Sunteți pe pagina 1din 132

Title Nonlinear finite element analysis of reinforced concrete beams

Advisor(s) Lo, SH

Author(s) Yao, Zhong; 姚钟

Yao, Z. [姚钟]. (2013). Nonlinear finite element analysis of


reinforced concrete beams. (Thesis). University of Hong Kong,
Citation Pokfulam, Hong Kong SAR. Retrieved from
http://dx.doi.org/10.5353/th_b5090002

Issued Date 2013

URL http://hdl.handle.net/10722/192857

The author retains all proprietary rights, (such as patent rights)


Rights and the right to use in future works.
NONLINEAR FINITE ELEMENT ANALYSIS OF
REINFORCED CONCRETE BEAMS

YAO Zhong

(姚钟)

M.PHIL. THESIS

THE UNIVERSITY OF HONG KONG


2013
NONLINEAR FINITE ELEMENT ANALYSIS OF
REINFORCED CONCRETE BEAMS

by

YAO Zhong

(姚钟)

B.Eng. (Civil Engineering),


Southwest University

A thesis submitted in partial fulfilment of the requirements for


the Degree of Master of Philosophy
at The University of Hong Kong

April 2013
Abstract of thesis entitled

Nonlinear Finite Element Analysis of Reinforced


Concrete Beams

Submitted by

YAO Zhong

for the degree of Master of Philosophy


at The University of Hong Kong
in April 2013

A nonlinear finite element program to simulate the behavior of reinforced


concrete (RC) members under the action of monotonic increasing loading has
been developed. The nonlinear response of the RC members is mainly due to the
nonlinear material characteristics including nonlinear biaxial stress-strain relations
and cracking of concrete and yielding of steel reinforcement.

A constitutive model of concrete under biaxial stress state is adopted in


this thesis. In this model, concrete fails and critical cracks occur when the tensile
strain of concrete exceeds the limiting tensile strain. The complete stress-strain
relationship of concrete under compression and tension are employed in the study
to investigate the post-peak behavior of reinforced concrete members. An
elaborate cracking model has been implemented which allows concrete to crack in
one or two directions. The tension stiffening effect of cracked concrete is also
incorporated into this model by including a descending branch in the stress-strain
curve of concrete under tension. Other nonlinear effects such as crushing of
concrete in compression and yielding or strain hardening of steel reinforcement
are also taken into account.

i
A nonlinear finite element program was developed, in which the above-
mentioned nonlinear effects have all been included in modeling the reinforced
concrete structures. The nonlinear equations of equilibrium are solved using an
incremental-iterative technique performed under displacement control. The
validity of the model including the confinement effect of secondary
reinforcements has been examined by analyzing three reinforced concrete beams.
The performance of the numerical model was assessed by comparing results with
those from available experimental data.

(253 words)

ii
DECLARATION

I declare that this thesis entitled “Nonlinear finite element analysis of reinforced
concrete beams” represents my own work, except where due acknowledgement is
made, and that it has not been previously included in a thesis, dissertation or
report submitted to this University or to any other institution for a degree, diploma
or other qualification.

Signed ______________________

YAO Zhong

iii
ACKNOWLEDGEMENT

I would like to take this opportunity to express my sincere gratitude to my


supervisor Prof. Lo Sai Huen for his time and energy devoted to my thesis and
research work.

Special thanks are also due to my dearest friends Mr. L. L. Gang, Mr. F.
Fu, and Ms. M. L. Chen for their valuable support and encouragement during my
depressing time.

The last but not the least, words cannot express the author’s immense
gratitude to his beloved parents. They have provided great support over the whole
period of my study. I gained from them a lot of courage and energy which are
always the sources of my progress.

iv
TABLE OF CONTENTS

ABSTRACT i
DECLARATION iii

ACKNOWLEDGEMENT iv
TABLE OF CONTENTS v
LIST OF FIGURES ix
LIST OF TABLES xi

CHAPTER 1 INTRODUCTION 1

1.1 General 1

1.2 Research Objectives 3

1.3 Organization of the Thesis 3

CHAPTER 2 MODELING OF CONCRETE 5

2.1 Introduction 5

2.2 Mechanical Behavior of Concrete under Different Short Term Static


Loading Conditions 5

2.2.1 Uniaxial loading behavior 7


2.2.1.1 Uniaxial compression 7
2.2.1.2 Uniaxial tension 8
2.2.2 Biaxial loading behavior 10
2.2.3 Triaxial loading behavior 11

2.3 Review of Numerical Models 14

2.3.1 Elasticity based models 14


2.3.1.1 Hyper-elastic model 15
2.3.1.2 Hypo-elastic model 15
2.3.2 Plasticity based models 16
v
2.3.3 Endochronic models 17

2.4 Concrete Model Used in This Study 18

2.4.1 Equivalent uniaxial stress-strain relation of concrete 19


2.4.2 Determination of peak stresses corresponding to the limiting
tensile strain 20
2.4.2.1 Limiting tensile strain failure criterion 20
2.4.2.2 Determination of peak stresses and the corresponding
limiting tensile strain 21
2.4.3 The stress-strain relationship beyond the peak compressive
stress 24
2.4.4 Modeling of concrete cracking 26
2.4.4.1 Crack initiation criteria 26
2.4.4.2 Crack description methods 27
2.4.4.3 Tension stiffening of concrete 31
2.4.4.4 Crack Interface Shear Transfer 35
2.4.5 The computational procedure of this concrete model 38
2.5 Summary 44

CHAPTER 3 MODELING OF REINFORCING STEEL 46

3.1 Introduction 46

3.2 Mechanical Properties of Reinforcing Steel in Static Loading


Conditions 47

3.3 Steel Constitutive Relations 48

3.4 Finite Element Representation of Reinforcing Steel 50

3.5 Summary 51

CHAPTER 4 INTERACTIONS BETWEEN CONCRETE AND


REINFORCEMENT 52

4.1 Introduction 52

4.2 Modeling Bond Action 53


vi
4.2.1 Introduction 53
4.2.2 Modeling of bond action in finite element analysis 54
4.2.3 Bond stress-slip relationships 55
4.2.4 Modeling of bond action based on tension stiffening 57
4.2.4.1 Model of tension stiffening by increasing the stiffness
of reinforcing steel after cracking 59
4.2.4.2 Model tension stiffening effect by retaining the
concrete contributions after cracking 59

4.3 Dowel Action 60

4.4 Modeling of Confinement Effects 63

4.5 Summary 66

CHAPTER 5 NONLINEAR SOLUTION TECHNIQUES 67

5.1 Introduction 67

5.2 Basic Techniques for Solving Nonlinear Equations 68

5.2.1 Basics for iterative methods 69


5.2.2 Iterative methods 70
5.2.2.1 Full Newton-Raphson method 70
5.2.2.2 Modified Newton-Raphson method 71
5.2.2.3 Initial stress method 71

5.3 Displacement Control Technique 73

5.3.1 Introduction 73
5.3.2 Formulation of displacement control method 74

5.4 Convergence Criteria 78

5.5 Termination of the Analysis 79

5.6 Summary 80

CHAPTER 6 APPLICATION TO REINFORCED CONCRETE


BEAMS 81
vii
6.1 Introduction 81

6.2 Classic Beam Specimens 82

6.2.1 Beam OA-1 82


6.2.2 Beam A-1 82
6.2.3 Beam J-4 82

6.3 Analysis of Beams 85

6.4 Parametric Study 91

6.4.1 Comparison study on confinement effect 91


6.4.2 Comparison study on finite element mesh size effect 93
6.4.3 Comparison study on tension stiffening effect and
convergence tolerance 95

6.5 Summary 101

CHAPTER 7 CONCLUSIONS AND RECOMMENDATIONS 103

7.1 Summary 103

7.2 Conclusion 104

7.3 Recommendations for Future Research 105

REFERENCES 107

viii
List of Figures

Figure 2. 1 Stress-strain relationship for aggregate, hardened cement paste, and


concrete 6
Figure 2. 2 Stress-strain behavior of concrete under uniaxial compression 9
Figure 2. 3 Stress-strain behavior of concrete under uniaxial tension 9
Figure 2. 4 Stress-strain relationships of concrete under biaxial compression
(Kupfer 1969) 12
Figure 2. 5 Stress-strain relationships of concrete under tension-compression
(Kupfer 1969) 12
Figure 2. 6 Stress-strain relationships of concrete under biaxial tension (Kupfer
1969) 13
Figure 2. 7 Kupfer’s Strength Envelop (Kupfer 1969) 13
Figure 2. 8 Saenz’s equation for concrete under compression 20
Figure 2. 9 The stress-strain relationship beyond the peak compressive stress 25
Figure 2. 10 Formation of secondary crack in singly cracked concrete 30
Figure 2. 11 Possible crack configurations 32
Figure 2. 12 Representation of tension stiffening 33
Figure 2. 13 Tension softening model of concrete (Guo and Zhang, 1987) 35
Figure 2. 14 Shear transferring by aggregate interlock 37
Figure 2. 15 Comparison of different expressions proposed for the reduced shear
stiffness of concrete 37
Figure 2. 16 Rotation of axis 41
Figure 3.1 Typical stress-strain curves for reinforcing steel (Chen 1982) 48
Figure 3.2 Idealization for stress-strain curves for reinforcing steel 49
Figure 3.3 Idealized stress-strain curve for reinforcing steel used in this study 50
Figure 4.1 Linkage Element to Represent Bond (Ngo and Scordelis 1967) 56
Figure 4.2 Comparison of average bond stress-slip relationships 56
Figure 4.3 Analytical bond stress-slip relationship suggested in CEB FIP 57
Figure 4.4 Stresses distribution in cracked reinforced concrete (Gerstle 1981) 58
Figure 4.5 Dowel action of reinforcing steel crossing a crack 62
Figure 4.6 Internal forces in a cracked beam 63
Figure 4.7 Stress-strain relationship proposed by Kappos (1991) 65

ix
Figure 5.1 Full Newton-Raphson method 72
Figure 5.2 Modified Newton-Raphson method 72
Figure 5.3 Initial stress method 73
Figure 5.4 Typical nonlinear structural response 76
Figure 5.5 Flowchat for displacement control technique 77
Figure 6.1 Details of beam OA-1 (Bresler and Scordelis, 1963) 83
Figure 6.2 Details of beam A-1 (Bresler and Scordelis, 1963) 83
Figure 6.3 Details of beam J-4 (Burns and Siess, 1962) 84
Figure 6.4 Finite element mesh used 86
Figure 6.5 Load-deflection curves of beam OA-1 and A-1 87
Figure 6.6 Crack pattern of beam OA-1 89
Figure 6.7 Crack pattern of beam A-1 90
Figure 6.8 The load-deflection curves of beam OA-1 and beam A-1 92
Figure 6.9 Concrete finite element mesh for beam OA-1 94
Figure 6.10 Comparison of load-deflection curves of beam OA-1 (fine and coarse
mesh) 95
Figure 6.11 Comparison of load-deflection curves of beam J-4 to study
convergence tolerance 99
Figure 6.12 Comparison of load-deflection curves of beam J-4 to study tension
stiffening 100
Figure 6.13 Load-deflection curves of beam J-4 under different conditions 101

x
List of Tables

Table 6.1 Material properties used in application 82

Table 6.2 Ultimate loads and the corresponding deflections of beam OA-1 and
A-1 86

Table 6.3 The ultimate loads of beam J-4 under different combinations 98

xi
CHAPTER 1

INTRODUCTION

1.1 General

Reinforced concrete has been widely accepted as one of the most


important building materials and is used for the construction of many types of
engineering structures. Reinforced concrete is a composite material made up of
concrete and steel. It is important to understand the individual materials properties
forming the reinforced concrete and their composite behavior. But to develop
analytical models of reinforced concrete structures is complicated because the
nonlinearity nature of concrete and the complex interactions between concrete and
steel. With the rapid development of more powerful analytical techniques and
high-speed computers, it is possible for structural engineers to analyze the
behavior of reinforced concrete structures.

In order to analyze the nonlinear behavior of reinforced concrete


successfully, the following aspects should be considered:

1. Realistic material models for concrete and reinforcing steel and their
interactions.
2. Efficient and reliable solution technique (the method used to solve the
nonlinear problem).
3. Reasonable selection of numerical parameters in nonlinear solution
process.

To design or analyze a reinforced concrete structure, sufficient knowledge


about the materials is required. In recent years, significant advances have been
made in the understanding of concrete behavior. However this is still incomplete.

1
Disparities in experimental results are often observed due to difficulties in
obtaining consistent test procedures and test specimens. Furthermore, this is also
due to the natural variability of concrete itself. Concrete is actually a composite
material composed of mortar, aggregates, and voids; therefore its properties vary
widely depending on the mixing conditions, placing and curing. It exhibits highly
nonlinear properties even under relatively low loads due to various factors
including cracking, environmental effects.

On the other hand, the behavior of steel reinforcement is much easier to


model compared with concrete. This is because the behavior of reinforcing steel is
well defined and predominantly uniaxial. However complexities do arise when
combine concrete with reinforcing steel. Bond-slip between steel and concrete,
dowel action under shear deformation, confinement effect can significantly
influence behavior. Extensive research work should be carried out to improve
understanding of these effects.

The finite element method has made possible the inclusion of highly
complex material behavior of reinforced concrete structures into analytical
solutions. Thus the finite element method has become a powerful computational
tool for the reinforced concrete structures. For examples cracking, softening and
crushing of concrete, multiaxial stress response of concrete, yielding of
reinforcement, bond and dowel action between steel and concrete, can be taken
into account by using the finite element method to analyze reinforced concrete
structures. Displacement- based finite element method was adopted in this thesis.

The numerical parameters can significantly influence the nonlinear


solutions. To understand how these parameters influence the solutions and how
they interact is important in analyzing the reinforced concrete structures with
finite element method. It is difficult and expensive to conduct comprehensive
experiments to study the influence of a certain parameter. However, with the help
of finite element method, it is convenient to isolate these specific parameters for
study.

2
The performance of the numerical model was assessed by comparing
results with those from available experimental data. In this thesis, the behaviors of
different reinforced concrete beams have been studied.

1.2 Research Objectives

The specific objectives of this research are:

1. To understand the material models for concrete and reinforcing steel and
the interactions between them.

2. To develop a nonlinear finite element program in Fortran which can


analyze various reinforced concrete beams subjected to short term loading.

3. To use this program to analyze several reinforced concrete beams, and


compare the analytical results with the experimental results, make
conclusions upon the analysis of the results.

4. To understand how numerical parameters influence the solutions and how


they interact.

In summary, the aim of this study is to develop a nonlinear finite element


program to analyze reinforced concrete beams subjected to short term loading,
and undertake parametric studies of reinforced concrete beam behavior.

1.3 Organization of the Thesis

There are seven chapters in this thesis and they are organized as follows:
3
In Chapter 2, a detailed survey of available literature on the experimental
behavior of concrete is presented and a discussion of the material models used for
concrete under biaxial states of stress is given. The modeling of cracking and
tension stiffening effect are also introduced.

In Chapter 3, a model for reinforcing steel is introduced. The analysis of


reinforcing steel is relatively simpler compared with concrete and it can be
modeled linearly. A bilinear stress-strain curve with possible strain hardening is
employed in this research. The finite element representation of the reinforcing
steel is also discussed in this chapter.

In Chapter 4, the modeling of the interactions between reinforcing steel


and the surrounding concrete is discussed. Reinforcing steel can influence the
behavior of concrete through the interactions between them. These interactions
include the bond action, dowel action, and the confinement effect.

In Chapter 5, the numerical solution techniques employed in this thesis are


presented. Details of the iterative methods, the displacement control technique, the
convergence criteria, and the termination of the analysis are given.

In Chapter 6, the theories, finite element program and related numerical


solution techniques have been used to analyze three different reinforced concrete
beams. These analytical results have been compared with experimental results.
And detailed discussion is presented for the parametric study in which the
confinement effect, the finite element mesh size effect, the convergence tolerance,
and the tension stiffening effect are included.

In Chapter 7, conclusions and findings of this study are summarized and


recommendations for future research are suggested.

