Вы находитесь на странице: 1из 266

SHRINKAGE, RESIDUAL STRESS, AND CRACKING IN HETEROGENEOUS

MATERIALS

A Thesis
Submitted to the Faculty
of
Purdue University
by
Jae Heum Moon

In Partial Fulfillment of the


Requirements for the Degree
of
Doctor of Philosophy

May 2006
Purdue University
West Lafayette, Indiana

ii

Dedicated to my family for their love and sacrifice

iii

ACKNOWLEDGEMENTS

I would like to express my deepest gratitude to Professor Jason Weiss for his
countless guidance, support, advice, and patience throughout the course of my research.
Without his supervision and advice, this research would not have been possible.
I would also like to thank Professors Menashi Cohen, John Haddock, and ChinTeh Sun for serving on my committee. The valuable suggestions and encouragement
given by them has helped me immensely.
I gratefully acknowledge the support received from the National Science
Foundation and Purdue University while pursuing this research work. This research was
conducted in the Material Sensing Laboratory and as such I gratefully acknowledge the
support that has made this work possible. Thanks are also due to Ms. Janet Lovell and
Mr. Mark Baker for the constant supports and help rendered by them in resolving various
issues and problems regarding laboratory works.
Very special thanks to Farshad Rajabipour, Bradly Pease, Jon Couch. This work
would not have been completed without them.

Any number of words would be

insufficient to express my appreciation to all my friends at Purdue, Aleksandra Radlinska,


Yogini Deshpande, Sulapha Peethamparan, and all my fellow students in Materials Area
of School of Civil Engineering for their help and company.

iv

Finally, I would like to express my sincere gratitude and love to my family


members for their constant support, encouragement and love, which has served as an
anchor for me. I will always treasure the sacrifices done by my parents for providing me
with every possible opportunity to excel, and the love and care of my sisters.

TABLE OF CONTENTS

Page
LIST OF TABLES............................................................................................................. ix
LIST OF FIGURES ........................................................................................................... .x
ABSTRACT.................................................................................................................... xvii
CHAPTER 1: INTRODUCTION ...................................................................................... 1
1.1 Background ................................................................................................................... 1
1.2 Proposed Research ........................................................................................................ 4
1.4 Organization of Contents ............................................................................................. 7
CHAPTER 2: AUTOGENOUS AND DRYING SHRINKAGE OF CONCRETE AND
THE RESIDUAL STRESS DEVELOPMENT IN THE RESTRAINED RING
SPECIMENS....................................................................................................................... 8
2.1 Introduction................................................................................................................... 8
2.2 Research Significance................................................................................................. 12
2.3 Analytical Approach ................................................................................................... 13
2.4 Moisture Gradients and Their Relationship with Shrinkage....................................... 15
2.5 Analytical Solution ..................................................................................................... 17
2.5.1 Stress Development in the Concrete Ring due to Differential Shrinkage:
,diff.shr.(r, ) .................................................................................................................... 17

2.5.2 Stress Development in the Concrete Ring due to Steel Ring Constraint: ,rest. ring(r)
...................................................................................................................................... 20
2.5.3 Superposition of Differential Shrinkage and Steel Ring Restraint ...................... 23

vi

Page................................................................................................................................Page
2.6 Finite Element Analysis (FEA)................................................................................... 25
2.7 Comparison of the Analytical Approach with FEA and Application to an Experiment
........................................................................................................................................... 26
2.7.1 Comparison with FEA ......................................................................................... 26
2.7.2 Application to Experiments ................................................................................. 33
2.8 Conclusions................................................................................................................. 38
CHAPTER 3: DEGREE OF RESTRAINT ..................................................................... 39
3.1 Introduction................................................................................................................ 39
3.2 Ring Specimen Geometry ........................................................................................... 42
3.3 Quantifying the Degree of Restraint under Uniform Shrinkage (Analytical Approach)
........................................................................................................................................... 44
3.4 Quantifying the Degree of Restraint in the Presence of Moisture Gradient ............... 52
3.5 A Practical Example of Geometry Selection for the Restrained Ring Test................ 62
3.6 Summary and Conclusions ......................................................................................... 66
CHAPTER 4: INTERNAL STRESS DEVELOPMENT IN HETEROGENEOUS
SYSTEMS......................................................................................................................... 68
4.1 Introduction................................................................................................................ 68
4.2 Effective Material Property Calculations................................................................... 70
4.3 Single Aggregate Prism Systems ............................................................................... 76
4.4 Multiple Aggregate Systems...................................................................................... 86
4.5. Summary................................................................................................................... 97
CHAPTER 5: ANALYSIS PROCEDURES FOR ASSESSING THE SHRINKAGE AND
CRACKING IN A CEMENTITIOUS COMPOSITE ON THE MESO-SCALE............. 99
5.1 Introduction................................................................................................................ 99
5.2 Overview of the Approach....................................................................................... 101
5.3 Analysis Procedures................................................................................................. 102
5.4 Preprocessing: Acquiring the Image from a Realistic Specimens ........................... 105
5.4.1 Specimen Preparation ........................................................................................ 105

vii

Page............................................................................................................................Page
5.4.2 Scanning............................................................................................................. 108
5.5 Finite Element Simulations...................................................................................... 115
5.5.1 General Approach .............................................................................................. 115
5.5.2 Failure Criterion................................................................................................. 116
5.5.3 Material Properties............................................................................................. 118
5.5.4 OOF Simulation (Performing Trials)................................................................. 119
5.6 Post-Processing: Analysis of Results....................................................................... 120
5.7 Results and Discussions............................................................................................ 122
5.7.1 Conventional Approach Using Effective Medium Assumption ........................ 123
5.7.2 OOF Simulation (Results) ................................................................................. 126
5.8 Summary................................................................................................................... 146
CHAPTER 6: PARAMETRIC STUDY OF SHRINKAGE CRACKING IN A
CONCRETE COMPOSITE............................................................................................ 147
6.1 Introduction............................................................................................................... 147
6.2 Importance of This Research .................................................................................... 148
6.3 One Directionally Restrained Boundary Condition.................................................. 149
6.3.1 Volume fraction of aggregate ............................................................................ 149
6.3.2 Size Distribution of Aggregate .......................................................................... 172
6.3.3 Bond Condition.................................................................................................. 180
6.4 Free Boundary Condition.......................................................................................... 190
6.4.1 Volume fraction of aggregate ............................................................................ 190
6.5 Discussion ................................................................................................................. 210
CHAPTER 7: SUMMARY............................................................................................. 212
7.1 Introduction............................................................................................................... 212
7.2 Quantification of the Residual Stress Development in Concrete Due to Non-Uniform
Shrinkage and the Degree of Restraint Exhibited by the Restrained Ring Test When
Different Specimen Geometries are Used....................................................................... 213

viii

Page................................................................................................................................Page
7.3 Discussion of the Comparison between OOF Simulation Results in a Heterogeneous
Material and Experimental Measurements ..................................................................... 216
7.4 Conclusion ................................................................................................................ 228
LIST OF REFERENCES................................................................................................ 230
APPENDIX..................................................................................................................... 244
VITA ............................................................................................................................... 245

ix

LIST OF TABLES

Table

Page

5.1

Material properties of cement paste.. 118

5.2

Material properties of aggregate....... 118

6.1

Information on the volume of aggregate in each composite shown in


Figure 6.11........ 174

LIST OF FIGURES

Figure

Page

1.1 The influence of aggregate on shrinkage and microcracking...

2.1 Conceptual illustration of the restraint components and a sample of stress


gradient in the concrete ring.....

10

2.2 Relative humidity (RH) gradients calculated using Eq. 2.3 (RHI=100%,
RHS=50%, thickness=75mm) ......

14

2.3 Stress gradient due to self-restraint in the concrete ring (i.e., without the
steel ring (Eq. 2.10)) (ROC=0.225m, RIC=0.150m, Econ=21GPa,
SH-const= -100 ).....

20

2.4 Conceptual illustration of the shrink-fit approach.

21

2.5 Stress gradient due to steel ring constraint in the concrete ring (Eq. 2.12)
(ROC=0.225m,
RIC=ROS=0.150m,
RIS=0.1406m,
Econ=21GPa,
Esteel=200GPa, SH-const= -100 ).

22

2.6 Stress gradient in the concrete ring exposed to circumferential drying


with steel ring (Eq. 2.13) (ROC=0.225m, RIC=ROS=0.150m, RIS=0.1406m,
Econ=21GPa, Esteel=200GPa, SH-const= -100 )

24

2.7 Comparison of stress gradients due to self-restraint in the concrete ring


without steel ring (FEA and Eq. 2.10) (ROC=0.225m, RIC=0.150m,
Econ=21GPa, SH-const= -100 )

27

2.8 Comparison of stress gradients due to self-restraint in the concrete ring


without steel ring (FEA and Eq. 2.10) (=0.002) (ROC=0.225m,
RIC=0.150m, Econ=21GPa, SH-const= -100 )..

29

xi

Figure

Page

2.9 Comparison of stress gradients due to self-restraint in the concrete ring


without steel ring (FEA and Eq. 2.10) (=0.02)
(ROC=0.225m,
RIC=0.150m, Econ=21GPa, SH-const= -100 )..

30

2.10 Comparison of Actual-Max due to self-restraint in the concrete ring without


steel ring (ROC=0.225m, RIC=0.150m, RIS=0.1406m, Econ=21GPa, SHconst= -100 )...

31

2.11 Comparison of stress gradients due to steel ring restraint in the concrete
ring (FEA and Eq. 2.12) (ROC=0.225m, RIC=ROS=0.150m, RIS=0.1406m,
Econ=21GPa, Esteel=200GPa, SH-const= -100 )...

32

2.12 Comparison of stress gradients due to circumferential drying in a


concrete ring with steel ring
(ROC=0.225m, RIC=ROS=0.150m,
RIS=0.1406m, Econ=21GPa, Esteel=200GPa, SH-const= -100 )

33

2.13 Stress gradients due to circumferential drying in the concrete ring with
steel ring (calculated using Eq. 2.13 and real test data) (0.50 W/C, 50%
Sand, ROC=0.225m, RIC=ROS=0.150m, RIS=0.1406m)

36

2.14 Predicted depth of microcracking in the concrete ring (0.50 W/C, 50%
sand, ROC=0.225m, RIC=ROS=0.150m, RIS=0.1406m).

37

3.1 Drying direction and the resulting stress development of restrained ring

42

3.2 A conceptual illustration of the degree of restraint..

45

3.3 The influence of specimen geometry on the degree of restraint and the
ratio st/SH...

50

3.4 Comparison between the analytical solution and the experimentally


determined degree of restraint before the through cracking (top and
bottom drying)..

51

3.5 The simulated humidity profiles corresponding to different values..

55

3.6 Degree of restraint as drying progresses; obtained from finite


element simulations (top and bottom drying)...

57

3.7 Influence of shrinkage condition on the deformation of concrete and steel


rings..

58

xii

Figure

Page

3.8 The effect of non-uniform deformation of the steel ring (simulations


results)..

60

3.9 Comparison on the degree of restraint obtained from the analytical


procedure (Eq. 6, uniform drying) and the values obtained from finite
element simulations (non-uniform drying)...

62

3.10 Graphical representation of Eq. 3.5 and 3.6 in terms of possible


combinations of geometry parameters for a desired degree of restraint..

65

4.1 Shrinkage exponent n as a function of the aggregate to paste stiffness


ratio...

74

4.2 Acoustic activity in mortars at early ages (50% aggregate volume)

74

4.3 An illustration of the two geometries simulated..

77

4.4 Stress development in an (a) unrestrained single aggregate specimen and


b) the externally restrained single aggregate specimen

80

4.5 Stress localization for different boundary conditions...

81

4.6 Stress development along the Y-axis of a restrained prism specimen over
an aggregate (restrained boundary condition) .

83

4.7 Experimental data from a restrained specimen containing an


Instrumented Aggregate....

85

4.8 Multiple aggregate system composed of unit cell matrix.

88

4.9 Comparisons of equivalent elastic modulus of composites..

89

4.10 Maximum stress development in an externally unrestrained composite..

91

4.11 Maximum stress development in a composite externally restrained in one


direction

93

4.12 Cracking tendency of a composite (horizontally restrained boundary


conditions) ...

96

5.1 Flow chart of OOF simulation on the meso-scale

102

5.2 Specimen preparation (surface treatments) .

107

xiii

Figure

Page

5.3 Flat-bed color scanner and concrete samples (saw-cut, polished, and
stained).

110

5.4 Image analysis (re-colorization, part of the scanned image Vagg = 55.7
%).

112

5.5 Meshed sample.

114

5.6 Failure criteria..

117

5.7 Color scaled image of stress distribution in a composite (Paste-1,


Horizontally restrained, paste= -140 )...

121

5.8 Cracked image (paste-1, horizontally-restrained, paste=140 )..

122

5.9 Equivalent elastic modulus of a composite (OOF simulation vs. Hansens


model)...

124

5.10 Equivalent stress vs. shrinkage strain of paste.

125

5.11 Shrinkage strain of paste vs. load (horizontally-restrained boundary


condition)..

127

5.12 Cracks in a composite (one directionally restrained boundary condition)...

129

5.13 Cracks in a composite (free boundary condition)

130

5.14 Color contour image of stress distribution in a composite (One


directionally restrained boundary condition)...

132

5.15 Cumulative area fraction vs. stress...

134

5.16 Strain energy density versus area fraction

136

5.17 Strain energy density level versus normalized stored energy..

138

5.18 Strain energy density level vs. cumulative values of normalized stored
energy...

140

5.19 Shrinkage strain of paste vs. cumulation of normalized stored energy

142

xiv

Figure

Page

5.20 Relationships among the percentage of the released energy, area fraction
of cracked zones, and the magnitude of shrinkage strain of paste (free
(free boundary condition).

144

5.21 Relationship between the area fraction of cracked zones and the
percentage of the released energy (free boundary condition).

145

6.1 Equivalent elastic modulus of a composite with the volume fraction of


aggregate..

150

6.2 Equivalent shrinkage strain of a concrete as a function of the volume


fraction of aggregate.

151

6.3 Equivalent stress of a concrete (paste = 225 )...

152

6.4 Series law for the calculation of the tensile strength of a concrete
composite..

153

6.5 Strength and equivalent stress of a composite with the variation of the
volume fraction of aggregate (paste 1 in Table 5.1)..

154

6.6 Cracking potential of composites based on the volume fraction of


aggregate and material properties.

156

6.7 Images with different volume fraction of aggregate

158

6.8 Images of composites with cracks (one directionally restrained boundary


condition)..

162

6.9 Average sectional stresses of composites vs. (a) shrinkage strain of


mortar and (b) shrinkage strain of paste (one directionally restrained
boundary condition).

167

6.10 a Stored strain energy vs. shrinkage strain of mortar..

169

6.10 b Stored strain energy vs. shrinkage strain of paste

170

6.10 c Energy release at failure vs. volume fraction of aggregate..

170

6.10 d Area fraction of cracked zone vs. percent-released energy..

171

6.10 e

171

Acoustic energy development as a function of time (Chariton et al., 2002)

xv

Figure

Page

6.11 Composites with different number of aggregates maintaining the volume


fraction of aggregate.

173

6.12 Stress development in composites

175

6.13 Sectional stress development in a composite under the restrained


boundary condition...
6.14 The sectional ratio of the total chord length of the aggregates to the
height of the concrete...

176
178

6.15 Images of cracked composites at failure (one directionally restrained


boundary condition).

182

6.16 Cracking behavior of a mortar composite (Vagg = 55.7 %) with different


bond conditions (one directionally restrained boundary condition).

185

6.17 Image of the cracked mortar composite (Vagg = 55.7 %) at failure (paste =
-190 ) (one directionally restrained boundary condition with perfectly
plastic failure criterion for bond phase) ..

187

6.18 Cracking behavior of a mortar composite with different failure criterion


on the bond phase (one directionally restrained boundary condition...

188

6.19 Images of composites with cracks (free boundary condition - paste = -250
).

190

6.20 Relationship between the area fraction of cracked zones and the
shrinkage strain of paste (free boundary condition).

194

6.21 Cracking behavior of mortar composites (free boundary condition)...

195

6.22 Relationship between the area fraction of cracked zones and the energy
release (Free boundary condition)

197

6.23 Images of cracked composites at paste = -250 (free boundary


condition)..
6.24 Volume fraction of cracked zones depending on the ratio of bond
material properties to paste material properties...
6.25 Cracking behavior of a mortar composite (Vagg = 55.7 %) with different
bond conditions (free boundary condition)..

200
202

203

xvi

Figure

Page

6.26 Relationship between the area fraction of cracked zones and the percent
released energy (bond phase, free boundary condition)...

204

6.27 Cracked image of a mortar composite (Vagg = 55.7 %) (paste = -250 )


(free boundary condition with perfectly plastic failure criterion for bond
phase)

205

6.28 Area fraction of cracked zones vs. shrinkage strain of paste...

206

6.29 Cracking behavior of a mortar composite with different failure criterion


on the bond phase (free boundary condition)...

207

6.30 Relationship between the area fraction of cracked zones and the percent
released energy.

208

7.1 A comparison of the shrinkage strain of mortar and the shrinkage strain
of paste as (OOF simulation, elastic)...

217

7.2 Relationship between the shrinkage strain of mortar and the shrinkage
strain of paste (OOF simulation, free boundary condition)..

218

7.3 Energy release and Area fraction of cracked zones versus shrinkage strain
of mortar and paste phase (OOF simulation, free boundary condition)...

219

7.4 Restrained (a) and free (b) shrinkage specimen geometries with locations
of acoustic emission sensors (Pease et al., 2004).

222

7.5 Autogenous shrinkage of a paste specimen and a mortar specimen, and


the dynamic elastic modulus of paste (w/c = 0.3)

224

7.6 Cumulative acoustic energies of mortar specimens (free boundary


condition)..

226

xvii

ABSTRACT

Moon, Jae Heum, Ph.D., Purdue University, May 2006. Shrinkage, Restraint, Residual
Stress, and Cracking in Heterogeneous Materials. Major Professor: W. Jason Weiss.

Concrete experiences volumetric changes as a result of material formation


(cement hydration), thermal variations, or moisture losses. If these volume changes are
prevented by the structure surrounding the concrete or volumetrically stable phases inside
the concrete, residual stresses can develop. In many cases, these residual stresses may be
large enough to result in cracking.

