Z. P., and Wu, S. T. (1973). Dirichlet series creep function for aging concrete. Journal of the Engineering Mechanics Division, ASCE, 99(EM2), pp. 367-387. Baant, Z. P., and Wu, S. T. (1974). Rate-type creep law of aging concrete based on Maxwell chain. Materials and Structures, RILEM, 7, pp. 45-60. Baant, Z. P., and Xi, Y. (1995). Continuous retardation spectrum for solidification theory of concrete creep. Journal of Engineering Mechanics, ASCE, 121(2), pp. 281-288. Baant, Z. P., and Xiang, Y. (1997). Crack growth and lifetime of concrete under long time loading. Journal of Engineering Mechanics, ASCE, 123(4), pp. 350-358. Belarbi, A., and Hsu, T. T. C. (1991). Constitutive laws of reinforced concrete in biaxial tension compression. Research Report UHCEE 91-2, Department of Civil Engineering, Houston, Texas. Belytschko, T., Fish, J., and Englandmann, B. E. (1988). A finite element with embedded localization zones. Computer Methods in Applied Mechanics and Engineering, 70, pp. 59-89. Benallal, A., Billardon, R., and Geymonat, G. (1988). Some mathematical aspects of the damage softening problem. Cracking and Damage, J. Mazars and Z. P. Baant, eds., Elsevier, Amsterdam and London, pp. 247-258. Bisschop, J. (2002). Drying shrinkage microcracking in cement-based materials, PhD Thesis, Delft University of Technology, Delft, The Netherlands.

332

Bocca, P., Carpinteri, A., and Valente, S. (1991). Mixed mode fracture of concrete. International Journal of Solids and Structures, 27(9), pp. 1139-1153. Borino, G., Fuschi, P., and Polizzotto, C. (1999). A thermodynamic approach to nonlocal plasticity and related variational approaches. Journal of Applied Mechanics, 66, pp. 952-963. Bradford, M. A. (2005). Viscoelastic response of slender eccentrically loaded reinforced concrete columns. Magazine of Concrete Research. (in press). Bresler, B., and Scordelis, A. C. (1963). Shear strength of reinforced concrete beams. Journal of ACI Proceedings, 60(1), pp. 51-73. Broms, B. B. (1965). Crack width and crack spacing in reinforced concrete members. ACI Journal Proceedings, 62(10), pp. 1237-1255. Brooks, J. J. (1989). Influence of mix proportions, plasticizers and superplasticizers on creep and drying shrinkage of concrete. Magazine of Concrete Research, 41(148), pp. 145-154. Buyukozturk, O. (1977). Nonlinear analysis of reinforced concrete structures. Computers and Structures, 7, pp. 149-156. Carlson, R. W. (1937). Drying shrinkage of large concrete members. Journal of American Concrete Institute, 33, pp. 327-336. ervenka, J. (1994). Discrete Crack Modeling in Concrete Structures, PhD Thesis, University of Colorado, Boulder. ervenka, V. (1970). Inelastic finite element analysis of reinforced concrete panels under in-plane loads, PhD Thesis, University of Colorado, Boulder. ervenka, V. (1985). Constitutive model for cracked reinforced concrete. ACI Journal Proceedings, 82(6), pp. 877-882. Chen, A. T. C., and Chen, W. F. (1975). Constitutive relations for concrete. Journal of Engineering Mechanics, ASCE, 101(4), pp. 465-481.