4
CHAPTER 2

MODELING OF CONCRETE

2.1 Introduction

Finite element analysis of reinforced concrete structures requires realistic


stress-strain relations of the plain concrete as an important input. In general, such
formulations are nonlinear and rely on a number of experimentally determined
material constants. The nonlinear behavior of concrete is very complex because
concrete is a composite made up of hydrated cement, sand, and coarse aggregates.
In addition, concrete contains numerous flaws and micro-cracks. The rapid
propagation of micro-cracks under applied loads is considered responsible for the
nonlinear properties of concrete as well. Various computational models derived
from different researchers and different theories have been proposed to represent
this nonlinear behavior of concrete.

It is the purpose of this chapter to present the concrete material model used
in this research for nonlinear finite element analysis of concrete structures under
general biaxial stress states. First, the mechanical behavior of concrete under
different short term loading conditions is discussed in Section 2.2. A brief review
of concrete models is presented in Section 2.3. Section 2.4 introduces the concrete
material model adopted in this study. Last, a brief summary is given in Section 2.5.

2.2 Mechanical Behavior of Concrete under Different


Short Term Static Loading Conditions

Concrete is a mixture of cement, water, fine aggregates, and coarse


aggregates. When the components of concrete mixed together, the cement and

5
water react to produce a cementing gel which bonds the fine and coarse
aggregates into a rock-like material. As a result, the strength of concrete is a
function of the strength of the cement, the aggregates, and the interaction between
the components.

Typical stress-strain curves of aggregate and cement paste are shown in


Figure 2.1. It is clear that the relationship is linear for aggregates and cement paste
except at a relatively high stress level. However, the combined response of these
materials is markedly nonlinear. This is mainly due to the nature of the cement-
aggregate bond and initiation of micro-cracking when this bond breaks down.

A review of the mechanical behavior of concrete under short term static


loading conditions will be presented in the next section. This will be discussed in
three different states of stress: uniaxial, biaxial and triaxial.

. Figure 2.1 Stress-strain relationship for aggregate, hardened cement paste,


and concrete

6
2.2.1 Uniaxial loading behavior

2.2.1.1 Uniaxial compression

A typical stress-strain curve for plain concrete subjected to uniaxial


compression is presented in Figure 2.2. From the origin to about 30% of ultimate
compressive stress, concrete can be idealized as a linear elastic material (Chen
1982). To some extend micro- and mesocracks (mostly bond cracks) already exist
before any mechanical load is applied to the concrete. These cracks are mostly
due to influences of temperature and moisture gradients within the material after
casting. However, they are quite stable at low stresses and have little tendency to
propagate.

Between about 30% and 50% of the ultimate stress, the cracks begin to
propagate at a very slow rate. It is observed that most of the crack growth
concentrates in the interfacial region between the aggregate and mortar, which is
the weakest link in normal strength concrete (Scrivener and Pratt 1987). At this
stage, the crack propagation is stable because the available internal energy is
almost balanced by the required crack release energy. The stress-strain behavior
of concrete begins to show increasing nonlinear properties.

Once the stress exceeds about 50% of the ultimate stress, the propagation
of cracks is not only remaining in the interfacial region, but also begins to extend
to the cement matrix. The newly generated cracks in the matrix begin to connect
the originally isolated bond cracks and to form a much more continuous crack
system.

At about 70 to 90 percent of peak stress these mesocracks join and start to


form continuous crack patterns, accompanied by a transition from volume
compaction to volume dilatation (Hsu et al. 1963). At higher stress levels instable
crack growth occurs, causing the stress-strain curve to bend sharply towards the
horizontal until it approaches the peak stress.

7
At peak stress, the mesoscopic crack patterns finally grow into
macroscopic cracks which dominate the behavior after peak stress (Kotsovos
1983). A softening post-failure range follows the peak stress and finally ends with
a complete crushing curve.

As a matter of fact, the stress-strain relationship is serious influenced by a


large number of factors. For examples the size and shape of the specimen, test
procedure and the strength of the concrete. A lot of experiments have been
conducted and various uniaxial stress-strain relationships of concrete have been
proposed.

2.2.1.2 Uniaxial tension

The typical shape of tensile stress-strain curve is illustrated in Figure 2.3


which shows some similarities with the compression curve. Stress and strain
increase proportionally up to elastic limit point, i.e., 40 to 60 percent of the
maximum stress. It is much more linear than the compressive curve because the
stable crack propagation stage is much shorter and cracks travel very rapidly
through the mortar matrix and around the interfacial region. Afterwards, strain
increases quickly and the curve becomes more nonlinear. After reaching the
maximum stress (peak or tensile strength), the curve descends sharply.

It is well known that the ratio of tensile strength to compressive strength of


concrete is usually about 0.1 (Hughes 1966). However, to measure the direct
tensile strength of concrete is really a difficult problem, mainly because of the
difficulty of applying a truly concentric pull. According to Raphael (1984), the
splitting tension test is easiest to accomplish and gives the most reliable results.
The splitting tensile strength is calculated with respect to the maximum load
applied to a cylinder and length and diameter of the specimen. On the other hand,
tensile strength can be computed from modulus of rupture test by multiplying by a
factor which takes into account the shape of specimen and the mode of failure. For
example, the factor is ¾ for rectangular cross section.

8
Figure 2.2 Stress-strain behavior of concrete under uniaxial compression

Figure 2.3 Stress-strain behavior of concrete under uniaxial tension

9
2.2.2 Biaxial loading behavior

Numerous experimental investigations into the strength, deformation and


cracking behavior of concrete under biaxial stress states have been conducted
(Kupfer 1969, Nelissen 1972, Slate and Nilson 1972). Unfortunately, the test data
reported by various investigators deviate from each other considerably. It is
believed that the discrepancies between test results from different sources can be
explained by the following factors: types of concrete, shapes of concrete
specimens, testing techniques, etc.

Typical stress-strain curves for concrete under biaxial states of stress in


compression-compression, tension-compression and tension-tension regions are
shown in Figures (2.4), (2.5) and (2.6). These figures were obtained from the
experimental tests of Kupfer et al. (1969). Several conclusions can be drawn from
the figures are as follows:

1. For concrete in compression, the presence of an orthogonal compressive


stress leads to a stiffer response and to greater strength. This is due to the
effect of Poisson’s ratio.

2. For concrete in tension, the presence of an orthogonal compressive stress


reduces the tensile stiffness and strength, as cracking is encouraged.

3. The strength under biaxial tension is almost the same as under uniaxial
tension.

The experiments by Kupfer (1969), Nelissen (1972), the group at Cornell


University, and Su and Hsu (1988) have demonstrated that plain concrete fails by
tensile splitting under uniaxial and biaxial compressive loading with the fractured
surface perpendicular to the direction of the maximum tensile strain. This seems
to indicate that the failure of concrete under uniaxial as well as biaxial stress states
is governed by the tensile property.

10
According to Kupfer et al. (1969), a failure surface of concrete under
biaxial states of stress is presented in Figure 2.7. A few conclusions drawn from
this envelop can be summarized as follows:

1. The ultimate strength increases under biaxial compression, which is


dependent on the ratio of principal stresses. Up to about 25% higher than
the uniaxial value at a stress ratio of about 0.5, and 16% higher than the
uniaxial value when the stress ratio is equal to 1.

2. The compressive stress at failure in the region of combined compression


and tension decreases almost linearly as the tensile stress increases.

3. The biaxial tensile strength of concrete exhibits constant or slightly


increased tensile strength compared with that under uniaxial tension, and
stress-strain curves are similar in shape in both uniaxial and biaxial tension.

In engineering projects, the stress state of many practical structures is a


biaxial one. As a result, biaxial strength envelope, directly derived from
experimental data, have been frequently cited and used.

2.2.3 Triaxial loading behavior

Since the triaxial stress behavior of concrete is not a primary concern of


this research, a brief description will be given for completeness. Triaxial
compressive stress states occur in a large number of concrete structures. However,
there are too many unknowns to describe it in a useful quantitative manner.
Although several researches on the behavior of concrete under triaxial stress state
have been reported (Liu 1985; Richard 1928), a large scatter of test results exists.

11
Figure 2.4 Stress-strain relationships of concrete under biaxial compression
(Kupfer 1969)

Figure 2.5 Stress-strain relationships of concrete under tension-compression


(Kupfer 1969)

12
Figure 2.6 Stress-strain relationships of concrete under biaxial tension (Kupfer
1969)

Figure 2.7 Kupfer’s Strength Envelop (Kupfer 1969)

13
2.3 Review of Numerical Models

In a nonlinear finite element analysis of a reinforced concrete structure, the


numerical modeling of the material properties requires the following: (1) a stress-
strain relationship to simulate the behavior of concrete, (2) a failure criterion,
generally in terms of strains, (3) a cracking criterion, and (4) a model for tension
stiffening.

Several approaches have been proposed by researchers used to define the


stress-strain behavior of concrete. They can be grouped into (1) elasticity based
models, (2) plasticity based models, and (3) endochronic theory based models. An
extensive review of the various models used in the finite element analysis of
reinforced concrete structures is given in an ASCE report (1981).Therefore, only
the basic concepts and the limitations of these models will be reviewed here.

2.3.1 Elasticity based models

The elasticity based models are those models whose constitutive relations
are deduced from the elasticity theory. Early constitutive models were based on
isotropic elasticity to represent concrete behavior (Ngo and Scordelis 1967). The
stress-strain relationship is linear. For Cauchy elastic models, the current state of
stress is uniquely expressed as a function of the current state of strain as

σ = F (ε ) (2.1)

where F is the elastic-response function of material and σ and ε are stress and
strain tensors, respectively.

However, the linear elasticity based model proved to be a good


approximation of the concrete behavior only in the tensile loading environment.

14
Nonlinear elastic models are needed to describe a material exhibits nonlinear
properties like concrete.

In general two different approaches are utilized in the development of


nonlinear elasticity-based stress-strain relationships: (a) finite material
characterization in the form of secant formulations (or termed hyper-elastic
models) and, (b) incremental model in the form of tangential stress-strain relations
(or termed hypo-elastic models).

2.3.1.1 Hyper-elastic model

In this model, an assumption of the existence of a strain energy density


function W is introduced. The stress-strain relationship is expressed as

∂W
σ= = S ε (ε )ε (2.2)
∂ε

The term S ε is the stiffness tensor, which is dependent on the current


strain. Based on a curve fitting of the experimental data, the shape of the function
W can be determined.

The behavior of the hyper-elastic model is reversible and path independent.


It can describe many of the concrete characteristics such as nonlinearity, dilatation
and strain or stress induced anisotropy, but is incapable of describing history
dependence and rate effects.

2.3.1.2 Hypo-elastic model

In this model, the stress-strain relationship is described incrementally,

dσ = S (σ )dε (2.3)
where the variable tangent stiffness, S, is a function of the current state of stress.

15
Various special, simplified forms of the hypo-elastic formulation have
been developed and utilized for modeling concrete behavior. Some of these
models assume the behavior of concrete to be incrementally isotropic, i.e. the
effect of stress-induced anisotropy is neglected (Gerstle 1981). Other models
include a particular type of stress-induced anisotropy, since concrete is
represented as an orthotropic material. One of the most widely used orthotropic
hypo-elastic models is that proposed by Darwin and Pecknold (1977) which has
been used to represent the behavior of concrete under cyclic loading. The model is
based on the concept of "equivalent uniaxial strain", whereby the effect of biaxial
stresses on the internal damage in concrete is represented by equivalent stress-
strain curves for each of the principal stress axes.

The strong points of elasticity based models are their reasonably good
representation of nonlinear concrete behavior and their conceptual simplicity in
comparison with other existing models.

2.3.2 Plasticity based models

It is well known that the classical theory of plasticity can successfully


applied to the analysis of metals. Consequently, extensive modifications should be
made when apply this model to simulate the behavior of concrete.

According to Chen (1982), the incremental theory of plasticity is based on


three fundamental assumptions: (a) the shape of an initial yield surface, (b) the
evolution of subsequent loading surfaces, the hardening rule, and (c) the
formulation of an appropriate flow rule. An initial yield surface in stress space
defines the stress level at which plastic deformation begins. A hardening rule
regulates the evolution of the subsequent loading surfaces during the course of
plastic flow. A flow rule defines an incremental plastic stress-strain relationship
using a plastic potential function.

16
Many plasticity based models have been developed, e.g. Chen and Chen
(1975), Han and Chen (1987) and Chan et al. (1994). These models may
accurately predict inelastic deformation, dilatancy, and hydrostatic pressure
sensitivity, but material softening, and, degradation of stiffness and hysteretic
behavior of concrete under cyclic loading may not be represented satisfactorily.
Since a decrease of all components of stress is impossible for strictly plastic
behavior satisfying Drucker's stability postulate, models which combine plasticity
theories and fracturing concepts have been proposed to account for softening.

2.3.3 Endochronic models

Bazant and his coworkers (1976) have extended the endochronic theory to
concrete structures with great success. The concept of the endochronic theory
consists in characterizing the inelastic strain accumulation by a certain scalar
parameter z, called intrinsic time, whose increment is a function of the
deformation strain increments and the real time increment. The endochronic
model appears to be capable of modeling aspects of the behavior of concrete
material, which were difficult to model by other constitutive models, such as, (a)
the hydrostatic pressure sensitivity of inelastic strain, (b) the inelastic dilatancy
due to shear straining, (c) the strain softening tendency at high stress. By virtue of
the last extension (c), the theory at the same time provides the failure criterion, in
which dependence on strain and stress histories is automatically accounted for.
The presence of real and intrinsic time in the constitutive relations also allows
simultaneous consideration of strain rate effects and nonlinear long-time creep.

Numerical examples gave a very close approximation to experimentally


observed behavior. However, the model was involved a large number of material
parameters, it seems to demand excessive programming and computer efforts.

17
2.4 Concrete Model Used in This Study

A nonlinear elasticity model based on the limiting tensile strain failure


criterion, covering the pre-peak and post-peak regimes, has been utilized for the
representation of the concrete behavior. This model is expressed in terms of secant
stiffness formulation. From the experimental results, Figure 2.7, obtained by
Kupfer et al (1969), it can be seen that concrete behaves as an orthotropic material
in the two principal directions, and the behavior in each direction depends on the
stress in the other direction.

Darvin and Pecknold (1977) have proposed a model of concrete which


represents the behavior of concrete under biaxial states of stress and assumes
concrete to be an orthotropic material. With the reference of the principal axes of
orthotropic, the stress-strain relationship of concrete can be described as:

 
σ1   E1 υ E1 E 2 0  ε 1 
  1 υ E E  ε 
σ 2  = 
E2 0
 2 
(2.4)
σ  1 − υ
2 1 2

 12   0

0
1
4
( )

 
β E1 + E 2 − 2υ E1 E 2  ε 12 

where σ 1 and σ 2 are stresses along two orthotropic axis 1 and 2 respectively; ε 1
and ε 2 are strains along two orthotropic axis 1 and 2 respectively; σ 12 and ε 12
are engineering shear stress and strain respectively; E1 and E 2 are the secant
stiffness, along the two orthotropic axis 1 and 2 respectively; υ is Poisson’s ratio
assumed as a constant. β is the parameter accounting for aggregate interlock,
termed of shear retention factor, will be discussed in detail later. It should be kept
in mind that E1 and E 2 are depending on the state of stress.

It was assumed for the model that stresses in the principal directions could
be calculated independently of each other, based on uniaxial stress-strain

18
relationships. The biaxial effect was assumed to be due to the interaction of the
two principal directions through the Poisson's ratio effect. Therefore, the concept
of “equivalent uniaxial strain” should be introduced in order to separate the
Poisson effect from the cumulative strain. In addition, equivalent uniaxial strain
can made a contribution to keep track of the degradation of stiffness and strength
of plain concrete and to allow actual biaxial stress-strain curves to be derived from
“uniaxial” curves.

2.4.1 Equivalent uniaxial stress-strain relation of concrete

The curves selected for compressive loading are based on an equation


suggested by Saenz (1964) and illustrated in Figure 2.8.