While the premature cracking of concrete is

significant enough to be considered in the design of concrete facilities, existing test


methods and design methodologies need to be updated to better quantify the cracking
potential especially when the concrete experiences non-uniform deformations.
Most design and testing approaches assume that concrete behaves like a
homogeneous material. However, concrete is a composite that consists of cement paste
and aggregates that have dissimilar material properties. When concrete changes its
volume, localized internal residual stresses develop due to heterogeneity (i.e., paste and
aggregates) that could lead to microcracking and premature cracking of concrete.
Therefore, to properly evaluate the cracking behavior of concrete, an analytical tool
considering the heterogeneous nature of concrete should be developed.

xviii

This research begins with quantifying the impact of non-uniform shrinkage on the
residual stress development in the restrained ring test considering that the concrete
experiences non-uniform moisture loss (drying). The role of the boundary conditions and
the degree of restraint on residual stress development is also discussed. Next, the internal
residual stresses that develop in a multi-phase composite system are examined by varying
the series of parameters (material properties of each-phase, volume fraction of aggregate,
and bond conditions between the phases). Analytical modeling is used to assess the
microcracking and cracking behavior of concrete composite systems using the objectoriented finite element code.
It is the hypothesis of this research that the cracking behavior of concrete can be
properly evaluated by assuming that concrete is a heterogeneous multi-phase composite
with the assistance of a recently developed object-oriented finite element code that
enables meshing a complex material meso-structure using optical mesostructural images.
Although this research focuses on shrinkage in cementitious composites only, this
software could have numerous applications in determining damage development in other
composites as well.

CHAPTER 1: INTRODUCTION

1.1 Background
Cementitious materials change volume in response to moisture variations,
temperature variations, and chemical reactions. If these volume changes are restrained,
residual stresses can develop and could lead to microcracking and cracking.
Several models exist to estimate shrinkage in mature concrete sections (Bazant
and Panula, 1979; Gardner and Lockman, 2002; and ACI 209 under revision). While
these models may be useful for predicting long-term maximum shrinkage for design
codes, they do not work well in predicting the rates of shrinkage that are important for
performance models, especially when these models are focused on describing early-age
behavior. Recently developed theoretical models describe the risk of cracking associated
with the restraint of autogenous, drying, and thermal shrinkage (Weiss et al., 1999; Yang
et al., 1999), however, these models are based on macroscopic properties and require the
direct input of measured shrinkage values.
Although many improvements have been made in concrete technology over the
last several decades, it is still common to observe cracking in concrete soon after it is
placed (Weiss et al., 1997; Burrows, 1999; Weiss and Shah, 2002; RILEM, 2002). This
premature microcracking and cracking is mainly caused by excessive volume changes at
an early age when the concrete does not have sufficient strength to resist these changes.

The heterogeneity of concrete results in internal residual stresses even if the net section
force is low.
Few approaches exist to explicitly quantify the role of each of the constituent
phases (paste, transition zone or bond, and aggregate) in contributing directly to
shrinkage and cracking. If such an approach exists it could be useful for performance
prediction and mixture optimization models.

Historically, the contribution of the

aggregate phase has only been considered through a series of approximate equations that
result in the shrinkage of concrete being considered as the product of the free shrinkage
of a pure paste phase and the volume fraction of the paste phase raised to an exponent
(usually between 1.2 and 1.7) (Pickett, 1956; Hobbs, 1974). While this type of approach
generally works well to describe long-term shrinkage, it does not describe local stresses
and microcracking that may develop inside the concrete. The approximate method for
predicting shrinkage described above does not account for changes in the relative
stiffness between the paste and the aggregates, aggregate size or shape, or damage
development (microcracking), which of itself is ultimately important for use in transport
models or through-section cracking predictions. In other words, previous models provide
averaged (equivalent) residual stress calculations using the average material properties of
a composite that would have higher internal residual stresses due to the stress
concentration induced by heterogeneity.
For example, Figure 1.1 (Pease et al., 2003b) illustrates the shrinkage strains and
the passive acoustic energy measured in paste and mortar systems. It can be seen that the
shrinkage strain of a mortar specimen with 45 percent aggregate (by volume) has lower
shrinkage strain than a paste specimen (Figure 1.1 (a)). It can also be seen in Figure 1.1

(b) that, with closely spaced aggregates (45% by volume), the microcracking is more
constant over time (i.e., a smooth curve). However, when the aggregates are spaced
further apart (15 percent by volume), discrete acoustic activities (sudden jumps in
acoustic energy) are observed, which is consistent with the idea that cracks propagate
through the paste between the aggregates.

However, it has been observed that the

cumulative acoustic energy does not linearly increase by increasing the volume fraction
of aggregate, regardless of the boundary condition as shown in Figure 1.1 (c). This result
implies that each phase (paste, aggregate, and bond) has an important role in the

Free Shrinkage ()

-100
-200
-300
-400
-500
-600
Free Shrinkage
Paste
45% Agg

-700
-800
0

8 12 16 2 0 24

Age of Specimen (Hours)

(a)

2.5
1.6

3
15% Aggregate
45% Aggregate

1.2

1.5

0.8
1

0.4
0.5

00

C umulative A coustic Energy


@ 24 H ours (nVs)

Cumulative
A c o u s Acoustic
t i c E n e r g Energy
y ( V s ) (nVs)

microcracking and cracking behaviors of concrete.

Restrained
Free

0
00

4
4

88 12
12 1616 202024 24

Age of
(Hours)
AgeSpecimen
of Specimen (Hours)

(b)

0
20
40
60
Aggregate Content (%)

(c)

Figure 1.1 The influence of aggregate on shrinkage and microcracking: (a) shrinkage
strain versus age, (b) cumulative acoustic energy versus age (restrained boundary
condition), and (c) cumulative acoustic energy (i.e., microcracking of damage) versus
volume fraction of aggregate

1.2 Proposed Research


To resolve the previously mentioned problems the restrained ring test will be
further investigated in this research. This research begins by considering the uniform
(autogenous) and non-uniform (drying) shrinkage behavior of concrete. This work begins
considering the concrete as a homogeneous material so that the effects of uniform and
non-uniform shrinkage can be assumed as it relates to the residual stress development in
the restrained ring test. Analytical solutions are developed for the calculation of the
residual stresses that are induced by non-uniform shrinkage. A method will also be
provided to quantifying the degree of restraint in the restrained ring test for specimens
with different geometries.
To better understand how mixture proportions influence the potential for
premature cracking in concrete, this work investigates the use of new simulation
techniques. A model was used that enabled the concrete to be considered as a multi
phase composite systems. This enabled the paste and aggregate phases to be explicitly
considered thereby enabling the stress concentrations around the aggregates to be
quantified. A recently developed object-oriented finite element code (OOF) (Fuller et al.
2003) is used in this research. OOF provides a user interface which enables the use of
actual 2-D images of concrete to be rapidly input in a finite element model. A procedure
was developed for taking an image of the cross-section of the concrete, developing a
finite element model for this composite computing the residual stresses that develop as a
result of volume change, and implementing a cracking criterion to describe damage
development.

This thesis describes methodologies to quantify residual stress development in


concrete when it experiences uniform and non-uniform shrinkage.

In addition to

assuming concrete to be a homogeneous material, this research shows how optical


microscopy could be combined with analytical tools (FEM) for predicting the stress fields
in a heterogeneous medium.

1.3 Objectives of the Study


The objectives of this thesis are to:

Provide a theoretical solution for quantifying the equivalent residual stress


development in concrete when it experiences non-uniform shrinkage in a ring
geometry due to moisture gradient.

Provide a theoretical solution to combine the equivalent residual stress


development for the specimen with non-uniform shrinkage and the residual stress
that develops from the boundary condition (degree of restraint) for the restrained
ring geometry.

Investigate the importance of heterogeneity on meso-structure scale as it relates to


the development of residual stress concentrations, and cracking in concrete
composite systems.

Evaluate the role of each parameter (volume fraction of each phase, material
properties of each phase, geometry of each phase, and bond condition on stress
development and cracking potential in concrete composites.

1.4 Organization of Contents


Chapter 2 provides a solution for the calculation of the equivalent residual stress in
a restrained ring test when it experiences non-uniform shrinkage due to moisture
gradients.
Chapter 3 provides a solution for the calculation of the degree of restraint in a
restrained ring test. This solution enables the degree of restraint to be quantified and used
in the selection of a test specimen geometry that best simulates a field structure.
Chapter 4 describes the concept of stress localization and concentration in a
composite system due to the existences of material phases that have different material
properties. This chapter describes the fundamental behavior of the composite system.
Chapter 5 describes the procedure of using new simulation techniques, to assess the
complex geometries that may be expected in an actual composite phase distribution. This
chapter also discusses the analysis procedures for OOF simulation results.
Chapter 6 discusses the parameters that affect the microcracking and cracking in a
composite.

The role of each parameter (volume fraction of each phase, material

properties of each phase, size and shape of the aggregates in the aggregate phases, and
the bonding condition between the aggregate and paste will be discussed.
Chapter 7 contains the summary and conclusions of this research.

CHAPTER 2: AUTOGENOUS AND DRYING SHRINKAGE OF CONCRETE AND


THE RESIDUAL STRESS DEVELOPMENT IN THE RESTRAINED RING
SPECIMENS

2.1 Introduction
Cementitious materials change volume in response to moisture variation,
temperature variation, and chemical reaction. If restrained, these volumetric changes
result in residual stress development that can lead to cracking. The restrained ring test
has recently become a popular method to assess a mixtures susceptibility to restrained
shrinkage cracking (Krause et al., 1995; AASHTO PP 34-99; Shah et al., 1992; Berke et
al., 1997; Grzybowski 1989; Lim et al., 1999). Over the last 80 years, researchers have
considered various adaptations of the ring test. For example, early work by Carlson and
Reading (1988) used the ring to qualitatively compare the shrinkage cracking potentials
of various cement compositions.

Malhotra and Zoldners (1967) proposed that a

pressurized ring test could be performed as a potential method for assessing the tensile
strength of concrete. Swamy and Stavrides (1979) suggested a ring test where strain
could be measured for assessing the behavior of fiber reinforced concrete. Grzybowski
and Shah (1989) used the ring test to investigate strain development in plain and fiber
reinforced concrete using strain gages on the concrete surface for the calibration of a
modeling approach. Kovler et al. (1993) combined the passive restraint from the classic
ring test with the active approach advocated by Malhotra for the development of a test
with an inner core ring that was made from a material with a higher thermal expansion

coefficient. In this test Kovler used the ring to apply passive restraint until a specific
time at which a tensile load was applied to the concrete ring by introducing a temperature
rise in the specimen assembly.

Kovlers approach enabled the testing time to be

shortened as the temperature rise that was required to cause fracture was related to the
potential additional stress capacity of the material. Weiss and Shah (2002) used the ring
geometry to demonstrate the benefit of a fracture mechanics approach for predicting the
geometry-dependent failure of the restrained concrete ring. More recently, in an effort to
increase the severity of the test, He et al. (2004) proposed an elliptical ring to increase the
stress concentration provided by the ring.
While the aforementioned research shows a wide range of applications for the
ring approach, this thesis will build on recent work to use measurements from an
instrumented ring test to make the test more quantitative. Attiogbe et al. developed
expressions based on a thin ring (See et al., 2003a; Attiogbe et al., 2003; See et al.,
2003b), while Weiss et al. (Weiss and Ferguson 2001; Hossain and Weiss 2003)
proposed a solution for a thick-walled concrete ring under uniform radial drying. This
work enabled the strain measured in a steel ring to be used to determine the maximum
residual tensile stress in the concrete ring using Eq. 2.1.

Actual Max = Steel ( t ) E s

2
2
2
R OS
+ R OC
R OS
R 2IS

2
2
2
R OC
R OS
2 R OS

Eq. 2.1

where ROS is the outer radius of steel ring, RIS is the inner radius of steel ring, ROC is the
outer radius of concrete ring, and ES is the elastic modulus of the steel ring. It should be
noted, however, that although steel was used in this expression, the restraining ring can
be made of any material provided it remains linear and elastic during the test. It has been

10

subsequently shown that this solution converges with the thin-ring solution for a
sufficiently small specimen (Attiogbe et al., 2004). Additionally, it was shown that the
combination of an elastic solution with Eq. 2.1 can provide an assessment of creep in the
ring (Hossain et al., 2003).
While Eq. 2.1 is appropriate for the case of uniform drying along the radial
direction, Figure 2.1 illustrates how a moisture gradient develops due to circumferential
drying, which can significantly influence the stress distribution (Weiss and Shah, 2002;
Bazant and Chern, 1985; Kim and Weiss, 2003; Weiss and Shah, 2001).

Drying from the outer


circumferential surface

Stress gradient due to


humidity gradient

Stress gradient due to


steel ring restraint

Stress gradient

(a) Total Stress Development

(b) Self Restraint

(c) Ring Restraint

Figure 2.1 Conceptual illustration of the restraint components and a sample of stress
gradient in the concrete ring

Moisture gradients were described as early as the 1930s by Carlson, who began
looking at the role of specimen thickness on drying shrinkage (Carlson, 1937). Bazant
and Najjar (1971) suggested that moisture diffusion analysis could be made more
accurate with the use of a diffusion coefficient that was non-linear with moisture content.

11

Bazant and Chern (1985) developed a general procedure for considering moisture
gradients, stress development, and microcracking. Grzybowski and Shah (1989a and
1989b) developed a model for restrained shrinkage cracking that considered moisture
gradients in the ring analysis that used a linear specimen approximation and a continuum
damage model, while Weiss and Shah (2001) used the moisture profile to illustrate the
specimen size dependency on the age of cracking. They described how the moisture
gradient resulted in complex stress fields that change shape over time (Weiss and Shah,
2001; Weiss, 1999). To quantify the moisture field that develops, Schiel et al. (2000)
suggested a procedure based on electrical impedance.

Grasley and Lange (2004)

demonstrated a novel sensor which can be used to measure the humidity profile directly.
Once the humidity profile is known, it can be used to compute the internal residual stress
in a specimen (Weiss and Shah, 2002b; Grasley and Lange, 2004; Bazant, 1986). The
existence of moisture gradients has been shown to result in the development of surface
microcracking based on analytical modeling (Bazant and Chern, 1985; Granger et al.,
1997), acoustic emission measurements (Weiss and Shah, 2001c), and optical microscopy
(Bisschop and Van Mier, 2001). While this research has pointed toward the importance
of moisture gradients in shrinkage and cracking studies, little has been developed in
understanding how moisture gradients influence the results that are obtained from the
ring test.
The residual stress development and cracking in the restrained ring depends on
the drying conditions. Stresses in the thick wall under uniform drying are highest at the
inner radius and the stresses decay as a function of r2 (Dally and Riley, 1991), while the
ring that dries from the outer circumference shows that the maximum actual residual

12

stress occurs at the outer radius (at least for short drying times) (Weiss and Shah, 2001c).
This difference in the shape of the stress field is important since it changes how the crack
initiates, develops, and propagates. Acoustic emission has been used to show that in
rings that dry uniformly along the radius, the cracking begins at the inner circumference
of the concrete rings while in specimens that dry from the outer circumference of the
rings, the cracks begin at the outer edge and propagate toward the center (Kim and Weiss,
2003; Hossain and Weiss, 2005). Analytical methods are needed to better characterize
the stress field for modeling failure in the ring. This thesis outlines a unique approach for
determining this stress field by considering the moisture gradient and the steel ring
restraint in the analysis of rings that dry from the outer circumference.

2.2 Research Significance


This chapter will propose a solution for determining the residual stress using the
restrained ring test geometry in which the residual stress that develops in the ring is
thought to be due to a combination of the self-restraint that develops due to non-uniform
drying conditions and the restraint that is provided by the steel ring. Based on this
approach, an analytical equation is presented for computing the residual stress field that
develops in the restrained ring. This expression can be used to describe stress gradients
that develop in a concrete ring. The development of this equation can enable the use of a
simple economic test (which has been proposed by AASHTO (AASHTO PP 34-99) and,
more recently, by ASTM (ASTM C 1581-04)), to provide information about the role of
moisture gradients in the shrinkage cracking behavior of concrete. In addition to use for
standard tests, this approach has implications in predicting the cracking behavior of slabs

13

and bridge decks in which the stress development is due to the concurrent effects of the
self-restraint and the restraint by the boundary conditions. It is fully anticipated that the
approach presented can be extended for general use in quantifying the cracking potential
of concrete under non-uniform drying.

2.3 Analytical Approach


It can be argued that the stress that develops in the restrained ring due to
circumferential drying can be considered as a combination of two components as shown
in Figure 2.1 (i.e., self-restraint due to non-uniform drying and restraint that comes from
external restraint supplied by the ring). Therefore, if the residual stress development can
be computed for each restraint component, all of the stress fields can be obtained.
An example of how the moisture gradient varies in concrete as a function of
drying time is shown in Figure 2.2 ( is the product of the moisture diffusion coefficient
and the drying time and is explained in detail in the following section) (Moon et al.,
2004). It can be noticed that for short drying times (i.e., low values of ), the moisture
gradient is very severe, with the majority of the drying taking place along the outer edge
of the specimen. As drying time increases (i.e., increases), the depth of the drying front
progresses further into the concrete and the moisture distribution is more gradual along
the radial direction.

14

100

RH (%)

90

80

Concrete

70

=0.002
=0.004
=0.008
=0.02
=0.04
=0.08
=0.2
=0.5
=100

60

50
0

0.02

0.04

0.06

0.08

Distance from the outer surface (m)

Figure 2.2 Relative humidity (RH) gradients calculated using Eq. 2.3
(RHI=100%, RHS=50%, thickness=75mm)

It should be noted that the loss of moisture causes the specimen to shrink. If the
concrete ring is unrestrained (i.e., no steel ring), it will shrink in the radial direction.
However, when the steel ring is present, the steel ring limits this shrinkage and residual
stresses develop (,rest.-ring). This restraint causes a pressure to develop at the interface
between the steel and concrete which can be calculated directly since the strain at the
inner surface of the steel ring (i.e., strain steel) is directly measured (Hossain and Weiss,
2003a; Shah and Weiss, 2003). In addition to the restraint from the steel, residual stress
develops due to self-restraint caused by circumferential drying (,diff.-shr.). The total

15

stress development can be expressed as the sum of the stress due to restraint and the
differential shrinkage.

( r , ) = ,rest . ring + ,diff .shr .

Eq. 2.2

The analytical solution of each component of this expression is further described


in section 2.5, however in order to enable a better understanding of the foundation of the
differential shrinkage behavior due to the moisture gradient, it is first necessary to
describe the moisture gradient and its relationship to drying shrinkage in Section 2.4.