333

Chen, W. F., and Han, D. J. (1988). Plasticity for Structural Engineers, SpringerVerlag, New York. Chen, W. F., Yamaguchi, E., Kotsovos, M. D., and Pan, A. D. (1993). Constitutive Models. Finite Element Analysis of Reinforced Concrete Structures II: Proceedings of the International Workshop, J. Isenberg, ed., ASCE, New York, pp. 36-117. Chivamit, P. (1965). Creep of plain concrete under axial compression, Thesis No. 97, Bangkok, Thailand. Chong, K. T., Gilbert, R. I., and Foster, S. J. (2004). Modelling time-dependent cracking in reinforced concrete using bond-slip interface elements. Computers & Concrete, 1(2), pp. 151-168. Ciampi, V., Eligehausen, R., Bertero, V. V., and Popov, E. P. (1981). Analytical model for deformed-bar bond under generalized excitations. Trans. IABSE Colloquium on Advanced Mechanics of Reinforced Concrete, Delft, Netherlands. Clark, A. P. (1956). Cracking in reinforced concrete flexural members. ACI Journal Proceedings, 52(8), pp. 851-862. Clark, L. A., and Spiers, D. M. (1978). Tension stiffening in reinforced concrete beams and slabs under short-term load. Technical Report No. 42.521, Cement and Concrete Association, Wexham Springs. Collins, M. P., and Porasz, A. (1989). Shear strength for high strength concrete. Bulletin No. 193, Design Aspects of High Strength Concrete, Comit EuroInternational du Bton (CEB), pp. 75-83. Comit Euro-International du Bton (CEB). (1997). Serviceability Models: Behaviour and Modelling in Serviceability Limit States including Repeated and Sustained Loads. Progress Report, CEB, Lausanne, Switzerland. Comit Euro-International du Bton-Fdration International de la Prcontrainte (CEBFIP). (1978). CEB-FIP Model Code for Concrete Structures, Paris.

334

Comit Euro-International du Bton-Fdration International de la Prcontrainte (CEBFIP). (1993). CEB-FIP Model Code 1990, Thomas Telford, London. Cook, R. D., Malkus, D. S., Plesha, M. E., and Witt, R. J. (2001). Concepts and Applications of Finite Element Analysis, John Wiley & Sons, Inc., New York. Cosserat, E., and Cosserat, F. (1909). Thorie des corps dformables, Herrman, Paris. Crisfield, M. A. (1982). Local instabilities in the nonlinear analysis of reinforced concrete beams and slabs. Proceedings, Institute of Civil Engineers on Part 2, 73, pp. 135-145. Darwin, D. (1993). Reinforced Concrete. Finite Element Analysis of Reinforced Concrete Structures II: Proceedings of the International Workshop, J. Isenberg, ed., ASCE, New York, pp. 203-232. Darwin, D., and Pecknold, D. A. (1977). Non-linear stress-strain law for concrete. Journal of the Engineering Mechanics Division, ASCE, 103(EM2), pp. 229-241. Davis, R. E., Davis, H. E., and Hamilton, J. S. (1934). Plastic flow of concrete under sustained stress. Proceedings of ASTM, 34, pp. 354-386. DD ENV-1992-1-1. (1992). Eurocode 2: Design of Concrete Structures, British Standards Institution. de Borst, R. (1987). Smeared cracking, plasticity, creep and thermal loading - a unified approach. Computer Methods in Applied Mechanics and Engineering, 62, pp. 89-110. de Borst, R. (1997). Some recent developments in computational modelling of concrete fracture. International Journal of Fracture, 86, pp. 5-36. de Borst, R. (2002). Fracture in quasi-brittle materials: a review of continuum damage-based approaches. Engineering Fracture Mechanics, 69, pp. 95-112. de Borst, R., Geers, M. G. D., Kuhl, E., and Peerlings, R. H. J. (1998). Enhanced damage models for concrete fracture. Computational Modelling of Concrete

335

Structures, R. de Borst, N. Bicanic, H. Mang, and G. Meschke, eds., Balkema, Rotterdam, pp. 231-248. de Borst, R., and Mhlhaus, H.-B. (1992). Gradient-dependent plascity: formulation and algorithmic aspects. International Journal for Numerical Methods in Engineering, 35, pp. 521-539. de Borst, R., Mhlhaus, H.-B., Pamin, J., and Sluys, L. J. (1993a). A continuum mechanics approach to concrete fracture. Numerical Models in Fracture Mechanics of Concrete, Wittmann, ed., Balkema, Rotterdam, pp. 115-127. de Borst, R., and Nauta, P. (1985). Non-orthogonal cracks in a smeared finite element model. Engineering Computations, 2, pp. 35-46. de Borst, R., Sluys, L. J., Mhlhaus, H.-B., and Pamin, J. (1993b). Fundamental issues in finite element analysis of localisation of deformation. Engineering Computations, 10, pp. 99-122. de Borst, R., Sluys, L. J., Van den Boogaard, A. H., and Van den Bogert, P. A. J. (1993c). Computational issues in time-dependent deformation and fracture of concrete. Creep and Shrinkage of Concrete, Z. P. Baant and I. Carol, eds., E & FN Spon, London, pp. 309-326. de Vree, J. H. P., Brekelmans, W. A. M., and van Gils, M. A. J. (1995). Comparison of non-local approaches in continuum mechanics. Computers and Structures, 55, pp. 581-588. Desayi, P., and Krishnan, S. (1964). Equation for stress strain curve of concrete. ACI Journal Proceedings, 61(3), pp. 345-350. Divakar, M. P., and Dilger, W. H. (1987). Influence of stirrups on crack spacing. Indian Concrete Journal, 61(3), pp. 73-76. Drr, K. (1980). Ein Beitrag zur Berechnung won Stahlbetonscheiben unter besonderer Bercksichtigung des Verbundverhaltens, PhD Thesis, Univerisity of Darmstadt.