ε iu E 0
σi = (2.5)
E ε ε
1 + ( 0 − 2) iu + ( iu ) 2
ES ε ic ε ic

where E 0 is the modulus of elasticity at zero stress; E S = σ ic / ε ic , is the secant

modulus at the point of maximum compressive stress; σ i is the stress along

principal direction i; ε iu is the equivalent uniaxial strain along principal direction

i; σ ic and ε ic are the peak stress and strain at peak stress respectively. It should be

noted that σ ic and ε ic are functions of the state of stress, have to be determined.

The detailed formulation of σ ic and ε ic will be discussed in Section 2.4.2.

In this model, concrete was assumed to carry compressive stress beyond


the strain corresponding to the concrete ultimate compressive strength, ε c , up to

an ultimate compressive strain ε cu . When concrete reaches its ultimate


compressive strain it was assumed to crush and fail to carry any force.

19
In tension, a complete stress-strain curve of concrete can be treated
similarly to one in compression. Two equations continued at peak are suggested to
describe the ascending and descending branches, respectively (Guo and Zhang
1987). In this investigation, the Saenz’s model, which is used to describe the
concrete behavior under compression, is used to describe the ascending branch of
the stress-strain curve under tension. The post-crack behavior of concrete under
tension will be discussed in detail later.

Figure 2.8 Saenz’s equation for concrete under compression

2.4.2 Determination of peak stresses corresponding to the limiting


tensile strain

2.4.2.1 Limiting tensile strain failure criterion

Concrete failure is due to the initiation of critical cracks. Critical cracks in


concrete may occur when the maximum tensile strain of concrete exceeds a
critical value. This critical value is termed the limiting tensile strain. In summary,
the limiting tensile strain criterion defines the concrete fails when the maximum
tensile strain exceeds the limiting tensile strain. This criterion is adopted in this
thesis.
20
The evaluation of limiting tensile strain of concrete under different stress
conditions will be discussed together with the peak stress in the next section.

2.4.2.2 Determination of peak stresses and the corresponding limiting tensile


strain

The value of maximum compressive stress, σ ic , for different combinations


of biaxial stresses has been the objective of many experimental studies. Kupfer
and Gerstle (1973), Liu, Nilson and Slate (1972), and Nelissen (1972) have
conducted biaxial loading tests of concrete; and the maximum strength criteria
that they obtained were quite consistent between each other.

Kupfer and Gerstle (1973) have suggested a strength envelop to determine


the peak stresses of concrete ( σ ic ) under biaxial state of stress. This criterion is
adopted in this study.

The peak stresses in the principal directions σ ic and the corresponding

strains ε ic are functions of the biaxial principal stress ratio, α . Darvin and

Pecknold (1977) gave expressions for the peak stresses in terms of α . For the
corresponding strains ε ic , expressions were obtained from the limiting tensile
strain criterion.

The biaxial strength envelope is divided into three regions, which depends
on the state of stresses, as represented by the biaxial principal stress ratio α . The
expressions for the peak stresses of the three regions of the strength envelope (T-T
zone, T-C zone, and C-C zone) are summarized as follows:

(i) Compression-compression region

For biaxial compression, Kupfer and Gerstle found that the strength
envelope was closely approximated by the equation
σ1 σ2 σ2 σ1
( + )2 − − 3.65 =0 (2.6)
fc fc fc fc

21
Equation (2.6) can be rewritten to give the maximum compressive strength
of the concrete σ 2c , in the form of α and f c .

1 + 3.65α
σ 2c = fc (2.7)
(1 + α ) 2

σ 1c = α σ 2c (2.8)

in which f c is the uniaxial compressive strength of the concrete.

The values of σ 1c and σ 2c are used to define the shape of the equivalent

uniaxial stress-strain curves for a given value of α . However, it should be kept in


mind that the shapes of the curves change continuously due to the ration of σ 1 to
σ 2 changes.

According to He (1999), the limiting tensile strain in this situation can be


summarized as following:

ε 2c = (1.55 − 0.46α )ε 0c when 0.2< α ≤ 1.0 (2.9)

ε 2c = (2.26α + 1)ε 0c when 0 < α ≤ 0.2 (2.10)

where ε 0 c is the strain corresponding to the peak stress at the stress-strain curve
under uniaxial compression.

(ii) Tension-compression region

For tension-compression, Kupfer and Gerstle suggest a straight line


reduction in tensile strength with increased compressive stress.

22
σ2
σ 1t = (1 - 0.8 ) ft (2.11)
fc

in which f t is the uniaxial tensile strength of the concrete.

Equation (2.11) can be rewritten in a more direct way as:

αf c f t
σ 1t = (2.12)
αf − 0.8 f t

This equation worked quite satisfactorily in material simulation of this


situation.

Equation (2.7) is slightly modified for negative ratios of σ 1 to σ 2 . For


tension-compression, the compressive stress-strain curves are modeled with the
following equation:

1 + 3.28α
σ 2c = fc (2.13)
(1 + α ) 2

In this situation, the limiting tensile strain can be defined as

 1+ α 
ε 1c =  A + B  × 10 −3 (2.14)
 0 .8 − m α 

where
0.5136 m
A= (2.15)
0 .8 + m
0.6336 + 0.15m
B= (2.16)
0 .8 + m
− fc
m= (2.17)
ft

23
(iii) Tension-Tension region

For tension-tension, Kupfer and Gerstle recommended a constant tensile


strength, equal to uniaxial tensile strength of the material. This is in close
agreement with the experimental data of many investigators. The expression for
the ultimate tensile strength for concrete under biaxial tension is defined as
follows:

σ 1t = f t (2.18)

σ 2t = σ 1t / α (2.19)

The limiting tensile strain under tension-tension biaxial state of stress is a


constant, which equals to the one under uniaxial tension. According to CEB-FIP
(1993), the magnitude of 0.00015 is adopted.

The equivalent stress-strain relationship describe by Equation (2.5) now


can be determined by using the peak stress and the limiting tensile strain defined
in this section. Furthermore, the secant modulus of elasticity E which defines the
nonlinear properties of concrete can be calculated.

2.4.3 The stress-strain relationship beyond the peak compressive


stress

The existence of the descending branch of the stress-strain curve beyond


the peak compressive stress has been long established. That means concrete
structure would not lose its loading-carrying capacity immediately until the out-
of-plane cracks propagate to a certain extent. Therefore, a falling branch of the
stress-strain relationship beyond the peak compressive stress should be
incorporate into this model.

24
Several mathematical models have been proposed for the descending curve
of the uniaxial stress-strain relationship of concrete in compression by different
researchers (Hognestad 1951; Saenz 1964; CEB-FIP 1993). A simple approach to
include the falling branch was adopted in the present work. Beyond the peak
compressive stress, the stress-strain relationship is assumed linear, proposed by
Kent and Park (1971), as shown in Figure 2.9.

In this model, Kent and Park suggest the slope of the descending part, u ,
can be calculated as follows,

0.50
u= (2.20)
3 + 0.29 | f c |
− ε 0c
145 | f c | −1000

A strain criterion is used in the present work to define the crushing of


concrete under biaxial compression. Crushing of concrete is assumed to occur
when the equivalent strain ε ic reaches the ultimate strain value ε cu , which defines
the maximum usable strain in calculation and it usually taken in the range 0.003 to
0.0035. Concrete can carry zero stress when it crushed.

Figure 2.9 The stress-strain relationship beyond the peak compressive stress

25
2.4.4 Modeling of concrete cracking

The phenomenon of concrete cracking is extremely important in the load-


deflection behavior of reinforced concrete structures. During monotonic loading
of a structure continued up to failure, cracking occurs at an early stage of loading.
Previous investigations have shown that treatment of post-cracking behavior plays
an important role in predicting the ultimate load and even failure pattern of
reinforced concrete structures. Consequently, an enormous amount of attention
has been given to the numerical representation of cracks. The following four
aspects should be considered:

(i) a criteria for crack initiation,


(ii) a method for crack representation,
(iii) a description of tension stiffening,
(iv) a model for crack interface shear transfer.

2.4.4.1 Crack initiation criteria

The maximum stress and strain theories are frequently used to determine
whether tensile cracking has occurred in the concrete (Chen 1982). If the
maximum principal stress or strain in a point of the structure reaches the uniaxial
tensile strength or tensile strain limit, cracking is assumed to form perpendicular
to the direction of the maximum tensile stress or strain. The stress in that direction
is subsequently reduced to zero.

The limiting tensile stress value which causes the crack is not a well-
defined quantity. For specimens cast from the same concrete, the flexural tensile
strength determined from a modulus of rupture test is higher than the tensile
strength of a split cylinder, which is in turn higher than the tensile strength
obtained from a direct tension test. Furthermore, for each type of test there is

26
significant scatter in the results. For concrete structures subjected to rapid loading,
the maximum strain criterion is more realistic, since uniaxial dynamic tensile tests
(Hatano 1960) indicate that an almost constant failure strain is observed
irrespective of the strain rate or loading rate. The limiting tensile strain criterion
has been employed with success to represent the tensile cracking of concrete
under static (Chen and Chang 1980) loading.

The detailed explanation of limiting tensile strain criterion has been given
in previous section.

2.4.4.2 Crack description methods

Cracks are represented in the finite element analysis in two different ways:
the discrete and smeared crack representations. The choice between these two
models depends on the purpose of the study. If bond slip, dowel shear effects and
aggregate interlocking across cracks, giving realistic crack patterns are of main
interest of the research, the discrete approach is convenient, since it can account
better for these. However, in cases where the overall behavior of a structure is
desired or where many cracks have formed, the smeared crack concept is
preferable.

2.4.4.2.1 Discrete crack model

The earliest application of the finite element method to cracked reinforced


concrete structures was made by Ngo and Scordelis (1967). In their research,
cracks were regarded as discrete gaps between the boundaries of elements, which
were predefined in the finite element mesh. As the loading increases, the nodal
“splitting” in the direction of crack propagation leads to a continuous change in
the topology of structure, which is very complex and time consuming. This model
was later improved by Nilson (1968) to include flexibility in specifying crack
directions and linkage elements along the separated nodes to simulate aggregate

27
interlock and dowel action. The discrete crack approach was further improved and
partially automated by Al Mahaidi (1979) who used a predefined crack utilizing
two nodes at one point connected by a linkage element.

Despite all these improvements, this approach has the disadvantage that it
involves changes in the topology of the finite element mesh following the
formation of a crack and the lack of generality in possible crack direction. Due to
these difficulties and restrictions the discrete crack models are not favored in
finite element analysis for general structural application.

2.4.4.2.2 Smeared crack model

The smeared crack approach which was proposed by Rashid (1968), offers
a more practical alternative for crack representation while using the finite element
method.

Instead of redefining the finite element mesh and nodal connectivities,


most previous works (Rashid 1968; Zienkiewicz 1974) took cracking into account
by modifying material properties within elements. After a first crack has occurred,
the concrete becomes an orthotropic material with one of the material axes being
oriented along the direction of cracking. The elasticity modulus in the direction
perpendicular to the cracking plane is reduced to zero. A reduced shear modulus
β G is assumed on the cracked plane to account for aggregate interlocking. The
value of β will be discussed later in this chapter. The tension stiffening effect can
be coarsely simulated by a gradual release of tensile stress after cracking.

This representation is more popular since it allows for automatic


generation of cracks without redefinition of the finite element topology and offers
complete generality in possible crack direction.

However, the smeared crack models also have appeared some arguments.
For example, smeared cracking representation is very sensitive to mesh
discretization, and also to the order of integration rule adopted, since the element

28
stiffness is determined from the contributions associated with each cracked or
uncracked integration point.

(i) Two crack models based on smeared crack approach

After a crack has formed its direction can be taken as fixed or can be free
to rotate. In the fixed crack approach the crack direction is defined by the
orientation of the initial crack and is then held fixed regardless of the change in
the principal directions with the changing loading.

The rotating crack approach was originally proposed by Cope et al. (1980).
This model is based on the assumption that the crack direction is always normal to
the direction of the maximum tensile principal stress. Milford and Schnobrich
(1985) stated that the cracks defined by the rotating crack model are not cracks in
the strict sense, but rather notational cracks defining the average crack rotation.

(ii) doubly-cracked concrete

In the present work, concrete is allowed to crack in two directions. The


following possibilities are allowed:

(a) In previously uncracked concrete, a smeared crack develops


perpendicular to the direction of principal stress if the tensile strain along this
principal direction exceeds the limiting tensile strain. Furthermore, smear cracks
in two orthogonal directions develop when both principal strain are tensile and
exceed the limiting tensile strain of concrete.

(b) On further loading of singly cracked concrete, another set of smeared


cracks may form when the strain state produces a tensile principal strain which
exceeds the limiting tensile strain of concrete. In this case, the secondary set of
cracks can occur as illustrated in Figure 2.10.

29
Figure 2.10 Formation of secondary crack in singly cracked concrete

30
(iii) Closing of cracks

In order to improve the reality of the present cracking model, the


possibility of crack closing is considered. This behavior may take place due to the
redistribution of stresses upon further loading. The fictitious strain normal to the
crack direction is used to assess the state of the cracks in the cracked concrete. If
this strain has a negative value, then the crack is assumed to be fully closed and
the modulus of elasticity E is retained in the direction normal to the crack.

In conclusion, concrete can crack along one or two directions, and cracks
are allowed to open or close in the present research. Figure 2.11 shows different
possible crack configurations in this research.

2.4.4.3 Tension stiffening of concrete

Concrete between the cracks is still capable of transferring some tensile


stress mainly due to bond effects, which is well known as the tension stiffening
effect of cracked concrete. It has been a common assumption in the early
numerical analysis studies that concrete is a purely brittle material in tension (Ngo
and Scordelis 1967). When cracking occurs, the stress normal to the crack
direction is immediately reduced to zero. However, it was soon discovered that
this procedure would lead to convergence difficulties, and more importantly, the
numerical results are inconsistent with the experimental data. Thus a descending
branch should be included in the stress-strain curve of concrete under tension.

Many different shaped curves have been proposed. Scanlon (1971) used a
stepped piecewise linear unloading relation. A similar approach was used by Lin
and Scordelis (1975), with a smooth unloading curve. Figure 2.12(a) and (b)
shows these two alternatives stress-strain relations for concrete in tension.

31
Figure 2.11 Possible crack configurations

32
(a) Scanlon’s stepped unloading curve (Scanlon 1971)

(b) Lin’s gradual unloading curve (Lin and Scordelis 1975)

(c) Modified stress-strain diagram for reinforcing steel (Gilbert and Warner 1978)

Figure 2.12 Representation of tension stiffening

33
Another approach to modeling tension stiffening effect is to ignore the
concrete after cracking and increase the reinforcing steel stiffness fictitiously
(Gilbert and Warner 1978), as shown in Figure 2.12 (c). The concrete is assumed
to carry no stress normal to a crack, but an additional stress will be added to the
reinforcing steel. This approach implies that the tension stiffening effect is
concentrated at the reinforcing steel level, whereas experimental observations
showed that this effect is distributed throughout the depth of the tension zone.

The inclusion of tension stiffening effects does not only simulate the real
behavior of concrete, but it also improves the stability of the numerical analysis.
However, there are still some arguments with this tension stiffening effect. For
example, this effect, if explained in terms of bond interaction with the reinforcing
steel, cannot be applied to plain concrete structures, or to concrete located at a
certain distance from the reinforcement. But it is well known that plain concrete is
not a perfectly brittle material that it has some residual load carrying capacity after
reaching the tensile strength. Therefore, the concept of “tension softening” is
introduced. This can be accomplished by adding a descending branch to the stress-
strain curve.

In this study, a relatively simple post-peak stress-strain relationship


proposed by Guo and Zhuang (1987) for uniaxial tension is adopted to estimate
the tension softening effect:

ε
ft
εt
σ= (2.21)
ε ε
0.312 f t ( − 1)1.7 +
2

εt εt

where σ and ε is the stress and strain in concrete respectively; f t is the tensile

strength of concrete and ε t is the strain of concrete at tensile strength. And this
model is shown in Fig. 2.13.

34
To summarize, the post-peak stress-strain relationship of cracked concrete
is dependent on the tension softening behavior of plain concrete and the bond
force transferred from the reinforcement. Therefore, in this research, a
modification of the model of tension softening effect is made to take the tension
stiffening effect into account. This will be discussed in detail in Chapter 4.