2.4 Moisture Gradients and Their Relationship with Shrinkage


The analysis described in this thesis considers a ring that is exposed to drying
from the outer circumference. The surface that is exposed to drying shrinks more rapidly
than the internal concrete, which loses water more slowly. Early work on moisture
gradients by Carlson in the 1930s considered the moisture gradient to be explained with a
diffusion-controlled process similar to that used in thermal analysis (Carlson, 1937).
Bazant and Najjar (1971) proposed that moisture loss in concrete can be better expressed
using a nonlinear form of Ficks second law. They proposed that in addition to being
age-dependent the diffusion coefficient was non-linear with respect to moisture (i.e., pore
pressure, relative humidity, or degree of saturation).
While the non-linear approach is more accurate, especially for long term drying,
this work will utilize a linear diffusion function (in the form of Eq. 2.3), which was fit to
physical tests to estimate the humidity profiles as shown in Figure 2.2 (Moon et al.,
2004). The linear diffusion approach is preferred here since it enables the use of the

16

closed-form solution for determining the relative humidity (RH) distribution which is
amenable to hand calculation. It should be noted, however, that a more general solution
will be pursued in future work. As a result, a simple moisture gradient can be determined
using Eq. 2.3.

x
RH(x, t) = RH I (RH I RH S ) erfc

2 D t

Eq. 2.3

where, RH(x,t) is the relative humidity at a depth x from the drying surface, t is the
drying time, erfc is the complementary error function, RHS is the relative humidity at the
surface of the specimen, RHI is the internal relative humidity, and D is the aging moisture
diffusion coefficient of concrete. To simplify the combined treatment of the diffusion
coefficient and drying time here, a single parameter () was introduced as twice the
square root of the product of the diffusion coefficient and time ( = 2 Dt ).
The internal relative humidity (i.e., pore pressure) could be used to predict the
induced drying shrinkage strain (Bazant, 1986). The relationship between shrinkage and
relative humidity can be approximated for the case of high relative humidity (i.e., RH >
50%) as linearly proportional (Weiss and Shah, 2002c). As a result, the drying shrinkage
strain can be expressed by a constant free shrinkage coefficient (SH-Const) and the change
of relative humidity RH as illustrated in Eq. 4:

( t ) = SH const RH

Eq. 2.4

The constant free shrinkage coefficient (SH-Const) can be considered as the slope
of the shrinkage versus the change in the relative humidity relationship.

The change of

17

relative humidity ( RH) represents the difference between 100% RH and the internal RH
of a concrete specimen with time.

2.5 Analytical Solution


As previously described, the stress development in the concrete ring is mainly due
to two components; self-restraint due to differential shrinkage caused by moisture
gradients and external restraint due to the effect of the steel ring. In this section, an
analytical solution for calculating each component is presented. It should be noted that in
its initial form of this solution it is assumed that the concrete ring specimen exhibits
linear creep (relaxation) without microcracking (Bazant and Chern, 1985; Kim and
Weiss, 2003; Bisschop and Van Mier, 2001; Granger et al., 1997).

2.5.1 Stress Development in the Concrete Ring due to Differential Shrinkage: ,diff.shr.(r, )
The equation for calculating stress development in the concrete ring due to
differential shrinkage can be obtained by using an approach that is similar to the approach
used for differential thermal effects (Boley and Weiner, 1988). It should be remembered
that the analysis of the effect of self-restraint does not consider external restraint by the
steel ring. Plane stress is assumed in the derivation and, due to the axi-symmetric
geometry, the shearing stress at the interface is zero. Using Hookes law and the drying
shrinkage strain from Eq. 4, the stress components (i.e., stress in radial direction: rr and
stress in circumferential direction: ) can be expressed in radial components as:

18

u
du

dr + r (1 + ) SH const RH

rr =

E con
1 2

E con u
du

+
(1 + ) SHconst RH
2
dr
1 r

Eq. 2.5

Eq. 2.6

where, u is the radial displacement ( rr = du / dr , = u / r ), r is the radial distance in the


cylindrical coordinates system, and Econ is the effective elastic modulus of concrete (note:
the effective modulus can be used to consider the effects of creep). Substituting Eq. 2.5
and 2.6 into the stress equilibrium equation (

rr / r + (rr ) / r = 0 )

results in the following

expression;

d 1 d (r u )
d (RH )
= SH const (1 + )

dr r dr
dr

Eq. 2.7

This expression can be solved to yield the general solution for the hollow ring geometry:

u (r ) =

(1 + ) SH const
r

R IC

RH rdr + C 1 r +

C2
r

Eq. 2.8

where C1 and C2 are constants of integration. Application of the boundary conditions


(i.e., no traction on the inner or outer surface of the ring) enables the constants of
integration (C1 and C2) to be obtained. Combining Eq. 2.5, 2.6 and 2.8 enables the
residual stress (

, diff . shr . ( r , )

) in the concrete ring caused by differential shrinkage to be

computed using Eq. 2.9 :


,diff .shr.

E
= SHconst2 con
r

r 2 + R 2IC
2
2
R OC R IC

R OC

R IC

r
erfc(A) r dr + erfc(A) r dr erfc(A) r 2
R IC

Eq. 2.9 (a)


where, A is given as;

19

A = (R OC r ) /

Eq. 2.9 (b)

While Eq. 2.9 (a) provides an expression for stress which can be integrated to yield Eq.
2.10 as a closed-form expression for the residual stress caused by different shrinkage.
,diff .shr .

E
= SH const2 con
r

r 2 + R 2IC

[
f
(
R
)

f
(
R
)]
+
f
(
r
)

f
(
R
)

erfc
(
A
)

r
2

OC
IC
IC
2
R OC R IC

Eq. 2.10 (a)


where,

R
1
2 A
f ( r ) = 2 erfc ( A ) A 2 + erfc ( A ) OC A +

+
2
2

4 eA

R OC
erf ( A )
+
A2
8
2e

Eq. 2.10 (b)


Figure 2.3 shows the example of resulting stress gradients in a concrete ring using
Eq. 2.10 for the case of a ring without steel restraint (ROC=225mm, RIC=150mm, SHconst=

-100 ). The stresses that are developed due to self-restraint have steep gradients

near the outer surface when drying begins (i.e., low values of ). With increasing drying
time (i.e., increasing ), the gradient flattens and the shrinkage becomes more uniform.
Decreasing the severity of the humidity profile implies that the strain (or stress)
developed by self-restraint will be decreased.

20

3
Outer
Surface
Of the Conc. Ring

Inner
Surface
Of the Conc. Ring

Stress (MPa)

=0.002
=0.004
=0.008
=0.02
=0.04
=0.08
=0.2
=0.5
=100

-1

0.02

0.04

0.06

0.08

Distance from the outer surface (m)

Figure 2.3 Stress gradient due to self-restraint in the concrete ring (i.e., without the steel
ring (Eq. 2.10)) (ROC=0.225m, RIC=0.150m, Econ=21GPa, SH-const= -100 )

2.5.2 Stress Development in the Concrete Ring due to Steel Ring Constraint: ,rest. ring(r)
The stress development due to the external restraint from the steel ring must also
be considered. Previous research has shown that the restraint from the steel ring can be
simulated by separating the steel and concrete ring and treating the problem as a shrinkfit problem. The concrete ring is permitted to shrink some amount (USH) that is equal
to that caused by drying and autogenous shrinkage. The composite cylinder can be
considered to have a fictitious pressure that is applied on the outer surface of the steel
ring that is equal to the pressure on the internal surface of the concrete ring (Weiss and
Shah, 2002; Weiss and Ferguson, 2001; Hossain and Weiss, 2003a; Hossain and Weiss,

21

2006). The pressure is adjusted until the steel ring is compressed (USteel) and the
concrete ring is expanded (Uconc) to compensate for the shrinkage as shown in Figure
2.4 (further details are provided in (Hossain and Weiss, 2003a)).

Free Shrinkage
(No Steel Ring)

ROC
RIS
ROS
(RIC)

USH

Uconc
USH
Usteel

(a) Before Shrinkage


Occurs

(b) Free Shrinkage


(No Steel Ring)

(c) Shrinkage with


Steel Ring

Figure 2.4. Conceptual illustration of the shrink-fit approach

The fictitious external pressure could be determined from Eq. 2.11 since the strain
in the steel can be obtained experimentally using the strain gage on the inner surface of
the steel ring (Hossain and Weiss, 2003a).
2
R OS
R 2IS
p i = steel ( t ) E S
2
2 R OS

Eq. 2.11

The fictitious pressure that can be thought to act on the steel ring could be related
to an internal fictitious pressure that acts on the concrete ring and as a result the stress
distribution in the concrete ring can be determined as shown in Eq. 2.12 (Hossain and
Weiss, 2003a).

22

2
2
R OC
R OS
,rest . ring ( r ) = p i 2
1+ 2
2
R OC R OS
r

Eq. 2.12

Figure 2.5 shows an example of the stress gradient in the concrete ring calculated
using Eq. 2.12 for the case of the steel ring restraint (ROC=225mm, ROS=RIC=150mm,
RIS=140.6mm, Esteel=200GPa, SH-const= -100 ) where steel is the strain that would be
measured in the steel ring. The maximum stress develops in the circumferential direction
along the interface between concrete and steel. It can be noticed that Eq. 2.12 is also able
to be used in the case of autogenous shrinkage because it can consider a uniform
shrinkage of the matrix.

1.6
Inner
Surface
of the Conc. Ring

=0.002
=0.004
=0.008
=0.02
=0.04
=0.08
=0.2
=0.5
=100

Stress (MPa)

1.2

0.8

0.4

0.02

0.04

0.06

0.08

Distance from the outer surface(m)

Figure 2.5 Stress gradient due to steel ring constraint in the concrete ring (Eq. 2.12)
(ROC=0.225m, RIC=ROS=0.150m, RIS=0.1406m, Econ=21GPa, Esteel=200GPa,
SH-const= -100 )

23

2.5.3 Superposition of Differential Shrinkage and Steel Ring Restraint


By combining Eq. 2.10, 2.11, and 2.12, the stress development in the concrete
ring can be expressed as the sum of the stress caused by the self-restraint (i.e., differential
shrinkage in Section 2.5.1) and the external restraint (i.e., stress caused by the steel ring
in Section 2.5.2). Eq. 2.13 enables the stress ( ) at any point in the concrete ring to be
calculated.

( r, ) = ,rest . ring + ,diff .shr .

= steel ( t ) E S

1 + R OC
2
R OS
)
r2

2
R OS
R 2IS

2 (R

2
OC

2
2

SH const E con r + R IC
(f ( R OC ) f ( R IC )) + f ( r ) f ( R IC ) erfc( A ) r 2

2
2
2
r
R OC R IC

Eq. 2.13

Figure 2.6 shows the total stress gradient in the concrete calculated using Eq. 2.13
where the concrete ring is restrained by a steel ring (ROC=225mm, ROS=RIC=150mm,
RIS=140.6mm, Esteel=200GPa, Econ=21GPa, SH-const= -100 ). As expected, the stress
development is dependent on the extent of drying time (i.e., ). It can be noticed that at
short drying times the maximum stress occurs on the outer circumference. As drying
continues, the position of Actual-Max changes from the outer circumference of the concrete
ring to the inner surface after exceeds approximately 0.2 (i.e., longer drying time
depending on the elastic modulus of the concrete) (Figure 2.6).

24

=0.002
=0.004
=0.008
=0.02
=0.04
=0.08
=0.2
=0.5
=100

Stress (MPa)

Inner
Surface
of the Conc. Ring
Steel
Ring

Econ / Esteel =0.105

-1
0

0.02

0.04

0.06

0.08

Distance from the outer surface (m)


Figure 2.6 Stress gradient in the concrete ring exposed to circumferential drying with
steel ring (Eq. 2.13) (ROC=0.225m, RIC=ROS=0.150m, RIS=0.1406m, Econ=21GPa,
Esteel=200GPa, SH-const= -100 )

This result can be explained by the fact that, at initial drying times , the stress
development in a concrete ring is mainly governed by self restraint, while it is mainly
governed by the external steel ring restraint for longer drying periods.
To verify the applicability of the superposition of stresses obtained from Eq. 2.10
and 2.12, each case of restraint (self restraint: Eq. 2.10 and steel ring restraint: Eq. 2.12)
will be compared with the results of finite element analysis separately, and Eq. 2.13 will
be finally compared with the finite element analysis in the following section.

25

2.6 Finite Element Analysis (FEA)


A series of finite element analyses were simulated using finite element analysis
(FEA) in ANSYS to verify the appropriateness of the developed approach. The ring
specimen was considered as an axi-symmetric geometry where the drying shrinkage
gradient occurs along a radial section of the ring specimen. Quadrilateral eight-node
elements were used for the simulations.

The geometry of the concrete specimens

simulated had a depth of 75 mm, an outer radius of concrete 225 mm, and an inner radius
of concrete 150 mm. The geometry and material properties of the steel ring were fixed
(thickness = 9.4 mm, Es = 200 GPa, and s = 0.3). Poissons ratio and the effective elastic
modulus of concrete were selected to be 0.18 and 21 GPa, respectively. To verify the
effect of self-restraint and steel ring restraint, two types of analyses were performed: with
the steel ring and without the steel ring.
For this analysis, it was assumed that the concrete ring dries only from the outer
circumference of the ring. Eq. 2.3 was used to determine the relative humidity profile
while RHI and RHS were assumed as 100% RH and 50% RH, respectively. A free
shrinkage strain constant (SH-const) was assumed to have a value of -100 .
Finite element analyses were performed by varying the severity of the drying (
=0.001 ~ 100). Theoretically, an infinite value of corresponds to a concrete ring,
which dries uniformly along the radius (as described in Eq. 2.1), however a value of
100 was found to be essentially equivalent to higher values, and, as such, it is the largest
simulated.
Because the FEA program does not support the drying shrinkage loading directly,
shrinkage was introduced using a temperature load variable substitution.

The

26

temperature distribution was assumed to vary as an error function (erf) (the change in
temperature in this thesis is analogous to a change in relative humidity) and a thermal
expansion coefficient was the input for the concrete (this can be thought of as being
analogous to the free shrinkage constant) to obtain the same effect as drying shrinkage.
The maximum stress was selected along a plane perpendicular to the depth direction at
midheight in the ring specimen to obtain the stress values that were reported.
The contact condition between steel and concrete for the FE analysis was chosen
to simulate an unbonded condition by using a pseudo-bar that essentially permitted a
friction-free condition along the vertical direction of the ring, which is compatible with
the results previously presented in the literature (Moon et al., 2004).

2.7 Comparison of the Analytical Approach with FEA and Application to an Experiment

2.7.1 Comparison with FEA


To verify the appropriateness of the equations that were developed, the stress
values calculated by Eq. 2.10, 2.12, and 2.13 were compared with those obtained from
the finite element analyses. Figure 2.7 shows the stress gradients in the concrete ring
without the steel ring restraint as compared to the solution presented in Eq. 2.10. A
generally good agreement was found resulting in minor differences (below 5%) between
the finite element analysis and the analytical solution (Eq. 2.10) especially for low
values (i.e., short drying times) along the outer circumference of the ring.

27

FEA =0.002
FEA =0.02
FEA =0.2
FEA =100
Eq. 10

Stress (MPa)

-1
0

0.02

0.04

0.06

0.08

Distance from the outer surface (m)

Figure 2.7 Comparison of stress gradients due to self-restraint in the concrete ring
without steel ring (FEA and Eq. 2.10) (ROC=0.225m, RIC=0.150m, Econ=21GPa,
SH-const= -100 )

To explain some of the differences along the outer circumference of the ring, it
should be remembered that the analytical solutions (Eq. 2.10, 2.12, and 2.13) were
developed considering plane stress conditions with a uniform stress field in the z (depth)
direction. Some of these differences occur due to the plane stress assumption which is
not especially well suited for very short drying times. To verify this, additional finite
element simulations were performed using both plane stress and plane strain conditions.
Figures 2.8 and 2.9 show the stress gradients in the concrete ring obtained from the plane
stress and strain analysis assuming a uniform stress field in the z direction. As expected,

28

the residual stress (Actual-Max) that was obtained directly from finite element analysis of
the actual geometry is located between the plane strain and plane stress approximations
(Figure 2.10). Since Eq. 2.10 was developed from the plane stress condition, the values
from Eq. 2.10 matched well with those from the plane stress condition finite element
analyses (Figure 2.10).

The results converge toward the plane stress solution as

increases (Figure 2.10). It is believed that for a first approximation, this approach is
reasonably accurate since it should be remembered that there is a high degree of
uncertainty associated with the actual ring specimen tests in which micro-cracking occurs
along the outer circumference.

29

FEA =0.002
Plane Strain
Plane Stress
Eq. 10

Stress (MPa)

-1
0

0.02

0.04

0.06

0.08

Distance from the outer surface (m)


Figure 2.8 Comparison of stress gradients due to self-restraint in the concrete ring
without steel ring (FEA and Eq. 2.10) (=0.002) (ROC=0.225m, RIC=0.150m,
Econ=21GPa, SH-const= -100 )

30

FEA =0.02
Plane Strain
Plane Stress
Eq.10

Stress (MPa)

-1
0

0.02

0.04

0.06

0.08

Distance from the outer surface (m)


Figure 2.9 Comparison of stress gradients due to self-restraint in the concrete ring
without steel ring (FEA and Eq. 2.10) (=0.02) (ROC=0.225m, RIC=0.150m, Econ=21GPa,
SH-const= -100 )

31

FEA (Exact Geometry)


FEA (Plane Strain)
FEA (Plane Stress)
Eq. 10

Stress (MPa)

0.001

0.01

0.1

10

100

Figure 2.10 Comparison of Actual-Max due to self-restraint in the concrete ring without
steel ring (ROC=0.225m, RIC=0.150m, RIS=0.1406m, Econ=21GPa, SH-const= -100 )

Figure 2.11 shows the stress gradients that developed as a result of the steel ring
restraint. The strain values (steel(t)) obtained from the finite element analyses were used
for the calculations of stress using Eq. 2.12, and it can be seen that the stress gradients
calculated by Eq. 2.12 corresponded well with those of the finite element analyses.
Since each component of stress development (self-restraint and steel ring
restraint) from the analytical solution was separately evaluated, the stress gradients
calculated using Eq. 2.13 (the sum of Eq. 2.10, self restraint and Eq. 2.12, steel ring
restraint) were compared with the finite element analyses and a reasonably good
agreement was also observed (Figure 2.12). As a result, it appears appropriate to utilize

32

the analytical approach described here, if the linear diffusion, linear creep, and no
microcracking assumptions are used.

1.2
FEA =0.002
FEA =0.02
FEA =0.2
Eq. 12

Stress (MPa)

0.8

0.4

0
0

0.02

0.04

0.06

0.08

Distance from the outer surface (m)

Figure 2.11 Comparison of stress gradients due to steel ring restraint in the concrete ring
(FEA and Eq. 2.12) (ROC=0.225m, RIC=ROS=0.150m, RIS=0.1406m, Econ=21GPa,
Esteel=200GPa, SH-const= -100 )

33

FEA =0.002
FEA =0.02
FEA =0.2
FEA =100
Eq. 13

Stress (MPa)

-1
0

0.02

0.04

0.06

0.08

Distance from the outer surface (m)

Figure 2.12 Comparison of stress gradients due to circumferential drying in a concrete


ring with steel ring (ROC=0.225m, RIC=ROS=0.150m, RIS=0.1406m, Econ=21GPa,
Esteel=200GPa, SH-const= -100 )

2.7.2 Application to Experiments


To assess whether the values predicted by this approach are logical and of the
right order of magnitude, Eq. 2.13 was used to describe a series of experiments that were
performed on a mortar with a water to cement ratio of 0.50 and an aggregate volume of
50% (Hossain and Weiss, 2005). The ring specimen had an outer radius of the concrete
of 225 mm, an inner radius of the concrete of 150 mm, and an inner radius of the steel
ring which was 140.5 mm (s=0.3, Es=200 GPa).