336

Eringen, A. C. (1965). Theory of micropolar continuum. Proceedings of the Ninth Midwestern Mechanics Conference, pp. 23-40. Eringen, A. C., and Edelen, D. G. B. (1972). On nonlocal elasticity. International Journal of Engineering Science, 10, pp. 233-248. fdration internationale du bton (fib). (2000). Bond of Reinforcement in Concrete, International Federation for Structural Concrete (fib), Lausane, Switzerland. Ferry-Borges, J. (1966). Cracking and deformability of reinforced concrete beams. Publications, International Association for Bridge and Structural Engineering, Zurich, 26, pp. 75-95. Filippou, F. C., Popov, E. P., and Bertero, V. V. (1983). Effects of bond deterioration on hysteretic behaviour of reinforced concrete joints. Report No. UCB/EERC. 83/19, Earthquake Engineering Research Center, College of Engineering, University of California, Berkeley, California. Foster, S. J. (1992). The Structural Behaviour of Reinforced Concrete Deep Beams, PhD Thesis, The University of New South Wales, Sydney, Australia. Foster, S. J., and Gilbert, R. I. (1990). Non-linear finite element model for reinforced concrete deep beams and panels. UNICIV Report No. R-275, School of Civil and Environmental Engineering, The University of New South Wales, Sydney. Foster, S. J., and Marti, P. (2002). FE modelling of RC membranes using the CMM formulation. Proceedings of the Fifth World Congress on Computational Mechanics (WCCM V), July 7-12, Vienna, Austria. Foster, S. J., and Marti, P. (2003). Cracked membrane model: FE implementation. Journal of Structural Engineering, ASCE, 129(9), pp. 1155-1163. Gambarova, P. G., and Karako, C. (1982). Shear-confinement interaction at the barto-concrete interface. Bond in Concrete, P. Bartos, ed., Applied Science, London, pp. 82-96.

337

Gambarova, P. G., and Rosati, G. P. (1996). Bond and splitting in reinforced concrete: test results on bar pull-out. Materials and Structures, RILEM, 29, pp. 267-276. Gambarova, P. G., and Rosati, G. P. (1997). Bond and splitting in bar pull-out: behavoural laws and concrete-cover role. Magazine of Concrete Research, 49(179), pp. 99-110. Gambarova, P. G., Rosati, G. P., and Zasso, B. (1989). Steel-concrete bond after concrete splitting. (I: Test results, II: Constitutive laws and interface deterioration). Materials and Structures, RILEM, 22, pp. 35-47, 347-356. Gergely, P., and Lutz, L. A. (1968). Maximum crack width in reinforced concrete flexural members. Causes, Mechanism, and Control of Cracking in Concrete, SP20, American Concrete Institute, Detroit, pp. 87-117. Gerstle, K. H. (1981). Simple formulation of biaxial concrete behaviour. ACI Journal, 78(1), pp. 62-68. Gerstle, W. H., and Xie, M. (1992). FEM modeling of fictitious crack propagation in concrete. Journal of Engineering Mechanics, ASCE, 118(2), pp. 416-434. Ghali, A., Dilger, W., and Neville, A. M. (1969). Time-dependent forces induced by settlement of supports in continuous reinforced concrete beams. ACI Journal, 66, pp. 907-915. Gilbert, R. I. (1979). Time-dependent Behaviour of Structural Concrete Slabs, PhD Thesis, The University of New South Wales, Sydney, Australia. Gilbert, R. I. (1988). Time Effects in Concrete Structures, Elsevier Science Publisher, The Netherlands. Gilbert, R. I. (2002). Creep and Shrinkage Models for High Strength Concrete Proposals for inclusion in AS3600. Australian Journal of Structural Engineering, Institution of Engineers, Australia, 4(2), pp. 95-106. Gilbert, R. I., and Nejadi, S. (2004). An experimental study of flexural cracking in reinforced concrete members under sustained loads. UNICIV Report No. R-435,