Figure 2.13 Tension softening model of concrete (Guo and Zhang, 1987)

2.4.4.4 Crack Interface Shear Transfer

Experimental results show that cracked concrete is still capable of


transmitting shear stresses due to aggregate interlock and dowel action.

When the crack is formed and subjected to shear forces, the crack
interfaces tend to override each other as a result of the shear slip. This tendency
may be constrained either by surrounding a portion of concrete or by the
reinforcing bars crossing the cracks. It has been shown experimentally that the
contact area between aggregate particles and cement paste plays a prominent role
in the shear transfer mechanism of aggregate interlock, as shown in Figure 2.14.

35
That is to say, increasing crack width will lead to reduction of the contact area,
which results in smaller shear stiffness.

Shear transmission due to aggregate interlock can be simply accounted for


in the smeared crack model by the introduction of a reduced value of concrete
shear modulus, β G. Hand et al. (1973) used a constant value for β throughout
the analysis. Later, Cedolin and Dei Poli (1977), in an analysis of beams falling in
shear, adopted a value of β linearly decreasing with the fictitious strain normal to
the crack (which represents the crack width). Al-Mahaidi (1979) has suggested a
hyperbolic variation of β with the fictitious strain normal to the crack, see Figure
2.15.

A reduced shear modulus which depends on the crack width seems more
realistic and was adopted in this research. A simple parabolic expression to
calculate the shear retention factor β , is adopted as the follows (He 1999):

2
 ε - ε 1c 
β = 0.2 1 -  ≤ 0.2 (2.22)
 ε 0c 

in which ε 1c is the limiting tensile strain. The quantity ( ε - ε 1c ) can be regarded as

a measure of the crack width. β decreases as the crack width increases and is
always larger than 0.

When secondary cracking occurs a further reduction is imposed on the


shear stiffness. This is achieved by multiplying the value of β , for the first crack,
by a second value calculated from Equation (2.22) using the strain normal to the
secondary crack.

36
Figure 2.14 Shear transferring by aggregate interlock

Figure 2.15 Comparison of different expressions proposed for the reduced shear
stiffness of concrete

37
2.4.5 The computational procedure of this concrete model

1. Enter with the values of strains {ε }n -1 , stresses {σ }n -1 , and initial crack


conditions. At the beginning of the analysis {ε }n -1 and {σ }n -1 are zero and the
material is uncracked.

2. Apply a load increment of ∆{P}n from which the incremental strains ∆{ε }n
and stresses ∆{σ }n are calculated.

3. Calculate the total strains and total stresses from:

{ε }n = {ε }n -1 + ∆{ε }n (2.23)
{σ }n = {σ }n -1 + ∆{σ }n (2.24)

4. If the concrete is initially uncracked then:

(a) Calculate the principal angle α c from the equation:

1 2τ xy
αc = tan −1 ( ) (2.25)
2 σ xx − σ yy

This angle is shown in Figure 2.16, and it varies from -45° to +45°.

(b) Calculate the total principal stresses {σ p }n and total principal strain {ε p }n

using the following equation:

{ε p }n = [T ] {ε }n (2.26)

[ ]
{σ p }n = T −1 {σ }n
T
(2.27)

38
in which [T ] is a transformation matrix defines by:

 cos 2 α c sin 2 α c cos α c sin α c 


[T ] =  sin 2 α c cos α c
2
− cos α c sin α c 

(2.28)
− 2 cos α c sin α c 2 cos α c sin α c cos 2 α c − sin 2 α c 

The matrix [T ] is used to calculate the principal stresses because it allows


the stresses to be orientated in the proper direction with respect to the angle α c .

(c) If {ε p }n ≥ ε 1c , a crack occurs perpendicular to the direction of the maximum

principal strain. Fix angle α c as the permanent angle of crack; and calculate the
new stresses in the direction perpendicular to the crack on the principal axis using
the tension stiffening equation, which will be introduced in Chapter 4. As a result,
ε i and σ i represents the strain and stress in the direction perpendicular to the
crack can be obtained.

(d) Calculate the stress in the direction parallel to the crack σ j and the assumed

retained shear stress τ from:

σ j = E jε j (2.29)
and
τ = β Gγ (2.30)

where j refers to the direction parallel to the crack and γ is the calculated shear
strain value. The updated stresses can then be written as:

σi 
 
{σ c }n =  σ j  (2.31)
τ 
 

if the crack occurred perpendicular to the axis 1 in Figure 2.16.

39
and

σ j 
 
{σ c }n =  σ i  (2.32)
τ 
 

if the crack occurred perpendicular to the axis 2 in Figure 2.16.

(e) The new Cartesian stresses can then be calculated from:

{σ }n = [T ] {σ c }n
T
(2.33)

(f) Steps (c) to (e) are also used to check whether two cracks have occurred at the
same load level and if so stresses will be updated accordingly in both directions.

(g) For the stiffness calculation in subsequent iterations [Dc ] is set equal to:

0 0 0 
[Dc ] = 0 E2 0  (2.34)
0 0 βG 

when a crack occurs perpendicular to axis (1) in Figure 2.16, or is set equal to:

 E1 0 0 
[Dc ] =  0 0 0  (2.35)
 0 0 β G 

if a crack occurs perpendicular to axis (2) in Figure 2.16.

If {ε p }n ≥ ε 1c in both directions then two cracks occur at the same time

perpendicular to each other. The angle α c is fixed as before and the new stiffness
matrix is updated to:

40
0 0 0 
[Dc ] = 0 0 0  (2.36)
0 0 β G 

E1 and E2 are the secant modulus of elasticity of concrete in the direction parallel
to the single crack, and β is a shear factor defining the shear resistance parallel to
the crack due to aggregate interlocking. The value of β can be determined by
Equation (2.22).

Figure 2.16 Rotation of axis

41
5. If cracks already exist then:

(a) Use the fixed angle of crack α c and transform the Cartesian strains and
stresses into the crack direction:

{ε c }n = [T ] {ε }n (2.37)

[ ]
{σ c }n = T −1 {σ }n
T
(2.38)

(b) If one crack exists then check whether the strain perpendicular to the crack is
changed from the descending to the ascending part of the stress-strain curve, then
update the stress on the principal angle from

{σ c }n = [Dc ] {ε c }n (2.39)

where

 E1 0 0 
[Dc ] =  0 E2 0  (2.40)
 0 0 βG 

and calculate the new Cartesian stresses from:

{σ }n = [T ] {σ c }n
T
(2.41)

For stiffness calculations in subsequent iterations [Dc ] is set equal to:

0 0 0 
[Dc ] = 0 E2 0  (2.42)
0 0 βG 

42
when the change of strain occurs perpendicular to axis (1) in Figure 2.16, or is set
equal to:

 E1 0 0 
[Dc ] =  0 0 0  (2.43)
 0 0 β G 

when the change of strain occurs perpendicular to axis (2) in Figure 2.16.

This criterion is used instead of using a criterion for crack closing, and
when the strains become negative then the secant modulus for the compression
zone is used.

(c) To check whether a crack exists in second direction the same procedure for
checking a crack in the first direction is used except that [Dc ] for the stiffness
calculation is set up so that:

0 0 0 
[Dc ] = 0 0 0  (2.44)
0 0 β G 

If a second crack already exists, then check if the strain in any single
direction or both directions changes from the descending to the ascending part of
the curve. Then calculate the stresses using:

{σ c }n = [Dc ] {ε c }n (2.45)

where
0 0 0 
[Dc ] = 0 E2 0  (2.46)
0 0 βG 

if the strain changes on axis (1) in Figure 2.16, or is set equal to:

43
 E1 0 0 
[Dc ] =  0 0 0  (2.47)
 0 0 β G 

if the strain changes on axis (2) in Figure 2.16.

If the strains change in both directions set to

 E1 0 0 
[Dc ] =  0 E2 0  (2.48)
 0 0 βG 

The new Cartesian stresses are then updated by using:

{σ }n = [T ] {σ c }n
T
(2.49)

For stiffness calculation [Dc ] remains the same as in Equation (2.44).

2.5 Summary

First, the mechanical behavior of concrete under different loading


conditions and a brief review of concrete models are presented in this chapter.
Then a detailed discussion of the concrete model adopted in this study is given.
The following points can be summarized in this model:

1. This is an orthotropic model based on the concept of equivalent uniaxial


strain.

2. The limiting tensile strain failure criterion defines concrete fails when the
tensile strain of concrete exceeds the limiting tensile strain. Cracking occur
perpendicular to the direction of the maximum tensile strain.

44
3. Doubly-crack model based on smear crack representation is incorporated
in this research. Concrete can crack along one or two directions. Moreover,
cracks are allowed to open or close.

4. The tension stiffening and shear retention are taken into consideration in
this model.

45
CHAPTER 3

MODELING OF REINFORCING STEEL

3.1 Introduction

In reinforced concrete structures, the behavior of reinforcing steel can


severely influence the behavior of the total structure. And in contrast with
concrete, the mechanical properties of steel reinforcement are well known. The
reinforcing steel bars are assumed to be capable of transmitting axial compressive
or tensile forces only. Thus, a uniaxial stress-strain relationship is sufficient for
monotonic or cyclic loading. Typical stress-strain curves for reinforcing steel bars,
loaded monotonically in tension, with different yield strengths are shown in
Figure 3.1. The stress-strain curves of steel are assumed to be identical in tension
and compression. Therefore, the same stress-strain relationship is adopted in this
research for reinforcing steel bars subjected to tension or compression.

A brief review of the uniaxial mechanical properties for reinforcing steel is


given in Section 3.2, while more details can be found in reviews published in
Mainstone (1975). Constitutive relationships for reinforcing steel used in the
previous investigations as well as the model used in this study are presented in
Section 3.3. Section 3.4 introduces the finite element representation of reinforcing
steel used in this study. A brief summary is given in Section 3.5.

46
3.2 Mechanical Properties of Reinforcing Steel in Static
Loading Conditions

Typical stress-strain curves for steel reinforcing bars under monotonic


static tensile loading are plotted in Figure 3.1. These curves are normally
characterized by the following features:

1. An initial elastic region up to the yield strain ε y of reinforcing steel exists,

with mean modulus of elasticity of 200 GPa.

2. A yield plateau followed by a strain hardening region from ε y to the

ultimate strain ε u , to the fracture strain ε f .

3. The ultimate strength is approximately 1.55 times the yield strength.

4. As the strength of the reinforcement is increased, its capacity to undergo


inelastic deformation, or its ductility, is decreased.

Under monotonic loading, it is generally assumed that the steel behavior is


identical in tension and compression. Unloading and reloading result in a response
path approximately parallel to the original elastic shape.

47
Figure 3.1 Typical stress-strain curves for reinforcing steel (Chen 1982)

3.3 Steel Constitutive Relations

Several constitutive theories have been proposed for steel in previous


studies. Indeed, most of the reviewed constitutive theories of concrete described in
Section 2.3 were primarily developed for steel or other metals. However, the most
common approach found in the literature for the material modeling of reinforcing
steel is the elastic-strain hardening constitutive model.

For simplicity in the analysis, it is often necessary to idealize the steel


stress-strain curve. Figure 3.2 shows 4 alternative idealizations for reinforcing
steel under monotonic loading. Figure 3.2(a) regards reinforcing steel as elastic-
perfectly plastic material. Bilinear (Figure 3.2b) or trilinear (Figure 3.2c)
idealization has been commonly used depending on the accuracy required. Figure
3.2 (d) shows a representation of a complete stress-strain curve.

The uniaxial stress-strain curve of steel idealized in the present model is


shown in Figure 3.3. A bilinear stress-strain curve with possible strain hardening
48
is employed in this research and elastic unloading is allowed. Following
properties are required to construct the stress-strain curve: E s is the initial

modulus of elasticity of reinforcing steel, E sh is the modulus of strain hardening,

σ y is the yield stress, and ε u is the ultimate strain.

If the modulus of the strain hardening, E sh , was set equal to zero, the
stress-strain relationship became an elastic-plastic one. Loading in tension and
compression was assumed to be elastic until the stress reaches the yield stress, σ y ,

beyond that yielding was assumed to have occurred. When the strain of the
reinforcing steel reached the ultimate strain ε u , failure was assumed to occur.
Then the reinforcing steel was assumed to rupture and carry zero stress.

Figure 3.2 Idealization for stress-strain curves for reinforcing steel

49
Figure 3.3 Idealized stress-strain curve for reinforcing steel used in this study

3.4 Finite Element Representation of Reinforcing Steel

Three alternative representations of the reinforcement, i.e., distributed,


embedded, and discrete representations (ASCE 1982), can be used to develop a
finite element model for the reinforcing steel.

Smeared model is adopted in this thesis. In this model, the reinforcement


is considered as a component of the composite material, instead of considering the
concrete and reinforcement individually. A composite concrete reinforcement
constitutive relation can be achieved as follows:

[D] = [DC ] + [DS ] (3.1)

Where [DC ] = material stiffness matrix for concrete,

[DS ] = material stiffness matrix for reinforcement,


50
and
ρx 0 0
[DS ] = ES  0 ρy 0
 0 0 0
(3.2)

in which E S = modulus of elasticity of the reinforcement,

ρ x , ρ y = reinforcement ratio of the element in x- and y-directions


respectively.

3.5 Summary
In this chapter, a brief introduction of the mechanical properties of
reinforcing steel is given in the first place. Next in this chapter, the stress-strain
relation and the finite element representation of reinforcing steel adopted in this
thesis are presented.

51
CHAPTER 4

INTERACTIONS BETWEEN CONCRETE AND


REINFORCEMENT

4.1 Introduction

The behavior of plain concrete and reinforcing steel bars have been
introduced and modeled separately in the previous chapters. However, reinforced
concrete structures are the structures that composed of these two materials, and
exhibit tremendous different properties due to the complex interactions between
the reinforcing steel and the surrounding concrete. These interactions can be
summarized as three aspects:

1. Bond action between reinforcing steel and surrounding concrete which is


considered to have the ability to influence the stiffness of the structure
tremendously.

2. Dowel action of reinforcement crossing cracks which influences the post-


peak behavior of reinforced concrete members significantly.

3. Confinement effects when transverse reinforcement exists which can


increase the load-carrying capacity and ductility of the reinforced concrete
members.

In this chapter, these three interactions between reinforcing steel and the
surrounding concrete will be discussed in detail.

52
4.2 Modeling Bond Action

4.2.1 Introduction

The bond action is one of the most important interactions between


concrete and reinforcing steel. The bond action between concrete and reinforcing
steel bars exists because of the chemical adhesion, friction, and mechanical
interlock between concrete and steel. Bond in plain bars depends mainly on
adhesion and friction with some mechanical interlock due to roughness of bar
surface while deformed bars depends mainly on mechanical interlock which gives
deformed bars superior bond strength.

Since bond is always related to slip, researchers have been engaged in


studying the bond-slip relationships. One of the most common methods used to
represent bond-slip is the use of linkage elements. The main idea is to include a
linkage element which connects the nodes of steel and concrete finite elements.
This linkage element is composed of two springs perpendicular to each other
which allow the transfer of forces between the separate nodes. Each spring must
be given a certain stiffness, obtained from experimental tests. Normally a
nonlinear bond-slip relationship, obtained from a pullout test, is assigned to the
spring parallel to the bar, and a very high stiffness is given to the spring
perpendicular to the bar to prevent any separation between the bar and the
concrete.

Another way to account for bond action is to use a gradual stiffening curve
for concrete after cracking has occurred, and is usually termed of "tension
stiffening". This approach is based on the fact that if the opening of a crack occurs
at the same time as bond failure, it will cause some movement between the bar
and the concrete. This will then cause the shear force at the contact surface
between the cracks to feed tension stresses into the concrete. The concrete,
attached to the bar will contribute to the overall stiffness of the system. This

53
contribution is accounted for by the gradual stiffening curve described in this
chapter. On the other hand, an alternative way of representing this stiffening effect
is to increase the steel stiffness and stress.

In conjunction with the smeared models of cracked concrete and


reinforcement, linkage elements are not used and the slip between reinforcement
and the surrounding concrete is ignore in this thesis. Instead, using a tension
stiffening stress-strain curve incorporated into this model to represent the bond
action between concrete and reinforcing steel.