The age-dependent diffusion

34

coefficient was determined for this mixture as described in Eq. 2.14 using electrical
conductivity measurements (Moon et al., 2004; Rajabipour et al., 2004).
D( t ) = d1 +

d2
t

(10-12 m2/sec)

Eq. 2.14

where, d1 had a value of approximately 9,0 d2 had a value of 112.64, and t is the drying
time, which is given in days. The age-dependent splitting tensile strength (fsp) was
measured and found to be described by Eq. 2.15 (Hossain and Weiss 2003a).
fsp (t ) = f

C4 ( t t 0 )
1 + C4 ( t t 0 )

Eq. 2.15

where, C4 was the rate constant with a value of 2.45 day-1, to was the setting time of 0.25
days, and f was the long-term strength, which had a value of 4.7 MPa.
Before the maximum residual stress Actual-Max was calculated using Eq. 2.13, two
additional material properties needed to be estimated. First, the free shrinkage strain
constant was estimated from unrestrained free shrinkage tests. It has been previously
shown (Hossain and Weiss 2003a) that the free shrinkage displacement (USH) can be
considered as the combination of the autogenous shrinkage displacement in a sealed
specimen (UAUTO) and the drying shrinkage displacement (UDRY) (Eq. 2.16).

U SH = U AUTO + U DRY

Eq. 2.16

Based on the experimental results, Eq. 2.17 was found to describe the autogenous
deformation (UAUTO) in an unrestrained sealed specimen as a function of time
U AUTO ( t ) = R IC AUTO ( t ) = R IC C1 ( t t 0 ) C 2

Eq. 2.17

where the coefficients C1=-52.83, C2=0.59, and t0=0.25 (mortar, 0.5 w/c, 50% sand)
correspond well to measurements in a sealed concrete (Weiss and Ferguson, 2001b). The

35

drying shrinkage deformation (UDRY) can be obtained using Eq. 2.18 which was derived
by integrating Eq. 2.8 (Hossain and Weiss, 2005).

U DRY ( t ) =

2 DRY const R IC
2
R OC
R 2IC

R OC

R IC

erfc(A) r dr

Eq. 2.18

It should be noted that the free shrinkage strain constant ( SH const ) in Eq. 2.8 is
renamed in Eq. 2.18 as the effective drying shrinkage strain constant ( DRY const ) to
denote that it only considers the drying shrinkage component since the autogenous
shrinkage component is already considered using Eq. 2.18.

The effective drying

shrinkage constant of -1,200 was estimated for this specific concrete, and the
deformations predicted by Eq. 2.17 and 2.18 were well correlated to physical
measurements.
Second, the effective elastic modulus was estimated using displacement
compatibility as illustrated in Figure 2.4 (Attiogbe et al. 2004)
Econ ( t ) =

pi ( t ) R IC
2
(1 + C )R OC
+ (1 C )R 2IC
2
2
Uconc ( t ) (R OC R IC )

Eq. 2.19

where, pi(t) and Uconc(t) can be calculated by Eq. 2.11 and Eq. 2.16.
With all the material properties now known, Eq. 2.13 was used to compute the
stress gradients along the radius of the specimen as shown in Figure 2.13. It can be seen
in Figure 2.13 that almost immediately upon drying the stress that develops on the outer
circumference of the concrete ring is higher than the tensile strength. It can be assumed
that in the region where the maximum stress exceeds the tensile strength microcracking
would occur (Bazant and Chern, 1985; Granger et al., 1997; Grasley et al., 2003; Weiss,

36

1999). It can also be seen that the depth of the concrete in which the maximum stress
exceeds the tensile strength increases with time.
20
Day of Drying
2 days
7 days
12 days
Strength

Stress (MPa)

16

12

2 days

7 days

12 days

0
0

0.02

0.04

0.06

0.08

Distance from the outer surface (m)

Figure 2.13 Stress gradients due to circumferential drying in the concrete ring with steel
ring (calculated using Eq. 2.13 and real test data)
(0.50 W/C, 50% Sand, ROC=0.225m, RIC=ROS=0.150m, RIS=0.1406m)

Figure 2.14 shows the progression of stable cracking (i.e., the depth where the
residual stress exceeds the tensile strength) in a concrete ring with time, thereby
suggesting that the crack has propagated at least 8 mm from the surface and this depth of
the stable crack is likely greater since these stresses can be expected to be redistributed
across the cross-section due to cracking. It should be noted that the laboratory specimen
developed a visible crack during the 12th day of drying (Hossain and Weiss, 2005), at

37

which time the magnitude of the average far-field stress is on the order of 2.5 to 3.0 MPa.
If it is assumed that these specimens have properties similar to other mortars (KIC ~ 0.5 to
1.0 MPa mm) (Shah and Jenq, 1989) this far field stress and crack depth could be

Depth from the circumferce where 1(t) reaches fsp(t) (m)

consistent with the level of stress that is required to cause through-cracking.

0.01

Visible Crack
0.008

0.006

0.004

0.002

0
0

12

Age of Drying (days)

Figure 2.14 Predicted depth of microcracking in the concrete ring


(0.50 W/C, 50% sand, ROC=0.225m, RIC=ROS=0.150m, RIS=0.1406m)

While this approach appears reasonable, further work is necessary to validate and
calibrate this procedure. Specifically, additional work is being performed to enable a
description of a wider range of moisture gradients (Bazant, 1986; Rajabipour et al., 2004;
Grasley and Lange, 2004). Acoustic emission testing is being refined to assess the extent

38

of the stable crack that can be expected to develop from the outer surface of the concrete
ring (Neithalath et al., 2005) and to document how these cracks develop over time. In
addition, further work is being performed to describe how these microcracks coalesce to
form a visible crack (Bisschop and Van Mier, 2001; Hossain, 2003d; Hossain and Weiss,
2005).

2.8 Conclusions
This chapter has described the development of an analytical approach for
calculating stress development in a restrained concrete ring test under circumferential
drying. An equation was developed that considered the stress that develops due to
restraint by separating the problem into two components: 1) self-restraint due to moisture
gradients (Eq. 2.10) and 2) external restraint due to the steel ring (Eq. 2.12). The
analytical residual stress equation (Eq. 2.13) was compared with the results from finite
element analysis and a reasonable agreement was shown. It was also verified that the
restrained ring that dries from the circumference has a high stress at the outer radius for
short drying time (i.e., low ) and a higher potential to crack from the outer
circumference. The developed solutions were applied to properties that correspond to an
actual test to show a potential application for the basic approach, which resulted in
reasonable stress fields. However, further work is needed to demonstrate the application
to a wider range of concretes and to improve how relaxation and microcracking are
considered.

39

CHAPTER 3: DEGREE OF RESTRAINT

3.1 Introduction
Cementitious materials change volume as a result of autogenous, drying, or
thermal shrinkage.

When these volume changes are prevented by the structure

surrounding concrete, residual tensile stresses can develop inside the material. If these
residual stresses exceed the tensile strength of concrete, cracking may occur. Over the
years, engineers have sought to develop simple tests to assess how susceptible a given
concrete mixture may be to shrinkage cracking.

While tests like ASTM C 157-04

(standard test method for the length change of hardened hydraulic-cement mortar and
concrete) are frequently used to measure the free shrinkage of a concrete mixture, free
shrinkage by itself is not sufficient to predict whether cracking will occur. Rather, the
potential for cracking is dependent on the interaction of several factors, including the
magnitude of free shrinkage, rate of shrinkage, elastic modulus, degree of restraint,
creep/stress relaxation, and fracture toughness (Weiss et al., 2000).
Several researchers have developed experimental procedures to study how
material properties influence the potential for shrinkage cracking (Swamy and Stavrides,
1979; Carlson and Reading, 1988; Weigrink et al., 1996).

Several of these test

procedures have been suggested to evaluate residual stresses that develop in concrete
when shrinkage is restrained. Examples of such tests include the restrained ring test
(Grzybowski and Shah, 1990; Hossain and Weiss, 2004), the passive linear restraint tests

40

(Springenschmidt et al., 1985; Weiss et al., 1998), and the active linear restraint tests
(Kovler, 1994; Toma et al., 1999; Altoubat and Lange, 2002;). While a review of these
test methods can be found elsewhere (Weiss, 1999), this paper focuses on assessing the
degree of restraint and the measurable strain in the restrained ring test.
Due to its simplicity and low cost, the restrained ring test has been used by
numerous researchers to assess the potential for shrinkage cracking in concrete mixtures
(Carlson and Reading, 1988; Grzybowski and Shah, 1990; Kovler et al., 1993; Weiss et
al., 1999; Attiogbe et al., 2004). The ring test consists of a concrete annulus that is cast
around a steel ring. As the concrete ring dries, it attempts to shrink. The steel ring
prevents this shrinkage causing tensile stress to develop in the concrete. If these stresses
are large enough, cracking may occur. Various geometries of the ring test have been
used by researchers.

AASHTO PP34-99 (AASHTO standard practice for cracking

tendency using a ring specimen) was developed as a provisional standard test method for
assessing the cracking potential of a concrete mixture when it is restrained. AASHTO
PP34-99 uses a 12.5 mm thick steel ring, a 75 mm thick concrete ring, and allows the
concrete to dry from the outer circumference. While this method has been used in several
studies, it has been reported that the low degree of restraint provided by the steel ring
results in a fairly long time before the first visible cracking is observed (Attiogbe et al.,
2003). As a result, an alternative test geometry was developed for inclusion as an ASTM
standard (ASTM C 1581-04: standard test method for determining age at cracking and
induced tensile stress characteristics of mortar and concrete under restrained shrinkage).
ASTM C 1581-04 uses a 12.5 mm thick steel ring, however, the concrete wall thickness
was reduced to 37.5 mm to encourage cracking to develop at an earlier age. This

41

restriction in specimen size, however, makes it difficult to test concrete with a larger
aggregate or fiber reinforcement.
While the development of these standards is a positive step forward, it should be
noted that the restrained ring test is not intended to measure a fundamental material
property; rather, the ring test measures the materials response to a specific stimulus (e.g.,
drying at a constant temperature and relative humidity) under specific boundary
conditions (e.g., specimen size, drying direction, and degree of restraint).

No one

specimen geometry or external stimulus can be used to simulate all the possible
conditions that may be encountered in the field. It is therefore important that procedures
are developed to enable the ring test to be interpreted for a variety of applications.
Specifically, this may include applications where the degree of restraint provided by the
structure may vary over a wide range (15-90%) (NCHRP Project 12.37, 1995) or when
the surface to volume ratio (this is analogous to varying thickness) of the concrete may
vary from structure to structure. For example, the developer of a repair material may
want to test their material under a high degree of restraint with a rapid rate of shrinkage
to simulate a thin fully-bonded repair. Alternatively, someone developing a bridge deck
mixture may be more interested in testing a thicker section (i.e., slower overall shrinkage
rate) with a lower degree of restraint.
This thesis will specifically look at how various geometric factors influence the
results of the ring test. An analytical solution will be presented that describes the degree
of restraint as a function of test geometry and material properties.

Finite element

simulations will be performed to quantify how material properties, geometry, and drying
direction influence the degree of restraint.

It is intended that the results of these

42

simulations will improve how the restrained ring test is used since it helps tailoring the
ring geometry for specific applications.

3.2 Ring Specimen Geometry


Figure 3.1 shows the typical geometry of the restrained ring test. The focus here
primarily will be on assessing two aspects of the restrained ring test: the influence of the
drying direction and the influence of the steel ring thickness.

Figure 3.1 Drying direction and the resulting stress development of restrained ring

43

First, the influence of the drying direction will be considered. The simplest case
which encountered consists of uniform shrinkage throughout the concrete in both the
radial and height directions. This may be observed in the cases of autogenous shrinkage
and in many of the basic analytical modeling approximations.

The second drying

condition considered is drying from the top and bottom of the concrete ring. This
geometry may be preferred compared to circumferential drying since it allows the
concrete ring to be sufficiently thick to enable large aggregate and fiber reinforced
concrete to be tested (Hossain and Weiss, 2006). Top and bottom drying results in
uniform shrinkage along the radial direction but not along the height direction. Since
both the uniform shrinkage and top and bottom drying conditions have uniform shrinkage
along the radial direction, the residual stresses that develop in these cases are the highest
at the inner radius of the concrete ring and decrease as a function of 1/r2 (where r is
radial distance) through the concrete wall thickness (Hossain and Weiss, 2004). The
third drying condition considered in this paper is the case of circumferential drying as
suggested by the standard testing procedures ASTM C 1581-04 and AASHHTO PP3499. In specimens that dry from the outer circumference, shrinkage is uniform throughout
the height of the specimen, but not through the radial direction. Since the specimen loses
the majority of water at the circumference (i.e., drying face), the stresses are highest at
the drying face (Weiss and Shah, 2001; Moon et al., 2004; Moon and Weiss, 2006),
which results in a complicated stress distribution that changes shape over time.
In addition to considering the drying direction, this thesis investigates the
influence of steel ring thickness. A thicker steel ring will provide a higher degree of
restraint, resulting in higher stress development in the concrete. On the other hand, the

44

increase of the steel ring thickness decreases the magnitude of strain that can be measured
at the inner surface of the steel ring, which is needed for stress calculations inside
concrete (Hossain and Weiss, 2004). Therefore, it is necessary to quantify how the
thickness of the steel ring changes the degree of restraint, as well as the stress
measurements.

3.3 Quantifying the Degree of Restraint under Uniform Shrinkage (Analytical Approach)
Figure 3.2 can be used to provide a basic illustration of the concept of the degree
of restraint (Weiss, 1999; See et al., 2003). In this conceptual example, a concrete prism
is thought to be axially restrained by a steel prism (Figure 3.2.(a)). If the concrete shrinks
and there is no connection between the concrete and the steel (free boundary condition), a
difference in the length of the specimens will exist as illustrated in Figure 3.2.(b). This
difference is what one would expect to measure in a free shrinkage test like ASTM C
157-04. If the concrete is restrained by the steel prism (assuming bending is prevented),
the concrete and steel will deform together (Figure 2.(c)).

45

Figure 3.2 A conceptual illustration of the degree of restraint

This will result in some of the shrinkage of the concrete being prevented, resulting
in a smaller length change than for a specimen under the free shrinkage condition. The
degree of restraint ( ) can be defined as;
=

L free L restrained
L restrained
= 1
L free
L free

Eq. 3.1 (a)

where Lfree is the displacement of concrete due to free shrinkage, and Lrestrained is the
displacement of concrete under restraint as shown in Figure 3.2. The degree of restraint
can be estimated for linear specimens using Eq. 3.1 (b) (ACI 207.2R, 1995):
=

EC AC
E C AC + ES AS

Eq. 3.1 (b)

46

where AC and AS are the cross-sectional areas, and EC and ES are the elastic moduli of the
concrete and steel elements, respectively. In the case of a restrained ring test, the degree
of restraint can be written as Eq. 3.1 (c):

U SH

R IC

(t) U S R (t)

U SH

Eq. 3.1 (c)

OS

R IC

(t)

where USH|RIC(t) is the free shrinkage displacement of the concrete ring (i.e., if no steel
ring was present) measured at the inner surface of the concrete ring (RIC) and US|Ros(t) is
the actual displacement of the concrete ring under the restraint provided by the steel ring.
The degree of restraint in Eq. 3.1 (c) is defined at the concrete-steel interface (RIC=ROS;
Figure 3.1) where the restraint is maximum. The degree of restraint can vary between
zero (corresponding to no restraint; USH|RIC = US|Ros) and one (corresponding to perfect
restraint; US|Ros = 0).
For the case of uniform shrinkage (i.e., no moisture gradient), the free shrinkage
displacement, USH|RIC(t) can be calculated using Eq. 3.2.a based on the linear free
shrinkage of the concrete, SH (t) (Weiss et al., 1999).
U SH R ( t ) = SH ( t ) R IC

Eq. 3.2 (a)

IC

The actual displacement of the restrained concrete ring at ROS, US|Ros(t), can be
calculated using the strain that is measured on the inner surface of the steel ring (st(t)) as
shown in Eq. 3.2.b.
U S R ( t ) = st ( t )
OS

1
2
2
R OS
+ R 2IS S (R OS
R 2IS )
2R OS

Eq. 3.2. (b)

47

where S is the Poissons ratio of steel. By substituting 3.2.a and 3.2.b into 3.1
(c), the degree of restraint can be written as follows:

1 ( t ) R 2IS
= 1 st
2 (1 + S ) + (1 S )
2 SH ( t ) R OS

Eq. 3.3

Meanwhile, the ratio st(t)/SH(t) can be related to the specimen geometry and the
material properties of the concrete and steel by considering the interfacial pressure that
develops at the concrete-steel interface. The magnitude of the interfacial pressure can be
calculated using either Eq. 3.4 (a), which is based on the measured steel strain from an
experiment or Eq. 3.4 (b), which is based on fundamental computation using a shrinkfit approximation that incorporates the free shrinkage strain of the concrete and elastic
properties of the concrete and steel (Hossain et al. 2003; Hossain and Weiss 2004):
2
R OS
R 2IS
p 0 = st ( t ) E S
2
2 R OS

SH ( t ) E 'C
p0 = '
2
2
2
E C [(1 + S )R 2IS + (1 S )R OS
] [(1 C )R OS
+ (1 + C )R OC
]
+
2
2
2
2
ES
(R OS R IS )
(R OC R OS )

Eq. 3.4 (a)

Eq. 3.4 (b)

In these equations, ES is the elastic modulus of steel, EC is the effective elastic


modulus of concrete (considering the creep effect), and C is the Poissons ratio of
concrete. Substitution of Eq. 3.4 (a) into 3.4 (b) enables the ratio st(t)/SH(t) to be
determined as:

48

st ( t )
E'
= C
SH ( t ) E S

2
R
1 IS
R OS

R
(1 + S ) IS
R OS
E 'C

ES
R
1 IS
R OS
2

+ (1 S )

R
(1 + C ) OC

R OS

+ (1 C )

R
1 OC
R OS

Eq. 3.5
Eq. 3.5 can now be substituted into Eq. 3.3, which results in an expression for the
degree of restraint as only a function of the geometry and material properties (namely the
stiffness and the Poissons ratio) of the concrete and steel.
= 1

E 'C

ES

Eq. 3.6

E 'C

ES

R
1 IS
R OS
R
1 OC
R OS

R
(1 + C ) OC

R OS

(1 + S ) IS

R OS

+ (1 C )

+ (1 S )

This enables one to determine the degree of restraint prior to conducting the
experiment and provides the opportunity to adjust the test geometry to satisfy the desired
degree of restraint that best approximates the actual field conditions.
In addition to determining the degree of restraint, an important parameter that
needs to be considered is the magnitude of the strain that will be measured from the inner
surface of the steel ring (st(t)). From a practical point of view, resolving the magnitude
of the residual stress in a typical concrete with an accuracy of better than 10 psi requires a
minimum steel strain value of 50 before the specimen fails, which can be ensured by
determining the minimum value of st/SH at failure for a given ultimate free shrinkage of
concrete (SH). For example, for concrete with an ultimate free shrinkage of SH = 400
, a minimum value of at least st/SH = 0.125 would be desired to obtain a strain in the

49

steel of 50 (st = 50 ). Eq. 3.5 can now be used to check whether the test geometry
satisfies this requirement.
It is now possible to evaluate the effect of specimen geometry on the degree of
restraint () and st/SH. Figure 3.3 (a) illustrates the influence of steel ring thicknesses
by varying RIS/ROS between 0 (a solid cylinder) and 1 (an infinitely thin ring). As
expected, the degree of restraint decreases with decreasing the thickness of the steel ring.
In addition, a stiffer concrete (higher effective elastic modulus) results in a lower degree
of restraint. It is shown that in all cases when RIS/ROS decreases below 0.6, only a slight
increase (less than 10%) in the degree of restraint occurs by further increasing the steel
ring thickness. Similarly, Figure 3.3 (b) shows how an increase in the concrete ring
thickness (ROC/RIC) reduces the degree of restraint.