338

School of Civil and Environmental Engineering, The University of New South Wales, Sydney. Gilbert, R. I., and Warner, R. F. (1978). Tension stiffening in reinforced concrete slabs. Journal of the Structural Division, ASCE, 104(ST12), pp. 1885-1900. Glanville, W. H. (1933). Creep of concrete under load. The Structural Engineer, 11(2), pp. 54-73. Glasstone, S., Laidler, K. J., and Eyring, H. (1941). The Theory of Rate Processes, McGraw Hill, New York. Goodman, R., Taylor, R., and Brekke, T. (1968). A model for the mechanics of rocks. Soil Mechanics and Foundations, Division 5, pp. 637-659. Griffith, A. A. (1921). The phenomena of rupture and flow in solids. Phil. Trans. Roy. Soc. of London, pp. 163-198 (Sec. 12.1). Han, D. J., and Chen, W. F. (1985). A nonuniform hardening plasticity model for concrete materials. Mechanics of Materials, 4, pp. 283-302. Hansen, W. (1987). Drying shrinkage mechanisms in Portland cement paste. Journal of American Ceramic Society, 70(5), pp. 323-328. Hatt, W. K. (1907). Notes on the effect of time element in loading reinforced concrete beams. Proceedings of ASTM, 7, pp. 421-433. Helmuth, R. A., and Turk, D. H. (1967). The reversible and irreversible drying shrinkage of hardened Portland cement and tricalcium silicate paste. Journal of the PCA Research and Development Laboratories, 9(2), pp. 8-21. Hillerborg, A., Modeer, M., and Peterson, P. E. (1976). Analysis of crack propagation and crack growth in concrete by means of fracture mechanics and finite elements. Cement and Concrete Research, 6, pp. 773-782.

339

Hognestad, E. (1951). A study of combined bending and axial load in reinforced concrete members. Bulletin No. 399, Engineering Experiment Station, University of Illinois, Urbana, Illinois. Hsu, T. T. C., and Zhang, L. X. (1996). Tension stiffening in reinforced concrete membrane elements. ACI Structural Journal, 93(1), pp. 108-115. Ingraffea, A. R., Gerstle, W. H., Gergely, P., and Saouma, V. E. (1984). Fracture mechanics of bond in reinforced concrete. Journal of Engineering Mechanics, ASCE, 110(4), pp. 873-889. Irwin, G. R. (1960). Plastic zone near a crack and fracture toughness. Proceedings of the 7th Sagamore Conference, 63 pp. Jirsek, M. (1998). Nonlocal models for damage and fracture: comparison of approaches. International Journal of Solids and Structures, 35(31), pp. 4133-4145. Jirsek, M., and Zimmermann, T. (1998a). Analysis of rotating crack model. Journal of Engineering Mechanics, ASCE, 124(8), pp. 842-851. Jirsek, M., and Zimmermann, T. (1998b). Rotating crack model with transition to scalar damage. Journal of Engineering Mechanics, ASCE, 124(3), pp. 277-284. Johnson, R. D. (1969). Structural Concrete, McGraw-Hill, London. Kachanov, L. M. (1958). Time of rupture process under creep conditions. Izv. Akad. Nauk. SSR, Otd. Tekh. Nauk., 8, pp. 26-31. Kang, Y. J., and Scordelis, A. C. (1980). Nonlinear analysis of prestressed concrete frames. Journal of the Structural Division, ASCE, 106(2), pp. 445-462. Kaufmann, W. (1998). Strength and Deformations of Structural Concrete Subjected to In-Plane Shear and Normal Forces, PhD Thesis, Swiss Federal Institute of Technology, Zurich, Switzerland. Kaufmann, W., and Marti, P. (1998). Structural concrete: cracked membrane model. Journal of Structural Engineering, ASCE, 124(12), pp. 1467-1475.