4.2.2 Modeling of bond action in finite element analysis

Ngo and Scordelis (1967) proposed the earliest work on modeling of bond
in finite element analysis of reinforced concrete. They introduced a linkage
element between concrete and steel to represent bond. According to Ngo and
Scordelis, the linkage element can be thought of conceptually as consisting of two
linear springs parallel to a set of orthogonal axes H and V, as shown in Figure 4.1.
Each of the springs is assigned a stiffness value from which the stiffness matrix
for the linkage element is obtained. If the springs in the H and V directions have
stiffness k h and k v respectively then the stress strain relation is given by:

σ h  k h 0  ε h 
σ  =  0 k v  ε v 
(4.1)
 v 

in which ε h and ε v are the relative displacements between points I and J in the H
and V directions.

Nilson (1968) pointed out that the relationship between bond stress and
bond slip is strictly nonlinear. He introduced a bond-slip equation which is
derived indirectly from experiments reported by Bresler and Bertero (1966). A
third order degree polynomial relating local bond stress µ to local bond slip d is
given as:

54
µ = 3606 × 10 3 d − 5356 × 10 6 d 2 + 1986 × 10 9 d 3 (4.2)

The spring linkage stiffness is found by differentiating µ with respect to


the displacement d .

Many researchers have utilized spring linkage elements to represent bond


in finite element analysis of reinforced concrete. For example, Robins (1971) used
spring linkage elements to simulate bond in the analysis of a reinforced concrete
deep beam by finite element method.

4.2.3 Bond stress-slip relationships

Bond stress can be thought as the shearing stress between reinforcing bar
and the surrounding concrete. Bond stress-slip relationship is strictly nonlinear
and it varies with the position along the bar. Experimental results of bond stress-
slip relationship are presented based on average or local values. To include such a
relation in analysis by finite element method normally an average bond stress-slip
relationship is used.

Different relations are obtained by different investigators. The difference


between these results is illustrated in Figure 4.2 which shows the average bond
stress-slip relationships obtained by a number of investigators.

CEB FIP Model Code (2000) suggests an analytical bond stress-slip model
under monotonic loading, which is largely based on the model developed by
Eligehausen et al. (1983), as shown in Figure 4.3. In Figure 4.3, τ mac is bond

strength, s1 , s 2 , s3 are the characteristic slip value related to bond slip behavior.
The shape of this curve is determined by the main influencing factors for bond-
slip behavior: roughness of the bar surface,confinement, bond condition, and
concrete strength.

55
Figure 4.1 Linkage Element to Represent Bond (Ngo and Scordelis 1967)

Figure 4.2 Comparison of average bond stress-slip relationships

56
Figure 4.3 Analytical bond stress-slip relationship suggested in CEB FIP

4.2.4 Modeling of bond action based on tension stiffening

The approach to model the bond action by considering slip between


reinforcing steel and the surrounding concrete is not adopted in this thesis,
because discrete crack model and discrete representation of reinforcement should
be incorporated. Moreover, the global load-deflection response of the reinforced
concrete structures, instead of the local behavior, is the main concern of this study,
perfect bond between the reinforcing steel and the concrete is assumed. However,
the bond forces transferred between reinforcing steel bars and the surrounding
concrete influence the total load-deflection behavior of reinforced concrete
structures severely and can not be simply neglected. Therefore, modeling the bond
action based on tension stiffening effect is introduced in this research.

57
Figure 4.4 shows the physical situation in the vicinity of a crack in a
reinforced concrete tension member. It indicates that at a crack the full load is
shared between steel and concrete. This ability of concrete between cracks to
share the tensile load with the reinforcement is called “tension stiffening”. As the
load is increased and the stress in the concrete between cracks reaches the ultimate
strength, the concrete will then rupture and a further crack forms between the
main cracks. Therefore, the average concrete stress over the cracked region will
progressively decrease with loading.

Figure 4.4 Stresses distribution in cracked reinforced concrete (Gerstle 1981)

As discussed in Chapter 2, two methods can be employed to represent the


tension stiffening effect in the finite element analysis.

58
4.2.4.1 Model of tension stiffening by increasing the stiffness of reinforcing
steel after cracking

The tension stiffening effect can be directly represented by modifying the


steel stress, as discussed in chapter 2. That is to say, the steel stiffness is increased
as illustrated in the modified stress-strain diagram in Figure 2.12(c). Thus the
force carried by the steel also includes the equivalent tensile force carried by the
concrete between the cracks. The added stress, due to the concrete tension
stiffening, is lumped at the steel level and oriented in the same direction as the
steel.

This method is not adopted in this research.

4.2.4.2 Model tension stiffening effect by retaining the concrete contributions


after cracking

By adding a descending branch to the uniaxial stress-strain curve of


concrete in tension to take the tension stiffening effect into account, as shown in
Figure 2.12(a), Figure 2.12(b). In conjunction with the smeared models of cracked
concrete and reinforcement used in this research, Stevens et al. (1991) is adopted,
but with some modifications.

In this model, Stevens et al. proposed a descending branch of the stress-


strain curve of cracked concrete as follows,

σ1
= (1 − β b )e −λ (ε −ε
t 1 1c )
+ βb (4.3)
σ 1c
ρ
β b = 75 s (4.4)
db
270
λt = (4.5)
βb

where σ 1c and ε 1c are concrete stress and strain when cracking, σ 1 and ε 1 are

stress and strain in the principal direction normal to the crack, λt is a parameter
controls the rate at which the post-peak stress-strain response decays to a limiting

59
value, ρ s is reinforcement ratio , d b is the diameter of steel bar, β b is a

parameter for the bond action of reinforcing bars crossing cracks, which reflects
the influence of amount and distribution of reinforcement.

Equation (4.3) shows the average tensile stress in the concrete reduces
exponentially from tensile stress at cracking σ 1c to a limiting value of β b σ 1c at
large tensile strains. In order to coincide with the tension softening equation
adopted in this thesis, according to He (1999), the following modification is made,

ε1
σ1 ε 1c
= (1 − β b ) + βb (4.6)
σ 1c 2 ε1 ε1
0.312 f t ( − 1) +
1.7

ε 1c ε 1c

4.3 Dowel Action

Dowel action is defined as the capacity of reinforcing bars to transfer


forces perpendicular to their axis. It is induced due to the counteraction of the
longitudinal bars bridging a crack to slip off the crack surfaces relative to each
other, as shown in Figure 4.5.

The phenomenon of dowel action as a shear transfer mechanism across


cracks has long been recognized as a component of the overall shear resistance
capacity of reinforced concrete beams. Shear transfer mechanisms in reinforced
concrete beams is provided by the shear resistance in the compression zone Vc ,

aggregate interlock force Va , stirrups tensile force Vs , and dowel action of

longitudinal reinforcing bars crossing the crack in the concrete, Vd , as shown in


Figure 4.6.

When seen in context of the other mechanisms, the contribution of dowel


action to the shear resistance can be relatively small (Hassan et al., 2008).

60
However, He and Kwan (2001) argued that dowel action can play an important
role if the other contributions to the shear resisting mechanisms become
insignificant. This may happen, for example, in the case of a beam with a small
amount of web reinforcement, or during the post-peak loading stages when, due to
widening of cracks, the contribution of aggregate interlock may decrease rapidly.
Dowel action may contribute significantly to the post-peak resistance and hence
contribute to the shear ductility of concrete members.

It is difficult to incorporate the effects of the dowel action in the nonlinear


analysis of reinforced concrete beams. The contributions of the compression zone,
the stirrups, and the aggregate interlock are fairly well understood in the literature,
but so far the dowel action of the reinforcing bars has not been well modeled. The
major difficulties in modeling the dowel action of reinforcement bars for finite
element analysis are presented as follows:

1. In experimental tests, the shear force transferred by the dowel action is


quite difficult to measure because it is embedded with other shear
transfer components. In fact, since the dowel action involves
interaction between the reinforcement bars and the surrounding
concrete, and the interaction stresses are extremely difficult to measure,
many details of the dowel action have never been investigated.
Consequently, experimental results on the dowel action have been
rather limited.

2. Even in finite element analysis, the mechanism of the dowel action is


too complicated to describe. To analyze the details of the dowel action,
only the linkage element used with discrete representation of
reinforcement and discrete crack can be adopted to represent dowel
action. Furthermore, a very fine mesh has to be used for the concrete.
As a result, the number of elements required would be very large.

3. Since the dowel action is usually more significant near peak load and
at the post-peak stage, experimental testing or theoretical analysis
extending into the post-peak range are needed to investigate the full
61
effects of the dowel action, but such testing and analysis are generally
quite difficult.

Though the modeling of the dowel action in the finite element analysis
with smeared cracks can be accomplished by making some assumptions, the
dowel action is simply neglected in this thesis.

Figure 4.5 Dowel action of reinforcing steel crossing a crack

62
Figure 4.6 Internal forces in a cracked beam

4.4 Modeling of Confinement Effects

Concrete restrained by transverse reinforcement is often termed of


confined concrete, which is in the form of closely spaced spirals or circular hoops,
or rectangular hoops with or without supplementary cross ties. Tests have shown
that concrete confined by suitable arrangements of transverse reinforcement can
result in a significant increase in both the strength and the ductility of compressed
concrete. In particular, the strength enhancement from confinement and the slope
of the descending branch of the concrete stress-strain curve have a significant
influence on the strength and ductility of reinforced concrete structures.

Numerous experimental tests have been conducted to study the behavior of


the confined concrete (Sheikh and Uzumeri 1980; Mander et al. 1988; and
Hoshikuma et al. 1997). The tests have demonstrated that the effect of
confinement is influenced by the following factors: the ratio of the transverse
reinforcement, the yield strength of the transverse reinforcement, spacing of hoops,
the hoop pattern, the longitudinal bars, and the additional supplementary cross ties.

63
Clearly it is important to be able to quantify these effects of confinement on the
stress-strain behavior of concrete.

The stress-strain model of Kent and Park (1971) for concrete confined by
rectangular transverse reinforcement neglected the increase in concrete strength
but took into account the increase in ductility due to rectangular confining steel.
More recently, Scott et al. (1982) and Park et al. (1982) modified the Kent and
Park (1971) stress-strain equations to take into account the enhancement of both
the concrete strength and ductility due to confinement and the effect of strain rate.
Monotonic stress-strain equations for concrete confined by rectangular-shaped
transverse reinforcement include those proposed by Vellenas et al. (1977) and
Sheikh and Uzumeri (1980). Park and Leslie (1977), Ahmad and Shah (1982) also
proposed various stress-strain relationships for confined concrete under
monotonic loading.

In all these models, the influence of various types of confinement is taken


into account by defining an effective lateral confining stress (confinement index),
which is governed by the configuration of the transverse and longitudinal
reinforcement.

In this thesis, the confinement effect of the transverse reinforcement is


introduced by adjusting the stress-strain relationship of concrete under
compression according to the confinement index proposed by Kappos (1991).
This model is represented in Figure 4.7 and can be summarized in the following
forms:

f cc = K f c (4.7)

ε 0 cc = K 2 ε 0c (4.8)

fy
K = 1 + a( ρ w )b (4.9)
fc

in which f cc is the confined concrete strength, f c is the unconfined compressive

strength, ε 0 cc is the corresponding strain to the confined concrete compressive


64
strength, ε 0 c is the corresponding strain to the unconfined concrete compressive

strength, K is the confinement index, ρ w is the volumetric ratio of transverse

reinforcement, f y is the yield strength of the transverse reinforcement, a and

b are the empirical coefficients which are determined by the hoop patterns. The
values of a and b suggested by Kappos are listed as follows,

a=0.55 and b=0.75 for single hoop patterns;


a=1.00 and b=1.00 for double hoop patterns;
a=1.25 and b=1.00 for multiple hoop patterns.

Figure 4.7 Stress-strain relationship proposed by Kappos (1991)

65
4.5 Summary

In this chapter, the interactions between reinforcing steel and the


surrounding concrete, i.e., bond action, dowel action and the confinement are
discussed.
The bond action is usually modeled by considering slip between
reinforcing steel and the surrounding concrete. However, this approach is not
adopted in the present study because discrete crack model and discrete
representation of reinforcement should be incorporated. Moreover, the local
behavior of slip is not the main concern of this study. Perfect bond between the
reinforcing steel and the concrete is assumed. The bond action is modeled based
on the tension stiffening effect in this research. This is accomplished by adding a
descending branch to the uniaxial stress-strain curve of concrete in tension. In this
thesis, some modifications of the equation of tension softening were made to take
tension stiffening effect into account.
Dowel action is defined as the capacity of reinforcing bars to transfer
forces perpendicular to their axis. However, only the linkage element used with
discrete representation of reinforcement and discrete crack can be adopted to
represent dowel action. Therefore, this effect is simply neglected in this research.
Concrete confined by transverse reinforcement can result in a significant
increase in both the strength and the ductility. In this thesis, the confinement effect
of the transverse reinforcement is modeled by adjusting the stress-strain curve of
concrete under compression according to the confinement index proposed by
Kappos (1991).

66
CHAPTER 5

NONLINEAR SOLUTION TECHNIQUES

5.1 Introduction

Due to its complicated properties, reinforced concrete cannot be treated in


the finite element method like a linear elastic material whose governing equations
can be solved in a direct manner. More sophisticated solution strategies have to be
employed when applying finite element procedures to the analysis of reinforced
concrete structures. One of the difficulties that may be encountered in a nonlinear
analysis is that numerical instabilities can occur if the chosen solution algorithm is
not suitable for the type of problem under consideration. For example, sudden
changes in element stiffness due to cracking and crushing of concrete and yielding
of reinforcement may cause extra difficulties in the analysis of reinforced concrete
structures. Therefore, accurate and efficient solution algorithms for the nonlinear
analysis of reinforced concrete structures are crucial for finding a meaningful
numerical solution.

An incremental finite element formulation, coupled with the use of an


iterative process until convergence to results of sufficient accuracy, is usually
used to trace out the entire structural response of reinforced concrete structures.
The numerical computation process can be highly dependent on the solution
techniques chosen in nonlinear finite element analysis. Reasons are discussed as
follows: (a) the choice of the load step sizes is indeed crucial, the load increment
size has to be made sufficiently small in order to achieve an accurate solution, but
the small increment can lead to a highly expensive computation system; (b) the
convergence process may be slow requiring a large number of iterations which
can again result in a high solution cost.

67
Many researchers have devoted for the development of efficient solution
algorithms for nonlinear problems. Some powerful alternatives to the traditional
Newton-Raphson and Modified-Newton-Raphson methods have been suggested.
These include the Quasi-Newton algorithm proposed by Matthies and Strang
(1979), the Secant-Newton methods advocated by Crisfied (1982), and the Arc-
length methods introduced by Riks (1979).

This Chapter describes the method of analysis used in the present work
and briefly introduces some other related methods. First, a brief description of the
incremental and iterative methods which are usually adopted in nonlinear analyses
is given in Section 5.2. Then a displacement control technique used in this study
to ensure the solution can be advanced in the post-peak range is introduced in
Section 5.3. Next in this chapter (Section 5.4), a discussion of the convergence
problems is given and residual load criterion is adopted in this study. In Section
5.5, criteria to define the collapse of the structure are suggested. Last in Section
5.6, a brief summary is given.

5.2 Basic Techniques for Solving Nonlinear Equations

The governing equation in this study is based on the equilibrium between


the external applied loads and the internal forces. After the finite element
discretization of the nonlinear governing equation, a system of algebraic equations
as the following form can be achieved,

{R(δ )} = {F (δ )} − {P} (5.1)

where
{R} is the residual force vector,
{F }= ∫VB T σ dv is the internal force vector,
{P} is the applied load vector,
{δ } is the unknown displacement vector, which expressed as follows,

68
{δ } = [k (σ , ε )]−1 {P} (5.2)

Equation (5.2) illustrates the basic nonlinear relationship between {δ } and


{P}, due to the influence of the nonlinear material law [k ] .

The satisfaction of equilibrium at the nodes requires that the external load
vector equals to the internal load vector, that is to say, the residual force vector
should be zero.