50

(a)

(b)
Figure 3.3 The influence of specimen geometry on the degree of restraint and the ratio
st/SH: (a) influence of the concrete stiffness and (b) influence of the concrete ring
thickness (EC = 21 GPa, ES = 200 GPa, uniform shrinkage)

51

Eq. 3.5 and 3.6 provide an analytical approach to determine the degree of restraint
and to ensure that an accurate strain measurement can be obtained from the steel ring so
that the stress computation can be performed. To illustrate the applicability of this
technique for predicting the degree of restraint prior to testing, the degree of restraint
determined using Eq. 3.6 was compared with experimentally-measured values (using Eq.
3.3), as illustrated in Figure 3.4.

Figure 3.4 Comparison between the analytical solution and the experimentally
determined degree of restraint before the through-cracking (top and bottom drying)

The experimental data used here was obtained from a series of mortar ring tests
with a w/c of 0.30 and 50% fine aggregate by volume (Hossain and Weiss 2004). Three

52

different restrained rings were used with varying steel thicknesses (3.2, 9.4, and 19 mm).
The geometry parameters were as follows: ROC = 225 mm, RIC = ROS = 150 mm. The
mortar rings were exposed to drying from the top and bottom in a 50 % relative humidity
environment. The effective elastic modulus of concrete was estimated to be 60% of the
actual measured elastic modulus to account for creep effects (Bazant and Wittmann
1982). The free shrinkage strain (SH(t)) and the steel ring strain (st(t)) were measured
experimentally and used in Eq. 3.3 to calculate the degree of restraint and these results
were compared with the values obtained from the analytical solution (Eq. 3.6). It must be
noted that the analytical approach discussed earlier is initially developed for a uniform
shrinkage condition. However, the results of a series of finite element simulations (as
described in the next section) suggest that the analytical approach can be used regardless
of the drying condition of the concrete ring. Figure 3.4 shows that there is reasonable
agreement between the analytical approach and the experimentally measured degree of
restraint. It is also evident that the degree of restraint decreases over time as the stiffness
of concrete increases.

3.4 Quantifying the Degree of Restraint in the Presence of Moisture Gradient


To consider how the degree of restraint is influenced by the presence of a
moisture gradient inside concrete (caused by preferential drying from exposed surfaces),
a series of finite element simulations were performed using the commercial analysis code
ANSYS. The ring specimen was considered to be axi-symmetric and was modeled using
quadrilateral eight-node elements. Both rings (concrete and steel rings) were considered

53

to have a height of 75 mm. A fixed value of RIC =ROS = 150 mm was used in the
simulations. Also, a base value of ROC = 225 mm and RIS = 140.5 mm were used,
however for geometrical analysis, the values of ROC and RIS were varied to provide an
RIS/ROS ranging from 0.5 to 1 and an ROC/RIC ranging from 1.25 to 2. Other parameters
used in the simulations were: EC = 21 GPa, ES = 200 GPa, c = 0.18, and s = 0.30.
No creep non-linearity or microcracking effects were considered in the finite element
analysis. A perfectly unbonded condition between the steel and concrete rings was
assumed (i.e., no transfer of tensile and shear stresses at the interface). This interface
condition can be approached experimentally by using a form release agent or a thin
plastic sheet between the concrete and steel (Hossain et al. 2003).
Three different cases of drying were considered: a) uniform drying, b) top and
bottom drying, and c) circumferential drying. In cases (b) and (c), a moisture gradient
develops inside the concrete as it dries, preferentially from the exposed surfaces. To
incorporate moisture gradients in the model, a linear moisture diffusion function was
used to describe the relative humidity at each point inside the concrete as a function of
the drying time (t), as described by Eq. 3.7 (Moon et al. 2004).

x
RH(x, t) = RH I ( RH I RH S ) erfc

2 Dt

Eq. 3.7

where, RH(x, t ) is the relative humidity at a depth x from the drying surface, erfc is the
complementary error function, RH S is the relative humidity at the surface of the
specimen (considered to be 50% in the simulations), RH I is the internal humidity if the
specimen was completely sealed ( RH I was assumed here to be 100% and, as such, the
effect of autogenous shrinkage was neglected), and D is the aging moisture diffusion

54

coefficient of concrete (m2/sec). It should be noted that although non-linear diffusion


coefficients are commonly suggested for use in concrete (Bazant and Najjar 1971), and
while these equations may be more scientifically correct, the application of a linear
coefficient here has enormous computational benefits with no loss in the generalities of
the conclusions of this thesis. By considering D to be only dependent on specimen age, a
single parameter can be introduced as = 2 Dt to represent the time and diffusion
coefficient variations in Eq. 3.7. It has been shown that this approach can be used to
reasonably describe the moisture gradient that develops inside mortar due to drying
(Moon et al. 2004). Figure 3.5 shows an example of relative humidity profiles that
correspond to different values of . A low value (i.e., < 0.01) refers to a very steep
moisture gradient, while the severity of the gradient decreases as increases. As
approaches infinity (practically this corresponds to a = 100), the entire cross-section of
the concrete has come to equilibrium with the ambient humidity.

55

Figure 3.5 The simulated humidity profiles corresponding to different values

The free shrinkage strain (SH) at each point inside the concrete can be obtained
from the relative humidity using an approach similar to that suggested by Bazant and
Wittmann (1982). The relationship between shrinkage and relative humidity can be
assumed to be linear for high relative humidities (i.e., RH > 50%) (Weiss and Shah 2001).
In this case, the free shrinkage strain can be estimated as:
SH ( t ) = SH

(100% RH ( x , t ))
(100% RH S )

Eq. 3.8

where SH- is the ultimate free shrinkage of the concrete at a specific ambient relative
humidity, RHS. For the simulations in this paper, a value of SH- = -100 was assumed

56

at a RHS = 50%, however, the simulation results can be simply scaled to consider other
shrinkage values.
The degree of restraint was determined using Eq. 3.1 (c) with values of USH|RIC
and US|ROS that were determined at the mid-height of the ring from the finite element
simulations. It was observed that in the case of uniform shrinkage the simulation results
match very well with the analytical procedure (Eq. 3.6). For the cases in which a
moisture gradient develops inside the concrete (i.e., non-uniform drying), the degree of
restraint was determined for different values of the drying time parameter, . Figure 3.6
shows the results for the case of top and bottom drying. It can be seen that a small
variation in the degree of restraint occurs as drying progresses. These variations are less
than 7 % over the entire duration of drying. Similar results were obtained for concrete
experiencing circumferential drying. These results indicate that the degree of restraint is
(to a reasonable extent) independent of the moisture profile inside the concrete and as
such, the analytical procedure developed for a uniform shrinkage condition (Eq. 3.6) can
be employed for other cases of drying to estimate the degree of restraint.

57

Figure 3.6 Degree of restraint as drying progresses; obtained from finite


element simulations (top and bottom drying)

The small variation in the degree of restraint in the cases of non-uniform drying
is primarily caused by a non-uniform deformation of the steel ring throughout the height.
Figure 3.7 provides a conceptual illustration of the deformations of the steel ring and the
pressure gradient that develops at the concrete-steel interface. When the concrete shrinks
uniformly, the interfacial pressure is (almost) uniform across the height (Figure 3.7 (a)).
However, in the cases of non-uniform drying, a pressure gradient is observed along the
height due to the non-uniform shrinkage inside the concrete (Figure 3.7 (b) and 3.7 (c)).

58

Figure 3.7 Influence of shrinkage condition on the deformation of concrete and steel
rings: (a) uniform shrinkage, (b) top and bottom drying, and (c) circumferential drying

To further investigate this behavior, the interfacial pressure gradient was obtained
using finite element simulations for a concrete ring experiencing top and bottom drying.
The geometrical parameters considered in this case were as follow: ROS = RIC = 150 mm,
ROC = 225 mm, steel ring thickness = 9.38 mm, and height = 75 mm. Figure 3.8 (a)
shows the simulation results in terms of pressure profiles as drying progresses. A higher
pressure is always observed in the middle region of the steel ring. Initially, the pressure
at the top and bottom regions is zero, which suggests that the top and bottom concrete is
not in contact with steel during this period. However, the pressure increases as drying
progresses (i.e., increasing ) and approaches a more uniform distribution at higher
values of . The pressure gradient causes a non-uniform strain profile at the inner surface

59

of the steel ring (Figure 3.8 (b)). Due to stress redistribution within the thickness of the
steel, the strain profile does not exhibit a significant gradient across the height and as
such, has a minor effect on the degree of restraint. Further investigation is required to
study the effect of bending for thin and tall steel rings for which bending is expected to
be more significant.

60

(a)

(b)
Figure 3.8 The effect of non-uniform deformation of the steel ring (simulations results):
(a) pressure gradients at ROS and (b) steel ring strain at RIS

61

Finally, Figure 3.9 provides a comparison of the degree of restraint calculated


from the analytical procedure (Eq. 3.6, uniform drying) and the results obtained from
finite element simulations for the cases of non-uniform drying. The results provided here
correspond to a ring with RIC=ROS= 150 mm, ROC = 225 mm, height = 75 mm, EC= 21
GPa, and Es= 200 GPa. The simulations were performed with different values of
ranging from 0.001 to 100 to determine the degree of restraint as the drying progresses.
A good agreement is exhibited between the results obtained from the analytical procedure
and the finite element simulations, suggesting that the analytical procedure is capable of
estimating the degree of restraint regardless of the direction and uniformity of drying.

62

Figure 3.9 Comparison on the degree of restraint obtained from the analytical procedure
(Eq. 6, uniform drying) and the values obtained from finite element simulations (nonuniform drying)

3.5 A Practical Example of Geometry Selection for the Restrained Ring Test
To illustrate the usefulness of the approach described here, an example is
provided in which the geometry of a restrained ring is determined to correspond with an
actual field application and to enable reasonable stress measurement. In this example, it
is assumed that the field concrete is a floor slab that is six inches (152.4 mm) thick and
the slab experiences drying from the top and bottom surfaces. The concrete has an
ultimate free shrinkage of approximately 400 at 50% relative humidity (i.e., the
shrinkage was measured using ASTM C 157-04). The actual elastic modulus is assumed

63

to be 25 GPa, however, the effective elastic modulus can be reduced to 15 GPa (60%) to
account for the creep effect. Further, the restraint that is observed in a typical floor slab
can be calculated to be approximately on the order of 60 %.
A reasonable restrained ring test geometry could consist of a concrete ring that is
six inches tall and is allowed to dry from the top and bottom surfaces to ensure similar
moisture gradients and shrinkage rates as those in the floor. Also, a concrete ring with
circumferential drying could be used, in which case the concrete ring thickness should be
three inches (i.e., half of the depth as drying occurs from only one side). In this example,
the ring with top and bottom drying is considered (however, the same procedure can be
applied to determine the desired geometry of the ring with circumferential drying).
From the analytical procedure described earlier (Eq. 3.5 and 3.6), a reasonable
target ring geometry can be determined to approximate the condition in the floor. Figure
3.10 is a graphical representation of Eq. 3.5 and 3.6 and can be used to illustrate different
combinations of geometry parameters that satisfy a desired degree of restraint while
meeting a minimum value of the ratio st/SH to ensure proper stress calculations. To
obtain a ring geometry, first the thickness of the concrete ring should be determined by
considering the maximum size of the aggregates. In this case, if a 25 mm aggregate is
used, the concrete ring thickness can be chosen to be 125 mm (five times the maximum
aggregate size). By assuming RIC = ROS = 152.4 mm (six inches), the value of ROC will
be 277.4 mm. From a practical point of view, a standard 279.4 mm (11 inches) radius
cardboard tube form can be used to provide the outer mold for the concrete ring. This
gives an ROC = 279.4 mm, which yields an ROC/RIC = 1.83. From Figure 3.10 (a), to
achieve a 60 % degree of restraint, RIS/ROS = 0.947 is obtained, which gives a steel ring

64

thickness of 8.08 mm. A commercially available steel ring with 9.5 mm (3/8 inch)
thickness can be chosen (outer radius is six inches). This corresponds to an RIS/ROS =
0.938 and a degree of restraint of = 63.8 %, which is quite similar to 60 % and is
therefore acceptable for practical purposes (as shown as a solid point in Figure 10 (a)).
The next step is to check the value of st/SH to ensure proper stress calculations.
Using the ultimate free shrinkage of concrete (SH = 400 ) a minimum acceptable value
of st/SH = 50/400 = 0.125 is desired.

From Figure 3.10 (b), the solid point

corresponding to ROC/RIC = 1.83, and RIS/ROS= 0.947 lies above the st/SH = 0.125 curve
and, as such, the strain ratio requirement is also satisfied.

65

(a)

(b)
Figure 3.10 Graphical representation of Eq. 3.5 and 3.6 in terms of possible combinations
of geometry parameters for a desired degree of restraint (a) and an acceptable strain ratio
(b) (assumed material properties: EC = 15 GPa, ES = 200 GPa, C = 0.18, and S = 0.3)

66

Finally, the actual maximum residual stress that develops inside the concrete ring
can be calculated using the method suggested by Hossain and Weiss (2004); Eq. 3.9.
2
R 2 + R OC
R 2IC R 2IS

Actual max = st ( t )E S IC
2
2
2

R OC R IC 2R IC

Eq. 3.9

It should be noted that this equation is applicable only when shrinkage is uniform
in the radial direction (i.e., uniform drying or top and bottom drying), and an alternative
formula (as described elsewhere; Moon and Weiss, 2006) needs to be used for the case of
circumferential drying.
To summarize, the final geometry of the restrained ring test can be selected as
follows; concrete ring with 279.4 mm (11 inches) outer radius (ROC) and steel ring with
152.4 mm (6 inches) outer radius (ROS) and 9.5 mm (3/8 inch) thickness. Both rings are
6 inches high. A standard 279.4 mm (11 inches) radius cardboard tube form is used as
the outer mold for the concrete ring.

3.6 Summary and Conclusions


The following conclusions can be drawn in this chapter:
To utilize the restrained ring test method to simulate an actual field condition
requires that an appropriate specimen geometry be determined prior to performing the
experiment.

The geometry of the ring test that should be selected for a particular

application must satisfy a desired degree of restraint, match the actual drying conditions
in the field, and ensure proper strain measurements to enable accurate stress calculations.
An analytical procedure was presented to determine the degree of restraint in the
restrained ring test for concrete under uniform shrinkage (Eq. 3.6). It was shown that the

67

degree of restraint depends only on the geometry of the rings and the material properties
(namely the elastic modulus and the Poissons ratio) of the concrete and the steel. The
degree of restraint increases with increasing the thickness and elastic modulus of the
restraining ring (steel) and decreases with increasing the thickness of the concrete ring.
In addition, it was shown that to ensure proper stress calculations inside the concrete ring,
a minimum acceptable strain ratio st/SH must be met. This imposes an additional
requirement on the restrained ring geometry (Eq. 3.5).
A series of finite element simulations were performed to examine the applicability
of the analytical solution to the cases of non-uniform (i.e., top and bottom or
circumferential) drying, in which a moisture gradient develops inside the concrete ring.
It was observed that the degree of restraint shows minor changes as the drying progresses
and is not significantly influenced by the presence of a moisture gradient inside the
concrete ring. This suggests that the analytical solution that was initially developed for
uniform shrinkage is also applicable to estimate the degree of restraint in the cases of
non-uniform drying.

It should, however, be noted that the drying direction can

significantly influence the distribution of the residual stresses inside the concrete ring
(Moon and Weiss, 2006) and must not be neglected during stress calculations.

68

CHAPTER 4: INTERNAL STRESS DEVELOPMENT IN HETEROGENEOUS


SYSTEMS

4.1 Introduction
This chapter provides an overview of a research program that is currently being
conducted to develop procedures to enable information from sensors placed in a concrete
structure or pavement to be utilized efficiently as real-time feedback for updating
performance simulation models (Weiss, 2001a). While this project has many facets, one
aspect of this work is to better understand how an embedded stress sensor may be used to
quantify residual stresses in concrete.

Toward this end, a series of preliminary

computations have been performed to better understand the residual stresses that develop
in a heterogeneous composite system when one phase (the paste phase) shrinks and the
other phase (the aggregate phase) does not.

Subsequent computations have been

performed to investigate stresses that develop when both phases experience differential
movement, though they will not be discussed here.
The overall research program is being conducted to try to resolve a few key
questions that need to be answered to enable residual stress sensors to be used more
efficiently in concrete.
Numerous researchers and engineers frequently simplify concrete by assuming it
behaves like a homogeneous material by using effective material properties. While this
is commonly done, the impact of simplifications made to implement effective material

69

properties remain unclear, especially as these simplifications relate to micro-cracking,


which influences system compliance and durability.
While residual stress sensors are being developed, research is needed to address
how the external boundary conditions influence the stress fields that develop around an
aggregate with respect to directionality and magnitude.
Numerous studies are being conducted to assess how the mixture proportions of
concrete can be optimized. The majority of these studies utilize parameters that can be
easily measured like strength or free shrinkage to optimize mixture proportions. This
study takes a more fundamental look at the influence of aggregate volume, shape, bond,
and stiffness on the resulting stress development, micro-cracking, and through-cracking.
It is the goal of this research that through further studies, improved guidelines may be
available to suggest how mixtures can be designed more efficiently to improve field
performance.
This chapter is divided into four main components. The first portion of this
chapter describes the overview and motivation for the research. The second section
describes the typical computations that are used for determining the effective material
properties of a composite by using the properties of each constituent. The following two
sections of this chapter (sections three and four) discuss preliminary investigations aimed
at understanding how external restraint influences the stresses around an aggregate and
how micro-cracking develops in these systems. Section three discusses simulations and
experiments that were performed in which a single aggregate is considered in a shrinking
matrix. Section four considers a collection of hexagonal unit cells, each consisting of an

70

aggregate in a paste matrix. Finally, section five provides an overview of the preliminary
observations that can be made from this work.
It should be noted that the majority of the calculations in this paper assume the
paste and aggregate to behave as non-aging, linear-elastic materials. This approximation
enables the models to remain relatively simple. As the models are refined over time, age
and time-dependent rheological material properties will be incorporated into these
models.