340

Klisinski, M., Runesson, K., and Sture, S. (1991). Finite element with inner softening band. Journal of Engineering Mechanics, ASCE, 117(3), pp. 575-587. Krner, E. (1967). Elasticity theory of materials with long range cohesive forces. International Journal of Solids and Structures, 3, pp. 731-742. Kupfer, H., and Gerstle, K. H. (1973). Behaviour of concrete under biaxial stresses. Journal of the Engineering Mechanics Division, ASCE, 99, pp. 552-866. Kupfer, H., Hilsdorf, H. K., and Rsch, H. (1969). Behaviour of concrete under biaxial stresses. ACI Journal, 66(8), pp. 656-666. Lasry, D., and Belytschko, T. (1988). Localization limiters in transient problems. International Journal of Solids and Structures, 24, pp. 581-597. Lemaitre, J., and Chaboche, J.-L. (1990). Mechanics of Solid Materials, Cambridge University Press, Cambridge. L'Hermite, R. (1960). Volume changes of concrete. Proceedings of the 4th International Symposium on the Chemistry of Cement, Washington D.C., 659-694. L'Hermite, R. G. (1978). Quelques problmes mal connus de la technologie du bton. Il Cemento, 75(3), pp. 231-246. Li, Y.-J., and Zimmermann, T. (1998). Numerical evaluation of the rotating crack model. Computers and Structures, 69, pp. 487-497. Liu, T. C. Y., Nilson, A. H., and Slate, F. O. (1972). Biaxial stress-strain relations for concrete. Journal of the Structural Division, ASCE, 98(ST5), pp. 1025-1034. Lofti, H. R., and Shing, P. B. (1994). Analysis of concrete fracture with an embedded crack approach. Computational Modeling of Concrete Structures, H. Mang, N. Bicanic, and R. de Borst, eds., Pineridge Press, Swansea, pp. 343-352. Lorman, W. R. (1940). The theory of concrete creep. Proceedings of ASTM, 40, pp. 1082-1102.

341

Lowes, L. N. (2001). A concrete-steel bond model use in finite modelling of reinforced concrete structures. Finite Element Analysis of Reinforced Concrete Structure, pp. 251-271. Lundgren, K. (1999). Three-dimensional modelling of bond in reinforced concrete, PhD Thesis, Chalmers University of Technology, Gteborg, Sweden. Marti, P., Alvarez, M., Kaufmann, W., and Sigrist, V. (1998). Tension chord model for structural concrete. Structural Engineering International, IABSE, 4/98, pp. 287-298. Martin, H. (1973). On the interrelation among surface roughness, bond and bar stiffness in the reinforcement subject to short-term loading (in German). Deutscher Ausschuss fr Stahlbeton, 228, pp. 1-50. Mazars, J. (1984). Application de la mcanique de l'endommagement au comportement non linare et la rupture du bton de structure, These d'Etat, Universit Paris VI, Paris. Mazars, J., and Pijaudier-Cabot, G. (1989). Continuum damage theory - application to concrete. Journal of Engineering Mechanics, ASCE, 115, pp. 345-365. McHenry, D. (1943). A new aspect of creep in concrete and its application to design. Proceedings of ASTM, 43, pp. 1069-1084. Mehlhorn, G., and Keuser, M. (1985). Isoparametric contact elements for analysis of reinforced concrete structures. Finite Element Analysis of Reinforced Concrete Structures, pp. 329-347. Mensi, R., Acker, P., and Attolou, A. (1988). Schage du bton: analyse et modlisation. Materials and Structures, RILEM, 21(121), pp. 3-12. Mindess, S., and Young, J. F. (1981). Concrete, Prentice-Hall, Inc., Englewood Cliffs, N.J. Mirza, S. M., and Houde, J. (1979). Study of bond stress-slip relationships in reinforced concrete. ACI Journal, 76(1), pp. 19-46.