{R(δ )}= 0 (5.3)

The solution of Equation (5.3) is usually attempted using one of the


following three basic techniques:

1. Incremental (step-wise procedure).


2. Iterative (Newton method).
3. Incremental-iterative (mixed procedure).

The different methods will in general lead to different load-deflection


paths influencing the final solution. In this study, the mixed incremental-iterative
method is adopted. In this method, the external loading is applied in increments
and the solution corresponding to each load increment is obtained by iterating
until convergence is achieved. A good review of iterative methods can be found in
Crisfield (1982); a brief review of them will be given next.

5.2.1 Basics for iterative methods

Much effort has been devoted to the development of efficient iterative


algorithms. In these methods the load is applied incrementally and a series of
iterations are carried out within each increment in order to eliminate the out of
balance forces {R}.

69
The first step is to calculate an initial estimate of the incremental
displacement {∆δ }1 , then additional corrective displacements {∆δ }i are obtained
n n

by iterations. (Note that superscript n denotes the load increment number while
subscript i corresponds to the current iteration number.)

An improved value of displacement {δ }i +1 can be obtained by computing


n

{δ }in+1 = {δ }in + {∆δ }in (5.4)

The incremental displacements {∆δ }i can be calculated from


n

{∆δ }in = [k ]in {R}in


−1
(5.5)

The solution is carried out by solving a new set of linearized equations,


Equation (5.5). Displacements are then updated using Equation (5.4) and the
residual forces are calculated from Equation (5.1). This process is repeated until
the solution converges.

5.2.2 Iterative methods

Based on the manner in which the stiffness matrix is formed, several


iterative methods can be classified.

5.2.2.1 Full Newton-Raphson method

This solution can be illustrated diagrammatically in Figure 5.1. It is clear


from Figure 5.1 that the Newton-Raphson method requires the updating of the
stiffness matrix [k ]i , and the solution of a new set of equations at each iteration.
This process can be very expensive. The repeated formations and decompositions
of the stiffness matrix require much more time than that required for system of

70
equations. To overcome this difficulty, a modification to the full Newton-Raphson
algorithm is made.

5.2.2.2 Modified Newton-Raphson method

This solution can be illustrated diagrammatically in Figure 5.2. In this


method, the initial linear elastic stiffness matrix of each increment is used and
held as constant throughout the entire increment, after which it is again updated.
This method is more economical than the Newton-Raphson method since it
involves fewer expensive reformations and decompositions of the stiffness matrix,
however it may reduce the convergence rate and require more iterations than those
required using the full Newton-Raphson method. This iterative method is adopted
in this research to form the stiffness matrix [k ] .

5.2.2.3 Initial stress method

This method was first advocated by Zienkiewicz and his coworkers for the
solution of elastic-plastic problems. The method can be thought of as a
modification to the Newton-Raphson algorithm where the initial elastic stiffness
matrix is maintained during the entire analysis, see Figure 5.3. This has the
immediate advantage of significantly reducing the computing cost per iterations,
however it shows less efficient than the previous Newton-Raphson methods
because the convergence rate is too slow.

71
Figure 5.1 Full Newton-Raphson method

Figure 5.2 Modified Newton-Raphson method

72
Figure 5.3 Initial stress method

5.3 Displacement Control Technique

5.3.1 Introduction

Two techniques were used to control the solution of the governing


equations. They are load control technique and displacement control technique. In
the former one, the independent variable is the applied load and the nodal
displacements are the unknowns in the solution of the governing equations. While
in the latter one, a dominant displacement component is chosen as an independent
variable instead of the usual load parameter, λ , and for a system of n degrees of
freedom the solution is carried out for (n-1) unknown displacements plus the load
parameter.

Consider the structural response shown in Figure 5.4, the load control
technique can be used to trace the ascending part of the response before the limit
point, and obviously fails to trace the falling part of the response. The arc length

73
approach and the displacement control techniques can be used to trace the
nonlinear response in the analysis of such problems. Although the arc length
method has been used by many investigators for nonlinear problems owing to its
ability to handle snap-through and snap-back problems, and its improved
convergence characteristics over the load control, it has been found to fail for
some problems involving the nonlinear analysis of reinforced concrete structures.
This is likely due to the fact that the arc length method may pick up unstable
positions along the falling part of the response.

Instead of making the problem more difficult, the arc length method is not
incorporated into this work. A displacement control technique similar to that
proposed by Batoz and Dhatt (1979) was adopted in the analysis. In this technique
an incremental displacement component is specified and the corresponding load
becomes one of the unknowns.

5.3.2 Formulation of displacement control method

A single incremental nodal displacement component is selected and


specified as a controlling parameter while the corresponding load value is
considered one of the unknowns. The nonlinear governing equations, Equation
(5.1), can be rewritten as

{R} = {F }− {P} (5.6)

in which the applied force vector {P} can be expressed in terms of an load factor

λ and a generalized force vector {P}* due to a unit load, such that

{P}= λ {P}* (5.7)

Substituting Equation (5.7) into Equation (5.6),

{R} = {F } − λ {P}* (5.8)


74
If at a load level λi the out of balance forces Ri (λi ) is obtained then the

out of balance forces at an adjacent load level λi + ∆λi is

Ri (λi + ∆λi ) = Ri (λi ) − ∆λi {P}


*
(5.9)

Applying the Newton-Raphson iterative technique to Equation (5.9), a


search direction can be found as

∆δ i (λi + ∆λi ) = ∆δ ia + ∆λi ∆δ ib (5.10)

where

∆δ ia = − k i−1 Ri (λi ) (5.11a)

and

∆δ ib = ki−1 {P}
*
(5.11b)

If the displacement is prescribed at the qth row of {δ }i , then

∆δ iq (λi + ∆λi ) is zero, and the qth row of Equation (5.10) can be written as

∆δ iq (λi + ∆λi ) = (∆δ ia ) q + ∆λi (∆δ ib ) q =0 (5.12)

from which the increment of the load factor is calculated as

∆λi = − (∆δ ia ) q / (∆δ ib ) q (5.13)

and the total load factor can be updated,

λi +1 = λi + ∆λi (5.14)

75
The main operations performed in the displacement control method are
shown in Figure 5.5. At the beginning of each increment a load factor λ is set
equal to zero and the incremental displacement component is assigned a preset
value. The incremental load factor was obtained from Equation (5.13) and hence
the incremental displacement vector can be determined from Equation (5.10). At
the end of each iterations, the load factor λ was updated by using Equation (5.14),
also the nodal displacements were then updated using Equation (5.4) and the
process was repeated until convergence occurred.

Figure 5.4 Typical nonlinear structural response

76
Form load vector due to unit load {P}
*

Initialize the load factor λ =0

Specify an incremental nodal


displacement component {δ }i
q

Determine ∆λi and {∆δ i }

Update the nodal displacement


{δ }in+1 = {δ }in + {∆δ }in

Update the load factor


λi +1 = λi + ∆λi

Determine the applied load vector


{P}= λ {P}*

Calculate internal load vector {F }

Calculate residual load vector


{R} = {F }− {P}

Convergence
Yes
No

Set the specified incremental


displacement component equal to zero
{∆δ i }=0
Figure 5.5 Flowchat for displacement control technique
77
5.4 Convergence Criteria

In an incremental-iterative solution technique, it is impractical and


unnecessary to satisfy the strict equality

{R(δ )} − F {δ } − λ (δ ){P}* =0 (5.15)

Therefore, the Equation (5.15) is replaced by the approximate condition

{R(δ )}i − F {δ }i − λ (δ ) i {P}* ≈ 0 (5.16)

where, during the iterative process, the progress of the approximate solution is
referred to specified convergence criteria. In the solution strategy, the solution
obtained at the end of each iterations is checked to see whether it has converged
within a preset tolerance. The criteria used are set so that the iterative process is
terminated when it is considered that additional iterations would not improve the
accuracy of the solution significantly. Thus, it is important to include reliable
criteria which will terminate the iterative process when convergence to the desired
accuracy has been achieved.

The convergence criteria, usually used for nonlinear structural analysis, are
based on (i) displacements, (ii) out-of balance forces, and (iii) internal energy. In
this thesis, a force convergence criterion is adopted.

It is difficult and expensive to check the decay of residual forces for every
degree of freedom thus some overall evaluation is preferable. This is achieved by
using norms. The residual load convergence criterion is that

R i
× 100% ≤ FTL (5.17)
λP *
i

where i
is the norm at iteration i, and R i = {R}Ti {R}i
78
and FTL is a specific convergence tolerance. The convergence tolerance must be
realistic. If the convergence tolerance is too loose, inaccurate results are obtained
and if the tolerance is too tight, much expensive effort is spent to obtain needless
accuracy.

The selection of convergence tolerance of the finite element method is


discussed later in Chapter 6.

5.5 Termination of the Analysis

Collapse of a structure takes place when no further loading can be


sustained. A criterion for total collapse is necessary to any nonlinear finite
element program as a mean to terminate the analysis. Several criteria may be used
to terminate an analysis, such as the maximum deflection, the maximum number
of iteration, and the growth of the dissipated energy.

In this work, the program is stopped when any of the following occur:

1. The number of iterations exceeds a preselected maximum number. A large


number of iterations in one step means that very large strain increments
take place during this step and the equilibrium can not be satisfied under
the applied loads. However, this criterion can be satisfied when the
solution is slowly converging or very tight convergence tolerances are
used. Therefore, a relatively large value of the maximum number of
iterations must be adopted.

2. The number of increments is greater than the maximum number of


increments. This is especially useful when displacement control technique
is used and the solution has passed the maximum load level.

3. The strain of reinforcing steel bars exceeds the ultimate strain permitted.

79
5.6 Summary
This chapter introduces the nonlinear solution techniques adopted in this
thesis. They can be summarized as follows:

1. Modified Newton-Raphson method in which the stiffness matrix is


updated at the first iteration of an increment is adopted. This method
shows a good performance of convergence and computational economy.

2. The displacement control techniques can be used to trace the whole


nonlinear response of reinforced concrete structures. A displacement
control technique similar to that proposed by Batoz and Dhatt (1979) was
adopted in the analysis. Detailed formulation of this method was given.

3. The convergence of the iterative schemes can be checked using the force
convergence criteria, the selection of the convergence tolerance will be
discussed in Chapter 6.

4. The analysis is terminated whenever the number of iterations or the


number of increments exceeds a preselected number or a steel bar has
fractured.

80
CHAPTER 6

APPLICATION TO REINFORCED CONCRETE


BEAMS

6.1 Introduction

In the previous chapters efforts have been made to discuss various models
for the finite element analysis of reinforced concrete structures. In this chapter the
computer program developed in this study is critically examined by carrying out
the numerical solution of a number of reinforced concrete structures. The results
of the analysis are compared with results obtained experimentally.

The application of previous introduced reinforced concrete model in this


chapter will be based mainly on three simply supported reinforced concrete beams.
These beams are specimen OA-1 and specimen A-1 tested by Bresler and
Scordelis (1963), and specimen J-4 tested by Burns and Siess (1962).

Moreover, a wide range of the parameters affecting the behavior of


reinforced concrete beams are considered in this chapter. It is important to
understand how the various numerical parameters influence the nonlinear
solutions of the finite element method. Otherwise, numerical problems could arise
which could cause divergence of a solution, or worse, give erroneous results even
though convergence had been achieved. A detailed discussion of finite element
mesh size effect, confinement effect of secondary reinforcement, tension
stiffening effect, and convergence tolerance will be given.

In all these case studies, the concrete is modeled by bilinear quadrilateral


elements with 2x2 Gauss integration, and the reinforcement is modeled by
smeared representation.

81
6.2 Classic Beam Specimens

6.2.1 Beam OA-1


Specimen OA-1 is a simply supported beam with a span of 3.66m which is
subjected to a concentrated load at mid-span. The geometry and the cross section
of beam OA-1 are shown in Figure 6.1 and the material properties are summarized
in Table 6.1.

6.2.2 Beam A-1


Specimen A-1 is similar to specimen OA-1 except that A-1 has transverse
reinforcement while OA-1 has none. The geometry and the cross section of beam
A-1 are shown in Figure 6.2 and the material properties are summarized in Table
6.1.

6.2.3 Beam J-4


Specimen J-4 is a simply supported beam with a span of 3.6m which is
subjected to a concentrated load at mid-span. The geometry and the cross section
of beam J-4 are shown in Fig. 6.3 and the material properties are summarized in
Table 6.1.

Examples Material Properties


f c (MPa) f t (MPa) f y (MPa) f y' (MPa) εt

OA-1 22.5 2.32 555 0.00015

Beam A-1 24.1 2.32 555 345 0.00015

J-4 33.32 3.0 309 0.00015

Table 6.1 Material properties used in application

82
Figure 6.1 Details of beam OA-1 (Bresler and Scordelis, 1963)

Figure 6.2 Details of beam A-1 (Bresler and Scordelis, 1963)

83
Figure 6.3 Details of beam J-4 (Burns and Siess, 1962)

84
6.3 Analysis of Beams

Beam OA-1 and beam A-1 are regarded as a classic test series, and have
been widely used as benchmark data for verifying the finite element models for
reinforced concrete. The details of these two beams are discussed in Section 6.2.
The finite element mesh boundary condition, and loading used in the analysis are
shown in Figure 6.4. Because of symmetry, only half of the beam will be
considered in this study.

The load-deflection curves of these two beams derived from the nonlinear
finite element analysis are shown in Figure 6.5. A good agreement with the
experimental results is achieved throughout the entire load-deflection range.

The ultimate loads and the corresponding mid-span deflections of the two
beams are listed in Table 6.2. The failure load of beam OA-1 tested by Bresler and
Scordelis is 333.6 kN with the corresponding mid-span deflection of 6.6 mm. The
numerical prediction of failure load in this study is 309 kN with the corresponding
mid-span deflection of 6.9 mm. And the failure load of beam A-1 tested by
Bresler and Scordelis is 467 kN with the corresponding mid-span deflection of
13.5 mm. The numerical prediction of failure load in this study is 502 kN with the
corresponding mid-span deflection of 14.7 mm. All the numerical results obtained
show a good agreement with that of the experimental study.

85
Figure 6.4 Finite element mesh used

Beam name Ultimate load (kN) Mid-span deflection


(mm)

OA-1 Experimental 333.6 6.6

Analytical 309 6.9

A-1 Experimental 467 13.5

Analytical 502 14.7

Table 6.2 Ultimate loads and the corresponding deflections of beam OA-1 and A-
1

86
400

350

300

250
Load (kN)

Experimental result
200
Analytical result

150

100

50

0
0 2 4 6 8
Mid-span deflection (mm)

(a) Load-deflection curve of beam OA-1

600

500

400
Load (kN)

Experimental result
300
Analytical resullt

200

100

0
0 5 10 15 20
Mid-span deflection (mm)

(b) Load-deflection curve of beam A-1


Figure 6.5 Load-deflection curves of beam OA-1 and A-1

87
The finite element simulation can also successfully model the development
of cracks and explain the failure pattern of the reinforced concrete beams. The
predicted development of cracks of beams OA-1 and A-1 are shown in Figure 6.6
and Figure 6.7, respectively. These load levels represent the most important stages
of behavior. First at a relatively low load level, for example 100kN for beam OA-
1, the typical flexural cracks would develop at the bottom of the beam running
towards the loading point. Then the diagonal tension cracks would appear, usually
in the middle third of the overall beam depth and at various sections along the
span. These diagonal cracks extended both upwards and downwards with further
increase in load and form the “critical diagonal tension crack.” The critical cracks
formed at a load of 267kN for both beams OA-1 and A-1 according to the
experimental data. The predicted crack patterns of beam OA-1 and A-1 at this
load level are shown in Figure 6.6 (b) and Figure 6.7 (b), respectively.

Beam OA-1, which has no transverse reinforcement, fails shortly after the
formation of the critical diagonal crack. The failures occurred as a result of
longitudinal splitting in the compression zone near the load point, and also by
horizontal splitting along the tensile reinforcement near the end of the beam. The
crack pattern at failure load is shown in Figure 6.6 (c).