4.2 Effective Material Property Calculations


Concrete shrinks in response to drying, self-desiccation, and chemical reactions.
The paste component is generally recognized as the component responsible for the
shrinkage, while the aggregate component is frequently thought of as an inert-filler that
reduces the overall shrinkage of the concrete system. Several researchers in the 1950s
discussed how the shrinkage of a concrete (Concrete) could be described as a fraction of
the shrinkage of the paste (Carlson, 1937; Pickett, 1956; LHermite, 1960). Pickett
developed an expression that is frequently used to describe the relationship between the
shrinkage of the paste and the shrinkage of the concrete (Eq. 4.1).

Concrete = Paste (1 VAgg ) n

Eq. 4.1

where Paste is the shrinkage of the paste, VAgg is the volume fraction of the aggregate,
and n is a constant for a particular system.

To arrive at this formulation, Pickett

considered the effect of spherical aggregate particles in a concrete composite. Pickett


computed the restraining effect of aggregates by assuming that the aggregate and the

71

concrete act elastically. Pickett assumed that the exponent n would be related to the
relative elastic properties of the aggregate and the concrete. He developed a theoretical
expression that showed that the value of n could range from 0 to approximately 2. This
theoretical expression is somewhat difficult to use appropriately, however, as it requires a
measure of the elastic modulus of the concrete composite.

Pickett experimentally

determined that the exponent n would have a value of 1.7 for mortars made using Ottawa
sand (ASTM C-190).
LHermite (1960) wrote an extensive summary of research on shrinkage and
summarized the work of Dutron (Dutron, 1934) where Eq. 4.1 had been written in a
slightly different form. In re-analyzing the results of Dutron using Eq. 4.1, LHermite
found that the exponent n ranged between 1.2 and 1.7 for normal strength mortars and
concretes made using different aggregate. Recently, simulations were performed (using
the planar geometry as described in Section 4.4) where the exponent, n was related to the
elastic modulus of the aggregates (Figure 4.1). A hyperbolic equation (Eq. 4.2) was fit to
the data shown in Figure 4.1

n = n

1
E
1 + C1 Paste
E Agg

Eq. 4.2

where n refers to the value of n for an infinitely stiff aggregate, which can be taken as
1.405, and C1 is a constant that can be taken as 0.25. EPaste and EAgg are the elastic
moduli of paste and aggregate, respectively. While Eq. 4.2 may need to be modified to
account for low paste moduli, different aggregate geometries, or differences in the
Poissons ratio between the two materials, it appears possible to use this expression to

72

estimate the free shrinkage of a concrete composite. It should be noted, however, that
this simulation was performed for a planar geometry (assuming plane stress) and some
differences may exist for a truly three-dimensional composite geometry.
This suggests that the overall shrinkage of a concrete can be reasonably estimated
using Eq. 4.1. Neville (1996) points out that the aggregate restrains the shrinkage and
this restraint causes residual stress to develop in the paste. However, while many authors
use Eq. 4.1 (Mindess et al., 1996; Mehta and Monteiro, 1986) to approximate the overall
shrinkage, others (Swazye, 1960) have cautioned against relying on such approximations.
An objection that has been raised to the use of Eq. 1 may be attributed in part to the role
of residual stresses that result in non-linearities associated with creep and microcracking
in the system.
To measure the residual stresses that develop in a composite system, Dela and
Stang (Dela and Stang, 2000) introduced an aggregate sensor to measure the pressure
on a spherical embedded inclusion inside a cement paste. They concluded that the effects
of stress relaxation were substantial at very early ages while the subsequent stress
development could lead to cracking. It has also previously been suggested that the region
around an aggregate may be susceptible to cracking as somehow attempted to measure
the extent of microcracking in cement mortars using acoustic emission measurements
(Pease et al., 2003b, 2004).
Figure 4.2 shows the results of early-age measurements in terms of cumulative
acoustic activity in externally-unrestrained sealed mortar samples with different water-tocement ratios (w/c) (Pease et al. 2003b). Substantial acoustic activity occurs in the
specimens with the lower w/c at early ages. The specimens with a w/c of 0.30 showed

73

the greatest number of acoustic events, followed by the specimens with a w/c of 0.35,
0.40, and 0.50, respectively. It was hypothesized that the lower w/c mixtures undergo
more autogenous shrinkage and, as a result, they are more likely to experience higher
residual stresses generated by the internal restraint from the aggregates. This implies that
concrete made using a lower w/c would be more prone to microcracking due to the
internal restraint of aggregates against autogenous shrinkage. The current paper builds
on these observations and describes a series of experiments (Pease et al., 2003b) in which
the role of aggregate inclusions is investigated by using different model systems.
To understand how micro-cracking occurs in low w/c pastes, Pease et al. (2003b)
and Moon et al., (2004) used a low w/c cylindrical paste specimen. A steel rod was
placed in the center of the paste specimen.

This system was used to simulate the

shrinkage of paste around an aggregate. Specimens were prepared with different pasteto-aggregate diameter ratios. An elastic shrink-fit theory (Timoshenko and Goodier,
1970) was used to try to interpret the results from the test. It was concluded that the
maximum residual stress that develops in the paste could be computed using Eq. 4.3
(after correcting the elastic modulus of paste to account for creep):

74

Shrinkage Exponent, n

1.50
1.25
1.00
0.75
0.50
0.25
0.00
0.01

0.1

10

100

1000

Ratio of Aggregate and


Paste Stiffnesses (EAgg/EPaste)
Figure 4.1 Shrinkage exponent n as a function of the aggregate to paste stiffness ratio

Cumulative Events (hits)

350
0.30 W/C
0.35 W/C
0.40 W/C
0.50 W/C

300
250
200
150
100
50
0
0

12

16

20

24

Age of Specimen (hrs)


Figure 4.2 Acoustic activity in mortars at early ages (50% aggregate volume)

75

= Paste

2
R OP
+ R 2OA
2
2
R OP
R OA
E Paste (t )
2
2
(1 Agg ) E Paste (t ) + Paste + R OP2 + R OA
2
E Agg
R OP R OA

Eq. 4.3

where corresponds to Poissons ratios of paste (denoted with a subscript Paste) and
aggregate (denoted with a subscript Agg), and R corresponds to the outer radius of the
aggregate (denoted with a subscript OA) and the paste cylinder (denoted with a subscript
OP), respectively.
Eq. 4.3 was used to estimate the maximum stress level that could develop around
the aggregate and to compare the cracking behavior of different specimens. As the
aggregate volume increased (i.e., lower ROP/ROA) there was an increase in both the
average residual stress level and the acoustic activity. As the paste radius decreased, the
potential for through-cracking increased. Through-cracking was observed to correspond
to a sudden rise in acoustic energy in these specimens.
Previous research assumed that acoustic activity is synonymous with the
development of microcracking. To measure this more directly, Lura et al. (2005) are
investigating the cracking around an idealized aggregate using gallium impregnation.
The main benefit of gallium impregnation is that it could be performed without inducing
additional cracking into the system; thereby, cracks are imaged as they were in the
unperturbed specimen, before damage from preparation could occur.
As mentioned earlier, a composite system can be described by equivalent material
properties, such as the equivalent elastic modulus. Hansen (1965) proposed an equation

76

for calculating the equivalent elastic modulus of a spherical composite, which contains a
spherical inclusion at the center of a sphere of the paste matrix (Eq. 4.4).

(1 VAgg ) E Paste + (1 + VAgg ) E Agg


E composite =
E Paste
(
1
+
V
)

E
+
(
1

V
)

Agg
Paste
Agg
Agg

Eq. 4.4

To provide a simplified solution, Hansen assumed that the Poissons ratio was the
same for each phase (0.20). Such equivalent material properties can be used directly to
estimate the residual stress that would develop when concrete is completely restrained
from shrinking freely by multiplying Eq. 4.1 and Eq. 4.4. This will be discussed in
greater detail later.

4.3 Single Aggregate Prism Systems


As previously discussed, there are questions as to how the external boundary
conditions influence the residual stress fields that develop around an aggregate when it is
suspended in a paste matrix. This section simulated a geometry in which a single
aggregate was placed inside a prism of paste. The paste was allowed to shrink and the
stresses were computed for an unrestrained and a horizontally-restrained specimen.
The single aggregate prism specimens used in this investigation are illustrated in
Figure 4.3. The length (L) of the specimen is five times its width (H). The specimens
contain one aggregate in the center, with a diameter (D=2ROA) that was varied from 0.1
to 0.2, 0.4, 0.5, 0.6, 0.8, and 0.9 times of the height (H) of the prism. Two external

77

boundary conditions were considered in this analysis: a) unrestrained and b) horizontallyrestrained, where the ends of the specimen were completely fixed so as to not permit any
movement in the x-direction, though movement in the y-direction was permitted.

L
D

H
Single Aggregate - Unrestrained

(a)

Single Aggregate - Restrained

(b)

Figure 4.3 An illustration of the two geometries simulated: a) an unrestrained single


aggregate system and b) a restrained single aggregate system

These specimens were simulated using finite element analysis (FEA) in the
commercial program ANSYS.

The model systems were meshed using quadratic

rectangular eight-node elements and analyzed using plane-stress approximations. The


effect of autogenous shrinkage was simulated using a temperature substitution analogy
(Moon et al., 2004).

A uniform temperature load was applied, while the thermal

expansion coefficient of the cement paste was set to 110-6 //C and the thermal
expansion coefficient of the aggregate was set to zero. The autogenous shrinkage strain
was assumed to be -100 microstrain ( which was applied using a temperature load (T
= -100 C). It should be noted that since linear elasticity was assumed, the stresses can be
scaled proportionally to reflect an increase or decrease in autogenous shrinkage. The

78

paste was assumed to have a modulus of 20 GPa and a Poissons ratio of 0.20, while the
elastic properties of the aggregate were 200 GPa and 0.30, respectively.
It should be noted that this solution does not consider the effect of creep or
changing the elastic properties of the paste. A perfect-bond between the aggregate and
the paste is assumed in the simulations unless it is specifically noted as otherwise. The
preliminary results showed that the bond condition has a minor effect on the system with
unrestrained boundary conditions, while it has a substantial effect with restrained
boundary conditions. A further discussion with respect to the bonding effect will be
presented later.
Figure 4.4 shows stresses that developed along the x-axis and along the y-axis for
two different external boundary conditions when the diameter of the aggregate (2ROA) is
0.1H. In the case of the unrestrained specimens, the residual stress level decreases as the
distance from the aggregate increases (Figure 4.4 (a)). This solution is similar to the
stresses that develop in the ring test (Pease et al., 2003b; Weiss, 1999) although a slight
correction is needed in the case of the larger aggregates to account for the top and bottom
edges of the prism specimen. Very similar stresses develop in the x and y directions. For
this particular geometry, the bond between the aggregate and the matrix has little
influence on the developed stresses.

Numerical simulations have shown that the

difference between the bonded and unbonded case is less than 7%. In pure cement paste
specimens (i.e., no aggregate), the specimen will not experience any internal stress
development under the unrestrained boundary conditions. However, when aggregates are
present, internal stresses that develop at the interface between the two phases could result
in internal microcracking.

79

In the externally restrained specimen (fully restrained in one direction) a stress


gradient develops along the x-axis (i.e., for element A in Figure 4.4) that is similar to the
stress gradient in the unrestrained specimen.

However, the externally restrained

specimen develops a stress gradient along the y-axis (i.e., for element B in Figure 4.4)
that is much greater than the stresses that develop in the unrestrained specimen. This is
not unexpected; however, if a pure paste specimen were restrained, the residual stress in
elements along the y-axis would be constant and equal to 2 MPa. As a result, the
maximum stresses that develop in Figure 4.4 (b) are approximately 14% higher than the
specimen without aggregate at a small distance from the aggregate surface and similar to
the specimen without aggregate at a sufficient distance away from the aggregate.
The results indicate that the inclusion of aggregate may increase the cracking
potential of a specimen by combining the influence of internal restraint with the stresses
that develop from the external restraint.

In the case of the unrestrained boundary

conditions, the stress level reduces quickly by moving away from the aggregate.
However, in the case of the specimen with externally restrained boundary conditions, a
high stress level is maintained even when moving away from the aggregate (in the ydirection). It can be imagined that if a small crack develops in the unrestrained specimen,
this crack could reduce a substantial amount of the stored residual stress. As a result, a
microcrack could initiate, yet there may not be enough stored energy to cause this crack
to propagate across the specimen. However, the externally restrained specimen stores
substantial energy, which may not substantially decrease as the crack develops.
Therefore, the externally restrained specimen will have the higher potential of throughcracking when the crack initiates.

80

Restrained Single
Aggregate Specimen

Unrestrained Single
Aggregate Specimen
2

A
B

Stress (MPa)

Stress (MPa)

0
0

10

20
B

-1

A
B

0
0

30

10

20
B

-1

DOA

DOA

A
Y

Y
-2

Distance from an aggregate (mm)

(a)

30

-2

Distance from an aggregate (mm)

(b)

Figure 4.4 Stress development in an (a) unrestrained single aggregate specimen and b) the
externally restrained single aggregate specimen

As previously discussed, the bond between the aggregate and paste has an
important role on the stress development in a composite system when the composite is
externally restrained. Figure 4 (b) shows that the maximum stress that develops along the
y-axis (element B) is not at the aggregate-paste interface. This may be explained by the
fact that the actual maximum stress develops along a diagonal (approximately 55o from
the y-axis for the geometry shown) when the aggregate is perfectly bonded with cement
paste (Figure 4.5 (a)), whereas it develops along the y-axis when the aggregate is

81

perfectly unbonded (Figure 4.5 (b)). Assuming that the aggregate and paste are initially
perfectly bonded, it could be expected that the tensile stresses that develop at the pasteaggregate interface could lead to microcracking (debonding).

Therefore, it can be

expected that the initial debonding initiates at the interface at approximately 55 (for the
geometry shown) from the y-axis and grows along the interface between the paste and
aggregate until the aggregate becomes fully debonded from the paste.

Externally Restrained
Boundary Condition

Externally Unrestrained
Boundary Condition

Max. Stress
Development

Agg.

(a)

Agg.

Agg.

(b)

(c)

Figure 4.5 Stress localization for different boundary conditions: (a) perfectly bonded
aggregate for the externally restrained specimen, (b) perfect unbonded aggregate for the
externally restrained specimen, and (c) perfectly bonded/unbonded aggregate for the
externally unrestrained specimen

This debonding causes a redistribution of stresses. Figure 4.6 shows the residual
stresses along the y-axis for different bond conditions (externally restrained specimen in
horizontal direction). In a perfectly bonded condition, a lower stress is observed at the
aggregate-paste interface, which can be attributed to the fact that the aggregate is

82

participating in transferring stress through the system. The local variations in stress arise
due to the differences in the elastic properties of the aggregate and the paste. When the
aggregate is unbonded, tensile stresses can no longer be transferred through the
aggregate. This results in the development of a maximum stress at the aggregate-paste
interface (element B). Figure 4.6 also shows the stress development around a spherical
air void, which can be considered as an aggregate with zero stiffness.

The main

difference between an air void and an unbonded aggregate is the ability of the unbonded
aggregate to transfer compressive stress. In a system containing an unbonded aggregate,
a radial pressure develops at the top and bottom of the aggregate as the paste shrinks.
This effectively implies that the aggregate wedges the void open. In the case of an air
void, no such stresses develop since the deformation of the paste in the y-axis direction is
not restrained by the presence of aggregate. This explains the slight difference between
the stress development in the paste around an air void and around an unbonded aggregate
inclusion (Figure 4.6).

83

6
Perfectly Bonded
Perfectly Unbonded
Air Void
No Agg.

Stress (MPa)

5
4
3
2
1
0

4
8
12
Distance from the Agg. (mm)

Figure 4.6 Stress development along the Y-axis of a restrained prism specimen over an
aggregate (restrained boundary condition)

To provide a physical model that is similar to the simulations, a prism specimen


was restrained in one direction with a single aggregate. It is believed that this physical
model will be able to be used to calibrate material parameters for future modeling
developments. The physical model consisted of a prismatic specimen that had a length of
250 mm, a width of 50 mm, and a height of 25 mm. This specimen geometry is similar to
the passive dog-bone restraint frame described in the literature (Chariton and Weiss,
2002). A cement paste specimen was placed in the frame with a w/c of 0.3. Specific
details on the cement and the free shrinkage of this cement paste are described in the
literature (Pease et al., 2003b). In addition to the cement paste, the specimen had an

84

instrumented aggregate that was placed in the center of the prism. The instrumented
aggregate consisted of a thin-walled copper cylinder with an outer diameter of 15.8 mm,
a height of 25 mm, and a wall thickness of 0.6 mm. Two strain gages were placed on the
inner surface of the copper cylinder at 90 from one another to measure the strain that
develops on the aggregate in different directions, as illustrated in Figure 4.7.
The specimen was sealed to prevent drying for the duration of the experiment. As
illustrated by Figure 4.7, the instrumented aggregate (copper ring) recorded a strain
development caused by the autogenous shrinkage of the cement paste. These strains were
measured by the strain gages and recorded by a Strainsmart acquisition system every five
minutes.
The results show an initial compressive strain recorded by both strain gages.
Since the paste is still in a plastic phase, it is highly unlikely that these strains are
attributed to a stress development in the system. Rather these strains could be thought to
be due to a slight temperature rise in the system (approximately 2C). After the time of
setting (approximately seven hours) a tensile strain is measured in the direction of the yaxis (B) and a compression strain in the direction of the x-axis (A) is observed. This can
be attributed to the development of a radial pressure at point B accompanied by
aggregate-paste debonding at point A. The strain increased over time due to the increase
in autogenous shrinkage. At an age of 27 hours, a visible crack was observed from the
aggregate to one of the specimen edges. It should be noted that the strain in the copper
ring does not drop to zero after cracking, presumably due to the bond between the ring
and paste and continued stress transfer from the mortar to the copper ring.

85

200
Position A
Position B

Strain ()

100
0
-100

B
A

-200

Copper Ring
(with strain gages)

-300
-400
0

10

20

30

40

Time (Hours)
(a)

Cummulative Acoustic
Emission Energy (nVs)

0.6

Jump in
Copper Ring
(with strain gages) Acoustic
Activity
to 10.3
nVS

0.4

0.2

0
0

10

20

30

40

Time (Hours)
(b)
Figure 4.7 Experimental data from a restrained specimen containing an Instrumented
Aggregate: a) strain development and b) acoustic activity

86

In addition to measuring strain development, acoustic emission was used to


estimate the extent of cracking that occurred in the specimen.