342

Miyakawa, T., Kawakami, T., and Maekawa, K. (1987). Nonlinear behaviour of cracked reinforced concrete plate element under uniaxial compression. Proceedings of the JSCE, 378, pp. 249-258. Mufti, A. A., Mirza, M. S., McCutcheon, J. O., and Houde, J. (1972). A study of the nonlinear behaviour of structural concrete elements. Proceedings of the Specialty Conference of Finite Element Method in Civil Engineering, Montreal, Canada. Mhlhaus, H.-B. (1991). Continuum models for layered and blocky rock. Comprehensive Rock Engineering, Vol. 2: Analysis & Design Methods, Pergamon Press, Oxford. Mhlhaus, H.-B., de Borst, R., and Aifantis, E. C. (1991). Constitutive models and numerical analyses for inelastic materials with microstructure. Computer Methods and Advances in Geomechanics, G. Beer, J. R. Booker, and J. P. Carter, eds., Balkema, Rotterdam, pp. 377-386. Mhlhaus, H.-B., and Vardoulakis, I. (1987). The thickness of shear bands in granular materials. Geotchnique, 37, pp. 271-283. Murray, D. W., Chitnuyanondh, L., Agho, K. Y., and Wong, C. (1979). A concrete plasticity theory for biaxial stress analysis. Journal of the Engineering Mechanics Division, ASCE, 105(EM6), pp. 989-1006. Nawy, E. G., and Orenstein, G. S. (1970). Crack width control in reinforced concrete two-way slabs. Proceedings, ASCE, 96(ST3), pp. 701-721. Needleman, A. (1988). Material rate-dependence and mesh sensitivity in localization problems. Computer Methods in Applied Mechanics and Engineering, 67, pp. 69-85. Nejadi, S., and Gilbert, R. I. (2004). Shrinkage cracking in restrained reinforced concrete members. UNICIV Report No. R-433, School of Civil and Environmental Engineering, The University of New South Wales, Sydney. Nelissen, L. J. M. (1972). Biaxial testing of normal concrete. HERON, 18(1), pp. 1-90.
343

Neville, A. M. (1995). Properties of Concrete, Longman, Essex. Neville, A. M., Dilger, W. H., and Brooks, J. J. (1983). Creep of plain and structural concrete, Construction Press, Longman Inc., New York. Ngo, D., and Scordelis, A. C. (1967). Finite element analysis of reinforced concrete beams. ACI Journal Proceedings, 64(3), pp. 152-163. Nilson, A. H. (1968). Nonlinear analysis of reinforced concrete by the finite element method. ACI Journal, 65(9), pp. 757-766. Nilsson, C. (1997). Nonlocal strain softening bar revisited. International Journal of Solids and Structures, 34, pp. 4399-4419. Nimura, A. (1991). Experimental research on failure criteria of ultra-high strength concrete under biaxial stress (in Japanese). Summaries of Technical Papers of Annual Meeting, pp. 473-474. Ortiz, M., Leroy, Y., and Needleman, A. (1987). A finite element method for localized failure analysis. Computer Methods in Applied Mechanics and Engineering, 61, pp. 189-214. Obolt, J., and Baant, Z. P. (1996). Numerical smeared fracture analysis: nonlocal microcrack interaction approach. International Journal for Numerical Methods in Engineering, 39, pp. 635-661. Pamin, J., and de Borst, R. (1995). Gradient-enchanced smeared crack models for finite element analysis of plain and reinforced concrete. Proceedings FRAMCOS2: Fracture Mechanics of Concrete Structures. Peerlings, R. H. J., de Borst, R., Brekelmans, W. A. M., de Vree, J. H. P., and Spree, I. (1996). Gradient-enhanced damage for quasi-brittle materials. International Journal for Numerical Methods in Engineering, 39, pp. 3391-3403. Petersson, P. E. (1981). Crack Growth and Development of Fracture Zone in Plain Concrete and Similar Materials. Report No. TVBM-1006, Division of Building Materials, Lund Institute of Technology, Lund, Sweden.