Beam A-1, which has transverse reinforcement, fails at loads substantially


greater than the load at which the initial diagonal tension crack occurred; and the
finial failure occurred by splitting in the compression zone but without splitting
along the tension reinforcement which is characteristic of beams without
transverse reinforcement. This is because the transverse reinforcement resists the
widening of the diagonal cracks and prevents the splitting of the concrete along
the longitudinal steel bars. The applied load can increase until the crush of
concrete near the loading point. The crack pattern at failure load is shown in
Figure 6.7 (c).

88
(a) Crack pattern of beam OA-1 at load= 100kN

(b) Crack pattern of beam OA-1 at load= 250kN

(c) Crack pattern of beam OA-1 at load= 309kN

Figure 6.6 Crack pattern of beam OA-1

89
(a) Crack pattern of beam A-1 at load= 180kN

(b) Crack pattern of beam A-1 at load= 300kN

(c) Crack pattern of beam A-1 at load= 502kN

Figure 6.7 Crack pattern of beam A-1

90
6.4 Parametric Study

Scordelis (1972), Gerstle (1981), Nilson et al. (1978) and Crisfield (1982)
made a comprehensive review of various finite element methods for analyzing
reinforced concrete structures. In their work, many numerical parameters such as
aggregate interlocking, tension stiffening, and dowel action have been studied.
Although much progress has been made in the analysis of reinforced concrete
structures, a lot of the numerical parameters have still not been studied in
sufficient detail.

To demonstrate the effect of some of the important numerical parameters


on the solution of nonlinear finite element problems, beam OA-1, bean A-1 and
beam J-4 have been employed in a parametric study.

6.4.1 Comparison study on confinement effect


This study is carried out by directly comparing the analytical results of
beam OA-1 and beam A-1 obtained already in Section 6.3. In this comparison
study, the following considerations are taken:

(a) Displacement control technique was used.


(b) Modified Newton-Raphson method was used.
(c) The load convergence criterion was used, the convergence tolerance was
5%.
(d) The same finite element mesh was used.
(e) The tension stiffening effect was incorporated.
(f) The slight difference of f c between these two beams was neglected.

The comparison of the load-deflection curves of these two beams is shown


in Figure 6.8.

91
600

500

400
Load (kN)

Beam OA-1
300
Beam A-1

200

100

0
0 5 10 15 20
Mid-span deflection (mm)

Figure 6.8 The load-deflection curves of beam OA-1 and beam A-1

Comparison of the deflections of beam OA-1 with beam A-1 indicates the
effect of transverse reinforcement on the deflections. It is obvious that the
ascending parts of these two beams are almost identical for low levels of load; this
is because for low levels of strain in concrete, the stress state in the transverse
reinforcements is very small and the concrete is basically not confined.
Conversely, when the stress in the concrete gets close to its uniaxial compression
strength, progressive internal cracking occurs and the transverse strain increases
remarkably. At this stage, transverse reinforcement produces a confinement action
which opposes the expansion of the concrete core. Thus, the specimen with
transverse reinforcement can continue to take external loads to a relatively high
level, and shows a much ductile load-deflection performance.

Moreover, the slope of the load-deflection curve of beam OA-1 is


approximately the same with that of beam A-1, which indicates the stiffness of
beams are not influenced appreciably by the addition of transverse reinforcement.

92
6.4.2 Comparison study on finite element mesh size effect

Finite element mesh size is one of the solution parameters, which can
significantly influence the finite element solution process. The effect of finite
element mesh size is analyzed based on beam OA-1. In order to study the effects
of finite element mesh size on the analytical results, two different mesh
configurations were investigated, as shown in Figure 6.9.

In this comparison study, the following conditions remain the same


throughout the analysis:

(a) Displacement control technique was used.


(b) Modified Newton-Raphson method was used.
(c) The load convergence criterion was used, the convergence tolerance was
5%.
(d) The tension stiffening effect was incorporated.

The analytical results with both fine mesh and coarse mesh and the
experimental results are shown in Figure 6.10. The numerical prediction of the
failure load of beam OA-1 with fine mesh has been discussed in section 6.3. It
shows a good agreement with the experimental results. However, for the coarse
alignment, the predicted failure load is 34.7% higher than that of the experimental
data. This is because when the element size is sufficiently small, the tensile strains
and stresses in the adjacent element just ahead of the crack tip could become
sufficiently large, and the crack could propagate under a sufficiently small load.
Consequently, the smaller the element size, the lower the predicted structural
capacity would be; the larger the element size, the higher the predicted structural
capacity would be.

In summary, the accuracy of the prediction by using finite element method


is largely reduced when coarse mesh is adopted. Thus a well-designed finite
element mesh is really crucial for analyzing reinforced concrete structures with
finite element method.

93
(a) Fine mesh (48 elements)

(b) Coarse mesh (8 elements)

Figure 6.9 Concrete finite element mesh for beam OA-1

94
500

450

400

350
Experimental result
300
Load (kN)

Analytical result (fine


250
mesh)
200 Analytical result
(coarse mesh)
150

100

50

0
0 2 4 6 8
Mid-span deflection (mm)

Figure 6.10 Comparison of load-deflection curves of beam OA-1 (fine and coarse
mesh)

6.4.3 Comparison study on tension stiffening effect and


convergence tolerance

Tension stiffening effect is an important characteristic of reinforced


concrete structures. It should be incorporated into the computational model if
accurate load-deflection responses of reinforced concrete structures are desired.
The aim of this section is to evaluate the effect of tension stiffening in the
reinforced concrete analysis. A comparison between the tension stiffening model
and no tension stiffening model is made in this section; and the interaction of the
convergence tolerance and the tension stiffening model is also investigated in this
section. The particular tension stiffening stress-strain curve (Equation 4.6) is used
for analysis.

95
Beam J-4 is an under-reinforced beam with longitudinal steel ratio equals
to 0.99%, which is used in this comparison study. And the following conditions
remain the same throughout the analysis:

(a) Displacement control technique was used.


(b) Modified Newton-Raphson method was used.
(c) The load convergence criterion was used.
(d) The same finite element mesh was used.

Since convergence tolerance and tension stiffening effect are two


parameters focused on this section, some selected load-deflection curves of beam
J-4 under different combinations of theses two parameters are plotted in Figure
6.11, Figure 6.12, and Figure 6.13. In addition, the numerical predicted ultimate
loads of beam J-4 under different conditions are listed in Table 6.3. For all these
figures and table, TS means tension stiffening and FTL means convergence
tolerance.

First see Figure 6.11, load-deflection curves are compared for convergence
tolerances of 10% and 5%. It is observed that no matter tension stiffening effect is
considered or not, the predicted ultimate loads of beam J-4 with convergence
tolerance= 10% are higher than those with convergence tolerance= 5%. For
example, when tension stiffening is considered, the predicted ultimate load with
tolerance of 10% is 13.6% higher than that with tolerance of 5%. This is because
using a high convergence tolerance is a method of maintaining a certain
proportion of residual forces in the structure as this slows down the rate of
cracking. Thus the predicted ultimate load increases. It is also observed that when
tension stiffening is used, a high convergence tolerance of 10% overestimate the
ultimate load, and a low convergence tolerance of 5% underestimate the ultimate
load. This shows again that convergence tolerance can influence the numerical
results significantly.

Then see Figure 6.12, when the convergence tolerance set to be identical,
the predicted ultimate loads with tension stiffening effect are higher than those
96
without tension stiffening effect. For example, when the convergence tolerance is
set to be 5%, the predicted ultimate load with tension stiffening effect is 12.5%
higher than that when neglecting tension stiffening effect. This phenomenon is
quite similar with the influence of convergence tolerance. That is to say, a high
convergence tolerance and consideration of tension stiffening can both lead to an
overestimation of the stiffness and the ultimate load. Therefore, a conclusion that
tension stiffening has the same effect of retaining residual forces which can lead
to an overestimation of the stiffness and the ultimate load can be drawn.

As shown in Figure 6.13, tension stiffening model with convergence


tolerance of 5% and no tension stiffening model with convergence tolerance of
10% gave the best fit to the experimental load-deflection curves. This indicates
that tension stiffening model can produce satisfactory results, however is very
interrelated to the convergence tolerance selected. For example, when tension
stiffening effect is included with a high convergence tolerance of 10%, the
predicted ultimate load is overestimated by 16.5%; but when tension stiffening
effect is included with a low convergence tolerance of 5%, the predicted ultimate
load is only overestimated by 2.5%. This is because high convergence tolerance
and tension stiffening both have the effect of overestimating the ultimate forces as
discussed before. Thus, when tension stiffening is used, a small convergence
tolerance should be selected; when tension stiffening is neglected, a relatively
high convergence tolerance should be selected.

97
Ultimate load (kN)

Analytical Experimental

TS, FTL=10% 184

TS, FTL=5% 162 158

No TS, FTL=10% 152

No TS, FTL=5% 144

Table 6.3 The ultimate loads of beam J-4 under different combinations

98
180

160

140

120
Load (kN)

100 No TS, FTL=10%


No TS, FTL=5%
80 Experimental

60

40

20

0
0 2 4 6 8 10 12
Mid-span deflection (m m )

200

180

160

140

120
Load (kN)

TS, FTL=10%
100 TS, FTL=5%
Experimental
80

60

40

20

0
0 2 4 6 8 10 12
Mid-span deflection (mm)

Figure 6.11 Comparison of load-deflection curves of beam J-4 to study


convergence tolerance

99
200

180

160

140

120
Load (kN)

No TS, FTL=10%
100 TS, FTL=10%
Experimental
80

60

40

20

0
0 2 4 6 8 10 12
Mid-span deflection (mm)

180

160

140

120
Load (kN)

100 No TS,FTL=5%
TS, FTL=5%
80 Experimental

60

40

20

0
0 2 4 6 8 10 12
Mid-span deflection (mm)

Figure 6.12 Comparison of load-deflection curves of beam J-4 to study tension


stiffening

100
200

180

160

140
Experimental
120
Load (kN)

No TS, FTL=10%
100 No TS,FTL=5%
TS, FTL=5%
80
TS, FTL=10%
60

40

20

0
0 5 10 15
Mid-span deflection (mm)

Figure 6.13 Load-deflection curves of beam J-4 under different conditions

6.5 Summary

In this chapter, the models discussed in the previous chapters, are validated
by analyzing three reinforced concrete beams. The good agreement between
analytical and experimental results demonstrates that the developed finite element
code can provide reliable predictions of the load-deflection response and the crack
pattern of reinforced concrete structures. Then a parametric study is conducted in
order to understand how confinement, finite element mesh size, tension stiffening,
and convergence tolerance influence the numerical results.

101
The findings from parametric study can be summarized as follows:

1. Reinforced concrete beams with transverse reinforcement fail at higher


loads and are capable of developing substantially higher deflections, thus
exhibiting greater "ductility". This is mainly due to the confinement action
provided by transverse reinforcement which opposes the expansion of the
concrete core and restricts the development of cracks under high level of
loads.

2. A well-designed finite element mesh is important for analyzing reinforced


concrete structures with finite element method. The accuracy of the
analytical results will be largely reduced if coarse mesh is adopted.

3. Tension stiffening has a significant influence on the post cracking


response of reinforced concrete structures. Neglecting tension stiffening
may significantly underestimate the ultimate load.

4. Tension stiffening was originally introduced to improve the accuracy of


the results; however, it has been given various physical interpretations. For
example, tension stiffening is a way of retaining residual forces indirectly.
Meanwhile, convergence tolerance has the same effect of retaining
residual forces. A higher convergence tolerance allows a higher level of
residual forces remained in the structure. Therefore, tension stiffening can
be replaced by using a no tension stiffening model with a high
convergence tolerance sometimes.

102
CHAPTER 7

CONCLUSIONS AND RECOMMENDATIONS

7.1 Summary

A nonlinear finite element program to simulate the behavior of reinforced


concrete beams under the monotonic loading has been developed in this research.
Nonlinear behaviors of the material and the interactions between concrete and
reinforcing steel have been considered. Furthermore, efficient and reliable
solution techniques are discussed. At last, the developed finite element program in
this study is used to analyze three reinforced concrete beams and parametric
studies are carried out.

The stress-strain relationship of concrete under biaxial loading conditions


is described as a nonlinear orthotropic model, in which the axes of orthotropy
coincide with the principal stress directions. The stress updating is carried out
using the secant modulus.

Cracking of concrete is controlled by the limiting tensile strain failure


criterion of concrete. A smeared crack representation is used. Concrete can crack
along one or two directions, and cracks are allowed to open or close at each
sampling point in the present research. Tension stiffening function, based on the
tension softening stress-strain relationship proposed by Guo and Zhuang (1987), is
incorporated in the cracking model. And shear retention model used in this study
is a simple parabolic function which depends on the crack width.

A bilinear stress-strain relationship is employed for reinforcing steel in this


thesis. And the smeared representation is used.

103
The nonlinear equations of equilibrium are solved by using incremental-
iterative techniques. The iterative technique used in this study is modified
Newton-Raphson method, and displacement control increment method is used.

The capabilities of the developed program in analyzing reinforced concrete


plane stress problems are examined and verified by using it to analyze several
reinforced concrete beams. The results obtained analytically are compared with
the experimental results and it shows a good agreement. Then the program is
utilized to study the effect of some material and numerical parameters.

7.2 Conclusion

Based on the results and discussions in the previous chapter, the following
conclusions can be made:

1. Transverse reinforcement can barely influence the stiffness of the


reinforced concrete beams. However, transverse reinforcement can have
significant effects on the ultimate loads and ductility of the reinforced
concrete beams. Both of the ultimate loads and the ductility of the beams
can be improved by adding transverse reinforcement into the beams.

2. Comparison study on the finite element mesh size effects was carried out.
Two different mesh configurations were investigated, one is a fine mesh
and the other is a coarse mesh. The comparison of the analytical results
obtained from these two meshes demonstrates that finite element mesh
size is one of the most important parameters which can influence the finite
element solution results. The fine mesh is capable of obtaining a much
better solution than the coarse mesh.

3. By comparing a series of the load-deflection curves of reinforced concrete


beam J-4, the effects of tension stiffening and convergence tolerance can

104
be found. Satisfactory results can be obtained when the tension stiffening
effects were incorporated in the computational model, however it is very
interrelated to the convergence tolerance used in the study.

4. The load-deflection curves obtained from tension stiffening model with


low convergence tolerance are close to those obtained from no-tension
stiffening model with high convergence tolerance. This is because both
approaches have the effect of retaining residual forces, the former
indirectly and the latter directly.

7.3 Recommendations for Future Research

1. Analytical results of load-deflection behavior of reinforced concrete beams


by using the finite element method is significantly influenced by element
types and the integration rules. It is well known that 2 × 2 Gauss rule,
which is adopted in this thesis, is the minimum requirement for a parabolic
element. In the future research, finer finite elements and higher order
integration should be incorporated in order to derive better results.

2. Because the main objective of this research is to analyze the load-


deflection behavior of reinforced concrete beams, the bond slip and the
local behavior are not discussed well in this study. Future research should
be conducted to develop properly models of bond.

3. Improving the cracking model by employing the non-orthogonal cracks


model or multiple cracks model in the future research.

4. More parametric studies can be carried out to investigate the effect of


different parameters such as shear retention factor and number of iterations
required to achieve an acceptable solution.

105
5. The adopted computational model is only suitable for monotonic loading.
Further improvement should be made to extend this model to include the
analysis of reinforced concrete structures under cyclic loading.

106
REFERENCES

Ahmad, S. H., and Shah, S. P. (1982). “Complete Triaxial Stress-Strain Curves for
Concrete.” Journal of the Structural Division, ASCE, 108(4), 728-742.

Al-Mahaidi, R. S. H. (1979). “Nonlinear Finite Element Analysis of Reinforced


Concrete Deep Members.” Report 79/1, Cornell University, USA.

Ansari, F. (1987). “Stress-strain Response of Microcracked Concrete in Direct


Tension.” ACI Material Journal, 84(6), 481-490.

Al-Manaseer, A. A. (1983). “A Nonlinear Finite Element Study of Reinforced


Concrete Beams.” PhD thesis, University of Glasgow, UK.

Andenaes, E., Girstie, K., and Ko, H. Y. (1977). “Response of Mortar and
Concrete to Biaxial Compression.” Journal of Engineering Mechanics Division,
ASCE, 103(4), 515-526.