Acoustic emission

describes a class of testing that relies on the use of piezo-electric transducers to measure
the vibration (or acoustic activity) that occurs when a disturbance (i.e., crack) occurs in a
material. This disturbance results in the release of energy and the propagation of an
elastic wave. For the sake of brevity, the reader is referred to existing literature on the
specific details of the acoustic emission equipment and the testing approach employed
herein (Kim and Weiss, 2003).
Figure 4.7 (b) shows the cumulative acoustic emission energy recorded during
this experiment. The first acoustic event was recorded shortly after setting. During the
initial period after setting, microcracking is expected at the aggregate-paste interface
which corresponds to a continuous release of acoustic energy (Pease et al., 2003b).
When the visible through-crack propagated, a rapid increase of the acoustic energy was
observed.

4.4 Multiple Aggregate Systems


A model was developed using a combination of hexagonal unit cells (Figure 4.8
(a)), each consisting of a cylindrical aggregate particle surrounded by a hexagon of the
paste matrix. This model system was used to begin to assess the behavior of a multiaggregate system.

The model system consisted of approximately 16 unit cells.

Simulation results for residual stress were acquired from the center hexagonal unit cell to
avoid boundary effects at the edges of the specimen. The length of each side of the
hexagon (lHex) was fixed as 16.7 % of the length of the system while the size of the

87

aggregates was varied to investigate the effect of aggregate volume fraction on stress
development. For this purpose, simulations were performed by varying the aggregate
radius (ROA) from 0.1ROP, to 0.2ROP, 0.4ROP, 0.5ROP, 0.6ROP, 0.8ROP, and 0.9ROP where
ROP is the equivalent radius of hexagon corresponding to the same cross-sectional area as
a cylinder (Figure 4.8 (a)) (Pease et al. 2003b). The elastic modulus of the cement paste
was assumed to be 20 GPa while the elastic modulus of aggregate was varied to observe
the effect of aggregate stiffness on stress development. The Poissons ratio of the paste
and aggregate was assumed to be 0.2 and 0.3 respectively. Two external boundary
conditions, unrestrained and horizontally restrained (Figure 4.8 (b)) were applied in the
simulations.
First, a series of simulations were performed to obtain the equivalent elastic
modulus of the composite system. A pseudo displacement load of -100 was applied to
the horizontal direction and the reaction force was obtained. The overall average section
stress and average section strain were used, along with Hookes law, to determine the
equivalent elastic modulus of the composite.

Figure 4.9 show the results of the

simulations for different aggregate volume fractions.

Figure 4.9 (a) provides a

comparison between the simulation results and the results obtained by using a series and
a parallel composite model, as well as the model used by Hansen (1965) (Eq. 4.4). The
results shown correspond to a case when the ratio of Eagg to Epaste is 10. Figure 4.9 (b)
provides a comparison between the simulation results (shown as data points) and the
results obtained from the Hansens model (shown as lines) for different Eagg/Epaste values.
It can be observed in both Figures that the simulation results are in good agreement with

88

the results from the Hansens model despite the differences in the Poissons ratio and the
differences in the 2d versus 3d model assumptions.

y
Unrestrained
Hexagonal Unit Cell Model

x
Agg.

lHex
ROA
ROP

Restrained
Single Unit Cell

Equivalent Cylinder

(a)

(b)

Figure 4.8 Multiple aggregate system composed of unit cell matrix

89

Equivalent Ecomposite (GPa)

200

EAgg / EPaste = 10
Parallel Model
Series Model
Hansen's Model
Simulation

160

120

80

40

0
0

20

40

60

80

100

80

100

Vol. of Agg. (%)


(a)

Equivalent Ecomposite ( GPa)

200
Eagg/EPaste
10
5
2
Simulation
1
0.5
0.1
Hansen's
Model

160

120

80

40

0
0

20

40

60

Vol. of Agg. (%)


(b)
Figure 4.9 Comparisons of equivalent elastic modulus of composites

90

A second series of simulations were performed to obtain the internal stress


development inside the composite system as the cement paste shrinks around the
aggregates. The autogenous shrinkage of the cement paste was assumed to be -100
microstrain (). Figure 4.10 shows the maximum stresses that develop in an externally
unrestrained system as a function of the aggregate volume fraction. Figure 4.10 (a)
shows the maximum principal stress in the cement paste for different values of Eagg/Epaste,
ranging from 0.1 to 10. A higher level of stress is observed as the volume fraction and
the stiffness of aggregate increase. Figure 4.10 (b) provides a comparison between the
simulation results and the maximum residual stress of the paste determined using Eq. 4.3.
A good agreement between the two is observed for composites with an aggregate volume
less than 60 percent. However, at higher volume fractions, the two methods predict stress
levels that are slightly different, which is mainly due to the fact that Eq. 3 is derived for a
single aggregate system (a cylindrical paste shrinking around a cylindrical aggregate).
As the volume fraction of aggregate increases, the distance between the aggregates
becomes smaller and the aggregates begin influencing one another.

91

Max. Principal Stress(MPa)

3
EAgg/EPaste (Simulation)
10
5
2

1
0.5
0.1

Externally Unrestrained
0
0

20

40

60

80

100

Vol. of Agg. (%)


(a)
2.2
Externally Unrestrained

Stress (MPa)

1.8

1.6
Simulation Eq. 3

EAgg 10
EPaste 5
2

1.4

1.2
0

20

40

60

80

100

Vol. of Agg. (%)


(b)
Figure 4.10 Maximum stress development in an externally unrestrained composite

92

Figure 4.11 show the maximum residual stresses that develop in the composite
system when it is horizontally restrained. In Figure 4.11 (a), the maximum principal
stress that develops in the paste is shown as a function of the aggregate volume fraction.
It is observed that for a value of Eagg/Epaste higher than 2, the maximum stress level is
relatively independent of the aggregate volume fraction. However, for lower stiffness
aggregates, the maximum stress initially decreases with increasing the volume fraction of
aggregate and reaches a minimum around a 65% volume fraction. After this point, a
slight increase in the stress level is observed as the volume of aggregates increases.
Alternatively, the maximum residual stress in concrete can be computed by
considering the system as a homogenous material with equivalent properties. In this case,
the equivalent shrinkage strain can be determined by applying Eq. 4.1 and 4.2, while Eq.
4.4 could determine the equivalent elastic modulus of the composite. The equivalent
maximum residual stress is determined by the application of Hookes law. Figure 4.11
(b) shows the results for composites with different volume fractions of aggregate. As the
aggregate volume increases, the equivalent shrinkage strain (calculated using Eq. 4.1 and
4.2) decreases, which leads to a decrease in the equivalent maximum residual stress.
However, such analysis does not consider the residual stresses that develop due to the
internal restraint provided by the aggregates. For comparison, the maximum principal
stress in the composite obtained from the FEA simulations is also included in Figure
4.11(b). These results, on the other hand, do not suggest that the residual stress level in
the composite decreases significantly with increasing the volume fraction of aggregate.

93

Max. Principal Stress (MPa)

6
Externally Restrained (Simulation)

EAgg / EPaste
10
5
2

0
0

20

40

60

1
0.5
0.1
80

100

Vol. of Agg. (%)


(a)
6

Stress (MPa)

Externally Restrained
Internal Equivalent
EAgg 10
EPaste 5
2

0
0

20

40

60

80

100

Vol. of Agg. (%)


(b)
Figure 4.11 Maximum stress development in a composite externally restrained
in one direction

94

While this may appear to contradict many field observations, such as those by
Darwin et al. (2004), who noted that visible cracking in the field was substantially
reduced in materials with a lower paste volume, it should be noted that these observations
do not consider the potential contribution of microcracking or interfacial cracking that
can be expected to occur in a heterogeneous composite system.
Up to this point, this chapter has focused on assessing the residual stress
development in the paste matrix (and aggregate). It is also necessary to consider the
effect of the bond condition between aggregate and cement paste to better understand the
potential for microcracking at the interface between aggregate and cement paste. As
previously discussed, the bond condition has little effect on the internal stress
development in a composite system for the unrestrained boundary condition because the
matrix shrinks around the aggregates, subjecting them to an almost uniform compressive
pressure along the interface. However, in the case of the restrained boundary condition,
the bond condition and the bond strength become much more significant parameters. The
influence of bond strength was investigated using finite element analysis. A composite
system consisting of three phases (cement paste, aggregate, and bond elements) was
prepared with an aggregate volume fraction of 15% (Figure 4.12). The time-dependent
values of autogenous strain, elastic modulus, and fracture energy of cement paste were
used in these simulations. For the sake of brevity, the exact time dependent functions are
not provided, however, they describe the behavior of a paste with a w/c of 0.30, which is
similar to that measured by Pease et al. (2004). For the trials described in this paper, the
bond elements were assigned stiffness and fracture energy that were 10% of the values
for cement paste. Further work is needed to characterize the bond stiffness and strength

95

more thoroughly for further simulations. To determine whether interfacial cracking


occurred, a fracture energy analysis was performed to determine whether the stored strain
energy in each element exceeded the fracture energy. When the stored strain energy
exceeded the fracture energy, the stiffness and fracture energy of those elements was
decreased to 10% of the previous values and the simulation was re-run until a stable
geometric configuration was obtained. This simulation technique is typically used when
a model has high non-linearity, such as cracking. The elastic modulus of aggregate that
was used was 100 GPa and the cracking of the aggregate was not considered in this study
in an effort to focus on observing the cracking tendency of cement paste with low bond
strength.
For the analysis shown in Figures 4.12 and 4.13, the shrinkage strain was applied
incrementally and the stress development and cracking was monitored. Figure 4.12 (a)
shows the initial status of the microcracking, where crack coalescence can be observed
during loading as the black-colored regions. These debonded regions developed along
the edges of the aggregates under the horizontally restrained boundary condition. The
debonding between the aggregate and the paste increased until the debonding reached the
y-axis position (i.e., the top and bottom of the aggregate as shown in Figure 4.12 (b)) at
which time a vertical through-crack developed (Figure 4.12 (b)).

96

(a)

(b)

Figure 4.12 Cracking tendency of a composite (horizontally restrained boundary


conditions)

It was previously discussed that the distance between the aggregates has an
important role in the internal stress development and cracking tendencies. However, it is
impossible to think that the simple hexagonal unit cell composite model can completely
represent a complex system like a concrete composite. In real cementitious composites,
the shape of the aggregate, the size distribution of the aggregate, and the spacing between
the aggregates would influence the internal stress development (Jones and Kaplan 1957).
Therefore, additional modeling approaches are needed.

97

4.5. Summary
This chapter describes the effect of internal and external restraint on residual
stress development in a concrete composite. Three model systems were evaluated
including 1) a single aggregate, 2) a composition of hexagonal unit cells, and 3) a real
composite image system. The following observations can be made:

The increase in aggregate volume reduces the overall free shrinkage of a


cementitious composite.

The addition of aggregate results in the generation of internal residual


stresses.

The bond condition between the aggregate and the cement paste has a small
role on stress development when a composite is not externally restrained.
The bond condition becomes more critical for describing the behavior of the
system when an external restraint is provided.

In the case of the externally unrestrained boundary condition, it was observed


that higher internal stresses develop with a higher volume fraction of
aggregate and higher stiffness of aggregate.

In the case of the externally restrained (in one direction) boundary condition,
the maximum residual stress level is not observed to vary significantly with
increasing the volume fraction of aggregate.

The local residual stresses that develop in a composite are higher than the
equivalent stresses that are computed using the equivalent shrinkage strain
and equivalent elastic modulus of a composite. This is due to the fact that the

98

latter technique does not consider the stresses generated by the internal
restraint of the aggregates, and indicates that it is possible to underestimate
the microcracking and cracking potential of concrete if estimation is
performed only using equivalent parameters.

99

CHAPTER 5: ANALYSIS PROCEDURES FOR ASSESSING THE SHRINKAGE AND


CRACKING IN A CEMENTITIOUS COMPOSITE ON THE MESO-SCALE

5.1 Introduction
Concrete changes volume in response to chemical reactions, moisture changes,
and temperature variations. Many documents highlight the importance of considering the
influence of volume change on the potential for cracking (ACI 209R-92; Kosmatka et al.,
2003), however, the majority of these documents and simulation models (Bernard et. al.,
2002; Foster, 2000; Weiss et al., 1998) treat the concrete as a homogeneous material and
focus only on the development of large, visible cracks.

Currently, few engineers

explicitly consider the heterogeneous nature of concrete in designing for shrinkage


cracking. Researchers (Grzybowski and Shah, 1990; Bisschop and van Mier, 2001; Kim
and Weiss, 2003; Moon et al., 2005) have shown that the combination of the external
restraint from the structure surrounding the concrete and the internal restraint provided by
the aggregate can lead to the development of microcracking even when visible cracking
is not predicted or observed. These microcracks may play a key role in relaxing stresses
(Pease et al., 2005), increasing fluid transport (Yang et al., 2005) and serving as an
initiation point for visible crack formation (Chariton et al., 2002). Chapter 4 showed that
the use of the conventional analysis approach of assuming that concrete acts as a
homogeneous material may not provide sufficient information about the microcracking
that would be experienced in the composite.

Chapter 4 introduced the concept of

100

assessing concrete as a heterogeneous system. Finite element simulations were proposed


that consider idealized meso-scale composite systems consisting of hexagonal unit cells.
While these models begin to illustrate the stress distribution in a composite due to the
stress concentration provided by the material heterogeneity, they do not provide enough
information to completely analyze the microcracking behavior of real heterogeneous
composite systems because aggregates do not have a single shape and are not distributed
uniformly.
This chapter describes an initial research effort to utilize a simulation technique to
assess the potential for microcrack development in concrete. The behavior of realistic
two-dimensional concrete composites system was simulated using an Object-Oriented
Finite Element (OOF ver. 1.1.21) code which was developed by the National Institute of
Standards and Technology (NIST, Langer et al., 2001). OOF is compatible with UNIX
or Linux operating systems and it provides an easy to use interface which enables users to
define each phase in a composite and to mesh elements of a composite system from an
optical scan of an surface of the material. The OOF program uses portable pixmap image
files that can be obtained by scanning the surface of real composites (actually, the
PPM2OOF code (ver. 1.1.29) which is the complementary code of OOF, provides this
interface). This approach is only applicable when each phase in a composite has a
different color (or color intensity) because the process of defining each phase is
performed based on the color information of each phase. Therefore, the color images of a
concrete (or mortar) need to be obtained where the concrete has paste and aggregates that
are different colors. Under these conditions, the OOF program can be used to perform
finite element simulations on multiple-phase meso-scale structures rather easily.

101

This chapter describes the simulation procedure in three sections.

The first

section describes the preprocessing step, indicating how the image files are obtained from
an actual concrete sample, how the colors of the paste and aggregate are used to define
each phase, and how this information can be used to develop a finite element mesh for
each sample. The second section of this chapter describes the finite element simulation
process, and the third section describes the postprocessing procedure indicating how
image analysis is performed for data analysis using color contour images of stress or
energy. The results and discussion sections of this chapter discuss how the results of the
simulations can be used for assessing the shrinkage cracking behavior of a heterogeneous
system.

5.2 Overview of the Approach


The importance of performing a meso-scale analysis to describe the cracking
behavior of concrete composites is discussed in this chapter. A realistic two-dimensional
composite system of aggregates in a cement paste matrix is simulated using OOF code.
The sequence of the meso-scale OOF simulation procedures is described in the flowing
chart (Figure 5.1), and the section thereafter describes the preprocessing, finite element
simulation, and postprocessing in detail.

102

OOF SIMULATION ON THE MESO-SCALE


(Object-Oriented Finite Element Simulation)
1. PREPROCESSING
1)
2)
3)
4)
5)
6)

Acquisition of concrete specimen


Surface preparation (saw-cut and polish)
Chemical treatment (phenolphthalein)
Acquisition of 2-D image (scanning)
Re-coloring
Meshing

2. FINITE ELEMENT SIMULATION


1)
2)
3)
4)

Modeling of shrinkage strain of paste


Determination of fracture criterion
Boundary condition
Load

3. POSTPROCESSING
1) Acquisition of images of stress/energy
distributions
2) Image analysis

DATA ANALYSIS

Figure 5.1 Flow chart of OOF simulation on the meso-scale

5.3 Analysis Procedures


The analysis procedures of OOF simulation consists of three sections:
preprocessing, finite element simulation, and postprocessing. While a more detailed
description of each section will be described later in this chapter, a brief description of
each process is provided below.

103

Preprocessing: The preprocessing procedure described in this chapter was used to obtain
an image of the concrete section that will be used for finite element simulation.
The preprocessing procedure consisted of obtaining a high-resolution color image.
The color in the image was used to define the phases (i.e., the aggregate and paste
matrix) of the composite. To obtain a quality image, the concrete or mortar
specimen was saw-cut and polished. The polished surface was stained using
phenolphthalein to change the color of the paste phase to pink, thereby enabling
each phase to be more clearly identified. The stained surface was then scanned
using a flat-bed type scanner. The scanned image was saved as a TIFF format file
that records the initial color of each pixel. The saved image file was re-colored
using a color histogram to identify similar phase areas (i.e., areas of similar color)
that can be converted to a grayscale image (i.e., black, gray, and white) to
represent any phase (aggregate, interfacial zone, or paste) with the assistance of
Adobe Photoshop. In this research, each phase was defined as having the same
material properties throughout the phase. The re-colored image was then saved as
a portable pixmap image file and transferred to PPM2OOF code (complementary
program of OOF code for pre-processing) for meshing.

The mesh was

successively refined until the boundaries between the phases had a shape that was
reasonably similar to the original image. The refinement of the mesh at the
boundaries was performed using the PPM2OOF code (Carter et al. 2000). In
summary, the preprocessing consisted of four components; a) concrete specimen
preparation, b) concrete specimen scanning, c) image analysis (re-coloring) of the
scanned image, and d) meshing.

104

Finite Element Simulation: This section begins by describing how the shrinkage of
concrete was simulated using the technique of thermal expansion substitution,
which is previously mentioned in Chapter 2.

For a composite system, a

coefficient of thermal expansion of zero is applied to the aggregate and a


coefficient of thermal expansion of 10-6 /oC is applied to the paste phase. A
negative temperature is applied to a meshed composite system, which introduce a
shrinkage strain in the paste phase (here, a change in temperature of -1 oC
represents a -1 shrinkage strain). Two different boundary conditions were
considered in this study.

The unrestrained specimen corresponds to a free

shrinkage specimen while the restrained specimen considers restraint in one


direction (i.e., the x-direction) similar to what may be expected in a bridge deck
or pavement. The magnitude of the shrinkage strain of paste was successively
increased until a meshed composite failed (externally restrained boundary
condition) or up to a designated value (free boundary condition). For cracking
analysis, a brittle failure criterion was used. Further discussion on the failure
criterion is provided later in this chapter.
Postprocessing: This section describes the developing methods used to assess the results
of the finite element analyses. In addition to using the stresses obtained directly
at the nodes from the finite element analysis, image analysis was performed using
the color contour images obtained from the OOF simulations, where each color
represents a specific level of stress or strain energy density. The area fraction of
each color level was obtained using Image-Pro PlusTM. This section discusses

105

how the information that was obtained from the results of the OOF simulations
were analyzed.
At the end of this chapter, the differences between the conventional analysis
approach which assumes that the concrete behaves as a homogeneous material and the
meso-scale approach (OOF simulation) which considers concrete as heterogeneous
material will be discussed. The application of the meso-scale approach for cracking
analysis of composites will also be illustrated.