344

Pickett, G. (1946). Shrinkage stresses in concrete. Journal of American Concrete Institute, 42, pp. 165-204, 361-400. Pickett, G. (1956). Effect of aggregate on shrinkage of concrete and a hypothesis concerning shrinkage. ACI Journal, 52(5), pp. 581-590. Pijaudier-Cabot, G., and Baant, Z. P. (1987). Nonlocal damage theory. Journal of Engineering Mechanics, ASCE, 113(10), pp. 1512-1533. Planas, J., Elices, M., and Guinea, G. V. (1993). Cohesive cracks versus nonlocal models: closing the gap. International Journal of Fracture, 63, pp. 173-187. Rabotnov, Y. N. (1969). Creep Problems in Structural Members, North-Holland, Amsterdam. Ragueneau, F., La Borderie, C., and Mazars, J. (2000). Damage model for concretelike materials coupling cracking and friction, contribution towards structural damping: first uniaxial applications. Mechanics of Cohesive-frictional Materials, 5, pp. 607-625. Ramaswamy, A., Barzegar, F., and Voyiadjis, G. Z. (1995). Study of layering procedures in finite-element analysis of RC flexural and torsional elements. Journal of Structural Engineering, ASCE, 121(12), pp. 1773-1783. Rashid, Y. R. (1968). Analysis of prestressed concrete pressure vessels. Nuclear Engineering and Design, 7(4), pp. 334-344. Rehm, G. (1961). On the fundamentals of stee-concrete bond (in German). Deutscher Ausschuss fr Stahlbeton, 138, pp. 1-59. Ross, A. D. (1958). Creep of concrete under variable stress. Journal of ACI Proceedings, 54(9), pp. 739-758. Rots, J. G. (1985). Bond-slip simulations using smeared crack and / or interface elements. Report No. 85-01, Structural Mechanics Group, Department of Civil Engineering, Delft University of Technology.

345

Rots, J. G. (1988). Computational Modeling of Concrete Fracture, PhD Thesis, Delft University of Technology, Delft, The Netherlands. Rots, J. G., Nauta, P., Kusters, G. M. A., and Blaauwendraad, J. (1985). Smeared crack approach and fracture localization in concrete. HERON, 30(1), pp. 1-48. Saenz, L. P. (1964). Discussion of Equation for stress-strain curve of concrete, by Desayi and Krishnan. ACI Journal Proceedings, 61(9), pp. 1229-1235. Saouma, V. E. (1981). Interactive Finite Element Analysis of Reinforced Concrete: A Fracture Mechanics Approach, PhD Thesis, Cornell University. Saouma, V. E., and Ingraffea, A. R. (1981). Fracture mechanics analysis of discrete cracking. Proceedings of IABSE Colloquium in Advanced Mechanics of Reinforced Concrete, Delft, pp. 393-416. Sathikumar, S., Karihaloo, B. L., and Reid, S. G. (1998). A model for ageing viscoelastic tension softening materials. Mechanics of Cohesive-frictional Materials, 3, pp. 27-39. Scanlon, A., and Murray, D. (1974). Time-dependent reinforced concrete slabs deflections. Journal of the Structural Division, ASCE, 100(ST9), pp. 1911-1924. Shima, H., Chou, L. L., and Okamura, H. (1987). Micro and macro models for bond in reinforced concrete. Journal of the Faculty of Engineering, the University of Tokyo, 39(2), pp. 133-194. Sigrist, V. (1995). Verformungsvermgen von Stahlbetontrgern [on deformation capacity of structural concrete girders]. Report No. 210, Institute of Structural Engineering, ETH, Zrich, Switzerland. Simo, J. C., and Oliver, J. (1994). A new approach to the analysis and simulation of strong discontinuities. Fracture and Damage in Quasibrittle Structures, Z. P. Baant, Z. Bittnar, M. Jirsek, and J. Mazars, eds., E & FN Spon, London, pp. 25-39.