ASCE Task Committee of Finite Element Analysis of Reinforced Concrete


Structures. (1982). State-of-the-Art Report on Finite Element Analysis of
Reinforced Concrete, ASCE Special Publications. ASCE, New York, USA.

Barros, M. H. F. M., Martins, R. A. F., and Ferreira, C. C. (2001). “Tension


Stiffening Model with Increasing Damage for Reinforced Concrete.” Engineering
Computations, 18(5-6), 759-785.

Batoz, J. L. and Dhatt, G. (1979). “Incremental Displacement Algorithms for


Nonlinear Problems.” International Journal for Numerical Methods in Engineering,
14(8), 1262–1267.

Bazant, Z. P., and Bhat, P. D. (1976). “Endochronic Theory of Inelasticity and


Failure of Concrete.” Journal of Engineering Mechanics Division, ASCE, 102(4),
701-722.

107
Beshara, F. B.A. (1991). “Nonlinear Finite Element Analysis Reinforced Concrete
Structures Subjected to Blast Loading.” PhD thesis, City University, London, UK.

Bresler, B. (1966). “Influence of Load History on Cracking in Reinforced


Concrete.” Department of Civil Engineering, Division of Structural Engineering
and Structural Mechanics, University of California, Berkeley, USA.

Bresler, B., and Scordelis, A. C. (1963). “Shear Strength of Reinforced Concrete


Beams.” ACI Journal, 60(1), 51-74.

Burns, N. H., and Siess, C. P. (1962). “Load-Deflection Characteristics of Beam-


Column Connections in Reinforced Concrete.” University of Illinois, Urbana.

Carino, N. J., and Slate, F. O. (1976). “Limiting Tensile Strain Criterion for
Failure of Concrete.” ACI Journal, 73(3), 160-165.

CEB-FIP. (1993). CEB-FIP Model Code 1990: CEB Bulletin d'Information 213-
214, Thomas Telford Service Ltd., London, England.

CEB-FIP. (2000). Bond of Reinforcement in Concrete, State-of-Art Report, FIP


bulletin 10, Switzerland.

Cedolin, L., and Dei Poli, S. (1977). “Finite Element Studies of Shear Critical
Reinforced Concrete Beams.” Journal of the Engineering Mechanics Division,
ASCE, 103(3), 395-410.

Cervenka, V. (1985). “Constitutive Model for Cracked Reinforced Concrete.”


ACI Material Journal, 82(6), 877-882.

Chan, H. C., Cheung Y . K., and Huang, Y. P. (1994). “Heterogeneous Hardening


Plasticity Model for Concrete.” Journal of Structural Engineering, ASCE, 120(1),
46-62.

108
Chen, A. C. T., and Chen, W. F. (1975). “Constitutive Relations for Concrete.”
Journal of Engineering Mechanics Division, ASCE, 101(4), 465-481.

Chen, W. F. (1982). Plasticity in Reinforced Concrete, McGraw-Hill, New York,


USA.

Chen, W. F., Suzuki, H., and Chang, T. Y. P. (1980). “End Effects of Pressure
Resistant Concrete Shells.” Journal of the Structural Division, ASCE, 106(4),
751-771.

Chen, Y. Y. (2007). “Nonlinear Finite Element Analysis of Certain Shear Failure


Mechanisms in Reinforced Concrete Beams.” PhD thesis, Heriot-Watt University,
UK.

Crisfield, M. A. (1982). “Accelerated Solution Techniques and Concrete


Cracking.” Computer Methods in Applied Mechanics and Engineering, 33(1-3),
585-607.

Dahlblom, O. and Ottosen, N. (1990). “Smeared Crack Analysis Using


Generalized Fictitious Crack Model.” Journal of Engineering Mechanics, 116(1),
55–76.

Darwin, D. (1985). “Crack Propagation in Concrete-Study of Model Parameters.”


Seminar on Finite Element on Finite Element Analysis of Reinforced Concrete
Structure, Tokyo, Japan.

Darwin, D., and Pecknold, D. A. (1977). “Nonlinear Biaxial Stress-Strain Law for
Concrete.” Journal of Engineering Mechanics Division, ASCE, 103(EM2), 229-
241.

Eligehausen, R., Popov, E. P., and Bertero, V. V. (1983). “Local Bond Stress-Slip
Relationships of Deformed Bars under Generalized Excitations.” UCB/ EERC-83/
23, University of California, Berkeley, USA.

109
Evans, R. J., and Pister, K. S. (1966). “Constitutive Equations for a Class of
Nonlinear Elastic Solids.” Solids and Structures, 23(6), 751-767.

Feenstra, P. H. and Borst, D. R. (1995). “Constitutive Model for Reinforced


Concrete.” Journal of engineering mechanics, 121(5), 587–595.

Fields, K. and Bischoff, P. H. (2004). “Tension Stiffening and Cracking of High-


strength Reinforced Concrete Tension Members.” ACI Structural Journal, 101(4),
447-456.

Ganaba, T. H. (1985). “Nonlinear Finite Element Analysis of Plates and Slabs.”


PhD thesis, The University of Warwick, UK.

Gerstle, K. H. (1981). “Simple Formulation of Biaxial Concrete Behavior.” ACI


Journal, 78(1), 62-68.

Gilbert, R.I. and Warner, R.F. (1978). “Tension Stiffening in Reinforced Concrete
Slabs.” Journal of the structural division, 104(12), 1885–1900.

Guo, Z. H., and Zhang, X. Q. (1987). “Investigation of Complete Stress


Deformation Curves for Concrete in Tension.” ACI Materials Journal, 84(4), 278-
285.

Han, D. J., and Chen, W. F. (1987). “Constitutive Modeling in Analysis of


Concrete Structures.” Journal of Engineering Mechanics Division, ASCE, 113(4),
577-593.

Hand, F. R., Pecknold, D. A., and Schnobrich, W. C. (1973). “Nonlinear Layered


Analysis of RC Plates and Shells.” Journal of the Structural Division, ASCE,
99(ST7), 1491-1505.

Hanson, J. H., and Ingraffea, A. R. (2003). “Using Numerical Simulations to


Compare the Fracture Toughness Values for Concrete from the Size-effect, Two-
parameter and Fictitious Crack Models.” Engineering Fracture Mechanics, 70(7-
8), 1015-1027.
110
Hassan, A. A. A., Hossain, K. M. A., and Lachemi, M. (2008). “Behavior of Full-
scale Self-consolidating Concrete Beams in Shear.” Cement and Concrete
Composites, 30(7), 588-596.

Hatano, T., and Tsutmi, H. (1960). “Dynamical Compressive Deformation and


Failure of Concrete under Earthquake Load.” Proceedings, 2nd World Conf. on
Earthquake Engng., Tokyo, 1963-1978.

He, X. G. (1999). “Constitutive Modeling of Reinforced Concrete for Nonlinear


Finite Element Analysis.” PhD thesis, the University of Hong Kong, Hong Kong.

He, X. G., and Kwan, A. K. H. (2001). “Modeling Dowel Action of


Reinforcement Bars for Finite Element Analysis of Concrete Structures.”
Computers & Structures, 79(6), 595-604.

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, USA.

Hoshikuma, J., Kawashima, K., Nagaya, K., and Taylor, W. (1997). “Stress-strain
Model for Confined Reinforced Concrete in Bridge Piers.” Journal of Structural
Engineering, 123(5), 624-633.

Hsu, T. T. C., Slate, F. O., Sturman, G. M., and Winter, G. (1963).


“Microcracking of Plain Concrete and the Shape of the Stress Strain Curve.” ACI
Journal, 60(2), 209-224.

Hsu, T. T. C., Zhang, L. X. (1996). “Tension Stiffening in Reinforced Concrete


Membrane Elements.” ACI Structural Journal, 93(1), 108-115.

Hu, H. T., and Schnobrich, W. C. (1989). “Constitutive Modeling of Concrete by


using Nonassociated Plasticity.” Journal of Materials in Civil Engineering, ASCE,
1(4), 199-216.

111
Hughes, B. P., and Chapman, G. P. (1966). “The Deformation of Concrete and
Microconcrete in Compression and Tension with Particular Reference to
Aggregate Size.” Magazine of Concrete Research, 18(54), 19-24.

Kappos, A. J. (1991), “Analytical Prediction of the Collapse Earthquake for R/C


Buildings: Suggested Methodology.” Earthquake Engineering & Structural
Dynamics, 20(2), 167–176.

Kent, D. C., and Park, R. (1971). “Flexural Members with Confined Concrete.”
Journal of the Structural Division, ASCE, 97(7), 1969-1989.

Kong, F. K., Robins, P. J., Cole, D. F. (1971). “Web Reinforcement Effects on


Lightweight Concrete Deep Beams.” ACI, 68(7), 514-520.

Kupfer, H. B., and Gerstle, K. H. (1973). “Behavior of Concrete under Biaxial


Stresses.” Journal of Engineering Mechanics Division, ASCE, 99(4), 853-866.

Kupfer, H., Hilsdorf, H. K., and Rusch H. (1969). “Behavior of Concrete under
Biaxial Stresses.” ACI Journal, 66(8), 656-666.

Kwan, A. K. H., Ho, J. C. M., and Pam, H. J. (2002). “Flexural Strength and
Ductility of Reinforced Concrete Beams.” Structures and Buildings, 152(4), 361-
369.

Lin, C. S., and Scordelis, A. C. (1975). “Nonlinear Analysis of RC Shells of


General Form.” Journal of the Structural Division, ASCE, 101(3), 523-538.

Liu, G. Q. (1985). “Nonlinear and Transient Finite Element Analysis of General


Reinforced Plates and Shells.” PhD thesis, The University of Swansea, UK.

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(5), 1025-
1034.

112
Mainstone, R. J. (1975). “Properties of Materials at High Rates of Straining or
Loading.” Material and Structures (Paris), 8(44), 102-116.

Mander, J. B., Priestley,M. J. N., and Park, R. (1988a). “Theoretical Stress-


Strain Model for Confined Concrete.” Journal of Structural Engineering, ASCE,
114(8), 1804-1826.

Mander, J. B., Priestley, M. J. N., and Park, R. (1988b). “Observed Stress-Strain


Behavior of Confined Concrete.” Journal of Structural Engineering, ASCE, 114(8)
1827-1849.

Matthies, H. and Strang, G. (1979). “The Solution of Nonlinear Finite Element


Equations.” International Journal for Numerical Methods in Engineering, 14(11),
1613-1626.

Milford, R. V., Schnobrich, W. C. (1985). “The Application of the Rotating Crack


Model to the Analysis of Reinforced Concrete Shells.” Computers & Structures,
20(1-3), 225-234.

Naji, J. H. (1989). “Nonlinear Finite Element Analysis of Reinforced Concrete


Panels and Infilled Frames under Monotonic and Cyclic Loading.” PhD thesis,
The University of Bradford, UK.

Nelissen, L.J.M. (1972). “Biaxial testing of normal concrete.” Heron Netherlands,


18(1).

Ngo, D., and Scordelis, A. C. (1967). “Finite Element Analysis of Reinforced


Concrete Beams.” ACI Journal, 64(3), 152-163.

Nilson, A. H. (1968). “Non-Linear Analysis of Reinforced Concrete by the Finite


Element Method.” ACI Journal, 65(9), 757-766.

Ottosen, N. S. (1977). “A Failure Criterion for Concrete.” Journal of Engineering

Mechanics Division, ASCE, 114(11), 1890-1910.

113
Park, R., Priestley, M. J. N., and Gill, W. D. (1982). “Ductility of Square-
Confined Concrete Columns.” Journal of the Structural Division, ASCE, 108(4),
929-950.

Polak, M. A., and Blackwell, K. G. (1998). “Modeling Tension in Reinforced


Concrete Members Subjected to Bending and Axial Load.” Journal of Structural
Engineering, ASCE, 124(9), 1018-1024.

Rahman, H. H. A. (1982). “Computational Models for the Nonlinear Analysis of


Reinforced Concrete Flexural Slab Systems.” PhD thesis, University College of
Swansea, UK.

Raphael, J. M. (1984). “Tensile Strength of Concrete.” ACI Journal, 81(2), 158-


165.

Rashid, Y. R. (1968). “Ultimate Strength Analysis of Prestressed Concrete


Vessels.” Nuclear Engineering and Design, 7(4), 334-344.

Raveendran, S. (1988). “Modeling of Reinforced Concrete Beams Subjected to


Both Static and Dynamic Loading.” PhD thesis, The North East London
Polytechnic Barking, UK.

Richart, F. E., Brandtzaeg, A., and Brown, R. L. (1928). “A Study of the Failure
of Concrete under Combined Compressive Stresses.” University of Illinois
Engineering Experimental Station, Bulletin No. 185, Illinois, USA.

Riks, E. (1979). “An Incremental Approach to the Solution of Snapping and


Buckling Problems.” International Journal of Solids and Structures, 15(7), 529-
551.

Saenz, L. P. (1964). “Discussion of ‘Equation for Stress-strain Curve for


Concrete’ by Desayi and Krishnan.” ACI Journal, 61(9), 1229-1235.

114
Scanlon, A., (1971). “Time Dependent Deflections of Reinforced Concrete
Slabs.” PhD thesis, University of Alberta, Edmonton, Canada.

Scordelis, A. C. (1972). “Finite Element Analysis of Reinforced Concrete


Structures.” Specialty Conference on the Finite Element Method in Civil
Engineering, Canada.

Scott, B. D., Park, R., Priestley, K. J. N. (1982). “Stress-Strain Behavior of


Concrete Confined by Overlapping Hoops at Low and High Strain Rates.” ACI
Journal, 79(1), 13-27.

Scrivener, K.L., Bentur, A. and Pratt, P.L. (1988). “Quantitative Characterization


of the Transition Zone in High Strength Concretes.” Advances in Cement
Research, 1(4), 230–237.

Shayanfar, M. A., Kheyroddin, A., and Mirza, M. S. (1997). “Element Size


Effects in Nonlinear Analysis of Reinforced Concrete Members.” Computers &
Structures, 62(2), 339-352.

Sheikh, S. A., and Uzumeri, S. M. (1980). “Strength and Ductility of Tied


Concrete Columns.” Journal of the Structural Division, ASCE, 106(5), 1079-1102.

Singh, B., and Chintakindi, S. (2012). “An Appraisal of Dowel Action in


Reinforced Concrete Beams.” Proceedings of the ICE- Structures and Buildings,
166(5), 257-267.

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), 49-59.

Su, E. C. M., and Hsu, T. T. C. (1988). “Biaxial Compression Fatigue and


Discontinuity of Concrete.” ACI Material Journal, 85(3), 178-188.

Vecchio, F. J. (2000). “Analysis of Shear-critical Reinforced Concrete Beams.”


ACI Structural Journal, 97(1), 102–110.

115
Vellenas, J., Bertero, V. V., and Popov, E. P. (1977). “Concrete Confined by
Rectangular Hoops Subjected to Axial Loads.” Report 77/13, Earthquake
Engineering Research Center, University of California, Berkeley, USA.

Wu, Y. (2006). “Post-crack and Post-peak Behavior of Reinforced Concrete


Members by Nonlinear Finite Element Analysis.” PhD thesis, the University of
Hong Kong, Hong Kong.

Yin, W. S., Su, E. C. M., Mansur, M. A., and Hsu, T. T. C. (1989). “ Biaxial Tests
of Plain and Fiber Concrete.” ACI Material Journal, 86(3), 236-243.

Zhao, Z. Z., Kwan, A. K. H. and He, X. G. (2004). “Nonlinear Finite Element


Analysis of Deep Reinforced Concrete Coupling Beams.” Engineering structures,
26(1), 13–25.

Zhu, R. R., Hsu, T. T. and Lee, J. Y. (2001). “Rational Shear Modulus for
Smeared-crack Analysis of Reinforced Concrete.” ACI Structural Journal, 98(4),
443-450.

Zienkjewicz, O. C., and Cormeau, I. C. (1974). “Viscoplasticity-Plasticity and


Creep in Elastic Solids: A Unified Numerical Solution Approach.” International
Journal for Numerical Methods in Engineering, 8(4), 821-845.

116

S-ar putea să vă placă și