346

Sluys, L. J. (1992). Wave propagation, localisation and dispersion in softening solids, PhD Thesis, Delft University of Technology, Delft, The Netherlands. Sluys, L. J., and Berends, A. H. (1998). 2D/3D modelling of crack propagation with embedded discontinuity elements. Computational Modelling of Concrete Structures, R. de Borst, N. Bicanic, H. Mang, and G. Meschke, eds., Balkema, Rotterdam, pp. 399-408. Stevens, N. J., Uzumeri, S. M., Collins, M. P., and Will, G. T. (1991). Constitutive model for reinforced concrete finite element analysis. ACI Structural Journal, 88(1), pp. 49-59. Suidan, M., and Schnobrich, W. C. (1973). Finite element analysis of reinforced concrete. Journal of the Structural Division, ASCE, 99(10), pp. 2109-2122. Tassios, T. P. (1979). Properties of bond between concrete and steel under load cycles idealizing seismic actions. Proc. AICAP-CEP Symposium, CEB Bulletin No. 131, Rome, pp. 67-122. Tasuji, M. E., Nilson, A. H., and Slate, F. O. (1979). Biaxial stress-strain relationships for concrete. Magazine of Concrete Research, 31(109), pp. 217-224. Taylor, G. I. (1938). Plastic strain in metals. Journal of the Institute of Metals, 62, pp. 307-324. Tepfer, R. (1973). A theory of bond applied to overlapped tensile reinforcement splices for deformed bars. Publication 73:2, Chalmer University of Technology, Division of Concrete Structure, Gteborg, Sweden. Tepfer, R. (1979). Cracking of concrete cover along anchored deformed reinforcing bars. Magazine of Concrete Research, 31(106), pp. 3-12. Thorenfeldt, E., Tomaszewicz, A., and Jensen, J. J. (1987). Mechanical properties of high-strength concrete and application in design. Proceedings of the Symposium on Utilization of High-Strength Concrete, Stavanger, Norway.

347

Truesdell, C. (1955). Hypoelasticity. Journal of Rational Mechanics Analysis, 4(1), pp. 83-133. Valente, S. (1995). On the cohesive crack model in mixed-mode conditions. Fracture of Brittle Disordered Materials: Concrete, Rock and Ceramics, G. Bakker and B. L. Karihaloo, eds., E & FN Spon, London, pp. 66-80. van Greunen, J. (1979). Nonlinear geometric, material and time dependent analysis of reinforced and prestressed concrete slabs and panels. Report No. UC SESM 79-3, University of California, Berkeley, California. van Mier, J. G. M. (1986). Fracture of concrete under complex stress. HERON, 31(3), pp. 1-90. van Zyl, S. F., and Scordelis, A. C. (1979). Analysis of curved prestressed segmental bridges. Journal of the Structural Division, ASCE, 105(11), pp. 2399-2417. Vandewalle, L. (1992). Theoretical prediction of the ultimate bond strength between a reinforcement bar and concrete. Bond in Concrete - From Research to Practice, Riga, Latvia, pp. 1-1 to 1-8. Vecchio, F. J. (1989). Nonlinear finite element analysis of reinforced concrete membranes. ACI Structural Journal, 86(1), pp. 26-35. Vecchio, F. J. (2000). Distributed stress field model for reinforced concrete: formulation. Journal of Structural Engineering, ASCE, 126(9), pp. 1070-1077. Vecchio, F. J., and Collins, M. P. (1986). The modified compression field theory for reinforced concrete elements subjected to shear. ACI Journal Proceedings, 83(2), pp. 219-231. Vecchio, F. J., and Collins, M. P. (1989). Compression response of cracked reinforced concrete. Journal of Structural Engineering, ASCE, 119(12), pp. 3590-3610. Volterra, V. (1913). Leons sur les Fonctions de Ligne, Gauthier-Villars, Paris.

348

Volterra, V. (1959). Theory of Functionals and of Integral and Integro-differential Equations, Dover, New York. Wells, G. N., and Sluys, L. J. (2001). A new method for modelling cohesive cracks using finite elements. International Journal for Numerical Methods in Engineering, 50, pp. 2667-2682. Wu, Z. S., and Baant, Z. P. (1993). Finite element modeling of rate effect in concrete fracture with influence of creep. Creep and Shrinkage of Concrete, Z. P. Baant and I. Carol, eds., E & FN Spon, London, pp. 427-432. Zhou, F. P. (1992). Time-dependent crack growth and fracture in concrete, PhD Thesis, Lund University, Sweden. Zhou, F. P. (1993). Cracking analysis and size effect in creep rupture of concrete. Creep and Shrinkage of Concrete, Z. P. Baant and I. Carol, eds., E & FN Spon, London. Zhou, F. P., and Hillerborg, A. (1992). Time-dependent fracture of concrete: testing and modelling. Fracture Mechanics of Concrete Structures, Z. P. Baant, ed., Elsevier Science Publish Ltd., London, England. Zienkiewicz, O. C., and Taylor, R. L. (2000). The Finite Element Method Volume 2: Solid Mechanics, Butterworth-Heinemann, Oxford.

349

Anda mungkin juga menyukai