5.4 Preprocessing: Acquiring the Image from a Realistic Specimens

5.4.1 Specimen Preparation


The simulations that were performed in this research used two-dimensional images. To
obtain a realistic cementitious composite, actual two-dimensional corss-sectional surfaces
of concrete (mortar) were prepared. The cross-sectional surface of the concrete was
obtained by saw-cutting and polishing the specimen. To avoid the destruction (raveling)
of the paste phase during cutting, the mortar specimen was allowed to hydrate until a
sufficient strength was achieved at the time of cutting. The saw-cut surface was polished
after cutting to obtain a flat and clean surface, which provided a clearly defined
composite image (Figure 5.2 (a)). Polishing was performed using equipments composed
of a flat-turntable with fine grits ranging from 6 m to 0.25 m. A general description of
polishing procedures are provided by Allman and Lawrence (1972) and John et al.
(1998). When the polishing process was finished, the polished surface was washed with

106

water and the specimen surface was dried using compressed air. After the specimen
surface was dried, phenolphthalein was applied on the polished surface using a cottontipped stick (Figure 5.2 (b)). The application of phenolphthalein changed the color of the
cement paste phase to pink, which helped to better define the paste phase during the
image analysis (re-coloring) procedure.
One alternative to the application of phenolphthalein is to use white cement with
dark color aggregates, which provide better color-defined phases.

1.27 c m

107

2.54 cm

(a) Cross section of a sample (saw-cut)

(b) Polished surface treated with phenolphthalein


Figure 5.2 Specimen preparation (surface treatments)

108

5.4.2 Scanning
A computer image file was obtained by scanning an actual concrete surface using
a computer scanner (Figure 5.3), which provided a two-dimensional color image scan
with a high resolution (600 dpi is used in this research). Images with high resolution
were needed for the meshing procedure to provide clear boundaries among phases
because meshing is performed based on the size of each pixel.
The resolution of an image also becomes important when the bond phase,
between the paste and aggregate phases, needs to be considered because the bond phase
will have very small thickness (m scale). The mechanical properties of the bond phase
are typically reported to be different (lower strength and stiffness) than those of bulk
cement paste due to increased porosity (Neville, 1996). Scrivener and Gariner (1988)
investigated the increase of porosity at the interfacial zone and found that the porosity at
the interface is three times higher than that of the bulk paste zone (between 0 to 50 m
from the boundary surfaces of aggregates).
The scanning resolution used in this research was 600 dpi (dot(pixel) per inch)
which provides 42 m per pixel. The size (thickness) of the bond phase was determined
by image processing that was based on a pixel unit.

Therefore, a one-pixel thick layer

that has 42 m thickness could be artificially placed around the aggregate phases, which
is compatible with the theoretical bond phase (interfacial zone). The details of ordering
the bond phase in a scanned image are described in the next section. While a higher
resolution would be better suited for the finer refinement at the bond phases, a high-end
computer (faster CPU (higher than Intel P4-3GHz) and higher RAM (random access
memory) capacity of more than 1GB) would be required for the preprocessing phase

109

because a higher resolution image requires more memory. The scanned image file is
saved as a TIFF format file that records the initial color of each pixel. Some file formats
(such as JPEG) compress the size of the original image file, resulting in the loss of the
initial pixels of data.

110

Figure 5.3 Flat-bed color scanner and concrete samples (saw-cut, polished, and stained)

111

5.4.3 Image Analysis (Re-Coloring)


The first step in the image analysis process was to distinguish between the two
phases of the specimen, namely the aggregates and the paste. This was done based on the
color difference between the aggregates and paste in the scanned image.

Natural

aggregates come in several different colors and the paste does not have uniform gray
color which makes it difficult to differentiate between the two phases simply by using
image analysis techniques. As a result, it was decided to stain the paste by applying
phenolphthalein before scanning (this gives paste relatively uniform pink color due to the
high alkalinity of the paste). A two phase image was generated from the scanned image
using the commercial image analysis program, Adobe Photoshop (Figure 5.4). First, the
(pink) cement paste phase was selected and reassigned uniformly as white. The rest of
the image that represents the aggregates was reassigned the color gray. In this research,
the air voids were not considered in the simulation. This implies that the areas of air
voids were assumed to belong to the paste phase.
Using the image analysis program, it was also possible to assign a third phase
(i.e., bond phase) as the paste-aggregate interface, which can have properties that are
different than the bulk paste. By using the expand or contract options in the image
analysis program, a third phase was generated at the perimeter of each aggregate with a
single pixel thickness.
At the end of the re-coloring process, the file format was converted to PPM so
that it could be used for the OOF simulation (PPM2OOF and OOF are compatible with
PPM format images).

1.27 cm

112

2.54 cm

Figure 5.4 Image analysis (re-colorization, part of the scanned image Vagg = 55.7 %)

113

5.4.4 Meshing
Meshing

was

performed

using

the

PPM2OOF

preprocessing program of OOF) (Carter et al., 2000).

code

(complementary

Before meshing, material

properties were ordered for each phase (i.e., the cement paste phase, aggregate phase, and
bond phase) which were previously grouped by color in the re-coloring process. In this
research, it was assumed that all aggregates have the same material properties and the
paste phases also have single values that describe their material properties.
Three node triangular elements were used in this research. The meshing and
refining processes were semi-automatically performed until the meshed boundaries
among the phases become reasonably the same as the original boundary shapes. The
details of the meshing and refining procedures can be found in the PPM2OOF Manual
(Carter et al., 2000). Careful application of the meshing process was performed to
optimize element refinement at the boundaries while avoiding excessive refinement
which can lead to element distortion and thus divergences of numerical calculations.
During the meshing procedures, the total number of meshed elements was maintained at
less than 200,000, which was found to be consistent with the computer (RAM-1.5 GB)
used for this research.
The total number of meshed elements is dependent on the size of an image and
the degree of refinement. Figure 5.5 shows an example of a meshed composite system
that contains about 70,000 elements (the size of the real sample for the portion of this
image was 2.54 cm * 1.27 cm).

1.2 7 cm

114

2.5 4 cm

Figure 5.5 Meshed sample

115

5.5 Finite Element Simulations

5.5.1 General Approach


For finite element simulations, the boundary condition, the type of load, and the
magnitude of the load need to be determined. In this study, two different boundary
conditions were considered (i.e., the free boundary and one directionally restrained
boundary conditions).

The unrestrained specimen corresponded to a free shrinkage

specimen while the restrained specimen considers restraint in a bridge deck or pavement,
and plane stress approximation was used in the simulations. The shrinkage loading
consisted of the application of a uniform volume change in the paste and no volume
change in the aggregates using a temperature substitution analogy as described in Chapter
2 (Moon et al., 2005). The elements representing cement paste were ordered to have a
non-zero thermal expansion coefficient ( = 1.0E-6 /oC in this research) while the
elements representing aggregates were determined as having a zero thermal expansion
coefficient. Therefore, by applying a temperature load (a negative temperature in the
case of the shrinkage of cement paste) onto the entire meshed elements, FE simulations
were performed that simulated concrete shrinkage. For example, a -1 oC of temperature
load will introduce the shrinkage strain of the cement paste phase as a magnitude of -1
when the thermal expansion coefficient of the paste is 1.0E-6 /oC. This approach is
compatible with the concept that the paste phase changes its volume due to autogenous or
drying shrinkage while aggregates do not change their volumes. Here, it should be kept
in mind that the thermal expansion coefficient and the temperatures used in this research

116

do not represent the real thermal expansion coefficient of paste nor the real temperature.
The magnitude of the temperature drop (which is compatible with the magnitude of the
shrinkage strain of cement paste) was successively changed (increased) to observe the
stress development, failure stress (strength), and cracking behaviors of composite
systems.

5.5.2 Failure Criterion


Cracking analysis in finite element simulations can be performed by applying
appropriate failure criterion, and the type of failure criterion can be divided in three
categories: perfectly brittle, quasi-brittle, and perfectly plastic (Figure 5.6). Concrete is
known to behave like a quasi-brittle material while each phase (cement paste and
aggregate) in the concrete behaves more closely to a perfectly brittle material. Therefore,
even though cement paste and aggregates are not perfectly brittle materials, for the
simplicity of the application, a perfectly brittle failure criterion was used for each phase
throughout the course of this research. The perfectly brittle failure criterion was applied
using damisotropic function in the OOF code (Langer et al., 2001). The damisotropic
function is composed of three parts: the basic material properties (elastic modulus,
poissons ratio, and thermal expansion coefficient), the maximum stresses (tensile and
compressive), and the knockdown factor (0~1). The knockdown factor is used to
reduce the elastic modulus of elements when the elements need to behave as cracked ones
in simulations. The knockdown factor is automatically multiplied to the elastic modulus
of an element if the stress in the element exceeds the strength value and the simulation is
restarted at that stage of loading using the new material properties of elements until the

117

simulation satisfies the convergence.

Therefore, the knockdown factor has a value

between 0 to 1, and it needs to be determined considering the failure criterion of a


material. To apply the perfectly brittle failure criterion, the knockdown factor should be
small enough to drop the elastic modulus close to zero but it should be large enough to
avoid the convergence error, which typically happens when the elastic modulus of
elements is too close to zero value, compared to the elastic moduli of other elements.
With the consideration of the range of the material properties (elastic modulus and tensile
strength) of cement paste, the knockdown factor in this research was chosen to be 10-5

Str e s s

Str e s s

Str e s s

through trial and error.

Str a in
(a) Perfectly brittle

Str a in
(b) Quasi-brittle
Figure 5. 6 Failure criteria

Str ain
(c) Perfectly plastic

118

5.5.3 Material Properties


In this study, three different material properties of pastes were chosen and used
for the simulations (Table 5.1).
The typical fine aggregate used in Indiana is likely to be composed of many kinds
of minerals, however, for this research, the average material properties of siliceous quartz
aggregate (Table 5.2) were chosen as siliceous quartz is the representative material for
typical fine aggregate (West, 1995).

Table 5.1 Material properties of cement paste


Paste-1

Paste-2

Paste-3

20

10

0.18

0.18

0.18

Tensile strength (MPa)

2.5

1.25

Compressive strength (MPa)

20

10

Elastic modulus (GPa)


Poissons ratio

Table 5.2 Material properties of aggregate


Fine aggregate
Elastic modulus (GPa)

50

Poissons ratio

0.2

Tensile strength (MPa)

20

Compressive strength (MPa)

200

119

5.5.4 OOF Simulation (Performing Trials)


A single composite image was used for simulations in this research. A mortar
specimen was made using sieved fine aggregates, which were retained between #16 and
#8 sieves (the size distribution of the fine aggregate: 1.19~2.36 mm). The original size of
the mortar specimen was 25.4 cm (10 inch) long with a 2.54 cm (1 inch) wide square
cross-section. The specimen was cured for two days and preprocessing (saw-cutting,
polishing, chemical treatment, scanning, re-coloring, and meshing) was performed
following the procedures described in Section 5.4. The size of the image used for
simulations was 2.54 cm (1 inch) wide and 1.27 cm (0.5 inch) high. The mixture had an
aggregate volume of 55 % (the obtained image was observed to contain 55.7 % of fine
aggregate) (Figure 5.4). The percentage of aggregates in an image was obtained using a
commercial image program (ImagePro PlusTM).
Three different material properties of cement paste were used for the simulations
(Table 5.1). The material properties of fine aggregates are shown in Table 5.2. For these
simulations, it was assumed that cement paste and aggregates are perfectly bonded.
Simulations were performed to observe the influence of autogenous shrinkage on
microcracking and cracking by increasing the magnitude of uniform shrinkage strain of
the paste phase up to the time of failure (in the case of the restrained boundary condition)
and up to -300 (in the case of the free boundary condition).
All simulation data files were saved and the postprocessing tasks were performed,
which are described in the next session.

120

5.6 Post-Processing: Analysis of Results


The post-processing portion of this work consisted of developing methods to
assess the results of the finite element analyses. In addition to using the stresses obtained
directly at the nodes from the finite element analysis, image analysis was performed
using the color contour images obtained from the OOF simulations, where each color
represented a specific level of stress or strain energy density. Numerical analysis using
color contour images can be performed. The benefit of this approach is that the stress or
energy development and distributions in a composite system can be easily quantified.
For the quantification of the specific stress or energy levels in a composite, the area
fraction of each color level was obtained using Image-Pro PlusTM program.
To acquire color scaled images, consistent maximum and minimum limits of scale
needed to be determined.

This research illustrates only the tensile stresses as the

minimum limit of the scale for the color contour of the stress distribution was fixed to
zero. Therefore, compressive stress is represented along with the zero stress elements.
After the determination of the maximum and minimum magnitude of scale, the
type of color scale was chosen. In this research, the Tequila_Sunrise option in OOF
was chosen (Figure 5.7). The Tequila_sunrise scale varies color from red (minimum
magnitude of scale) to yellow (maximum magnitude of scale), changing the RGB (red,
green, blue) values from 255, 0, 0 to 255, 221, 0. The value of red is always 255 and blue
is always zero in the scale. Therefore, each color level can be defined by observing the
intensity change of the green color value.

121

The total number of intervals in the color scale was selected. Theoretically, it is
possible to order the total intervals up to 256, however, only 80 scale intervals were used
in this research.
The color contour images from the simulations were saved as portable pixmap
image format (ppm) by OOF code (OOF saves images in ppm format) and the image files
were converted to bitmap image files (bmp) for post processing. During this process, the
image files were not compressed since this will cause information to be lost. Using
Image-Pro PlusTM, the color contour image was analyzed regarding the area fraction of
each color level, referring to the intensity change of the green color. As a result of image
analysis, Image-Pro PlusTM provided a table of the total number of pixels for each color
level.

Yellow

Figure 5. 7 Color scaled image of stress distribution in a composite


(Paste-1, Horizontally restrained, paste= -140 )

122

The OOF code automatically changed the color of elements to black when the
elements reached to the maximum stress (strength) values (Figure 5. 8). Therefore, from
the image analysis using these images, the area fraction of cracked zones was determined.

Figure 5. 8 Cracked image (paste-1, horizontally-restrained, paste=140 )

5.7 Results and Discussions


In this section, two analytical approaches were discussed. The conventional
approach for the calculation of residual stress development in concrete using equivalent
parameters, which assumes the concrete composite to be a homogeneous effective
medium, and a meso-scale OOF simulation approach, which considers the heterogeneity
of the composite. First, using the conventional approaches, equivalent parameters are
calculated, hence, the equivalent residual stresses are calculated as previously discussed
in Chapter 4. The meso-scale approach will show the results of the OOF simulations,
showing the stress and energy distributions and cracking behaviors in a heterogeneous

123

composite system. Discussions follow regarding the differences between the approaches
and the importance of the meso-scale approach.

5.7.1 Conventional Approach Using Effective Medium Assumption


Chapter 4 showed the conventional approach for the calculation of the residual
stress using equivalent parameters (equivalent shrinkage strain and equivalent elastic
modulus). An actual composite system was shown in Figure 5.4. The volume fraction of
aggregate was determined from Figure 5.4 (55.7 %) and the material properties shown in
Tables 5.1 and 5.2 were used to compute equivalent properties. The equivalent strain
was calculated by Eq. 4.1 and Eq. 4.2, and the equivalent elastic modulus was calculated
by Eq. 4.4. All of these calculated equivalent parameters were compared with the values
directly obtained from the finite element simulations. It was observed that the difference
in the equivalent elastic modulus between the OOF simulation and Hansens model was
less than 5 percent (Figure 5.9). Therefore, it can be said that the conventional equations
for the equivalent strain and equivalent elastic modulus developed by Pickett (1956) and
Hansen (1965) can be used not only for the idealized composite system (hexagonal unit
cell composite), but also for the cases of real composite systems.

124

EquivalentElastic
elasticModulus
modulus
(GPa)
Equivalent
(GPa)

40

30

20

10
OOF Simulation
Hansen's Model
0
0

Paste number (Table 5.1)


Paste number (Table 5.1)

Figure 5.9 Equivalent elastic modulus of a composite


(OOF simulation vs. Hansens model)

Figure 5.10 shows the equivalent stress values for each type of cement paste
(Table 5.1). The equivalent stress was calculated by multiplying the equivalent strain
with the equivalent elastic modulus obtained from Eqs. 4.1, 4.2, and 4.4. As can be seen,
equivalent stress increases when the shrinkage strain of paste and elastic modulus of
paste increases. Here, it can be noticed that the equivalent stress for any type of paste
does not reach the tensile strength of paste (5 MPa) up to - 160 .

125

Equivalent stress (MPa)

2
Paste-1
Paste-2
Paste-3

1.6

1.2

0.8

0.4

0
0

40

80

120

160

Shrinkage strain of paste ()

Figure 5.10 Equivalent stress vs. shrinkage strain of paste

As previously discussed in Chapter 4, the conventional approach cannot account


for the microcracking behavior in a composite system appropriately because the
conventional approach transforms a composite system to an equivalent homogeneous
system, which neglects stress localization. This aspect also becomes important when a
concrete specimen is in the free boundary condition because the conventional approach
assumes that concrete will not experience stress development in a specimen in the free
boundary condition while stresses develop and stress localizations occur in a real
concrete specimen due to the existence of aggregates, which restrain the volume change
of the paste phase due to shrinkage. In the case of the meso-scale approach, the stress

126

development and localization in a composite can be observed, and the strengths of the
different composites can be evaluated from simulations if the material properties of each
phase are known.

5.7.2 OOF Simulation (Results)


Three different material properties were used in the simulations to describe the
properties of the cement paste phase (Table 5.1). OOF simulations were performed by
systematically increasing the shrinkage strain of the paste phase. Two different boundary
conditions (horizontally-restrained and free boundary conditions) were applied. In the
case of the restrained boundary condition, simulations were performed until the reaction
force of the composite at the boundaries decreased to nearly zero. In the case of the
unrestrained boundary condition, simulations were performed with shrinkage in the paste
phase of up to -300 .
Figure 5.11 shows the reaction force that developed at the horizontal boundary of
each composite with different paste properties in the case of the restrained boundary
condition. By knowing the cross-sectional length (height) of the composite, which is
used for the OOF simulation (in this study, 1.27 cm) and assuming that the composite has
a unit thickness (1 m), the average cross-sectional stress was calculated as shown in
Figure 5.11. Here, the average sectional stress along the restrained direction at failure
was determined to be 1.7 MPa, 0.9 MPa, and 0.5 MPa for paste-1, paste-2, and paste-3,
respectively, for the restrained specimens.

As can be seen in Figure 5.11, paste-1

developed the highest reaction force and failed when the shrinkage strain of cement paste

127

reached to -140 . The cases of paste-2 and paste-3 failed at -125 and -120 ,
respectively.
Referring to the material properties of cement paste in Table 5.1, the maximum
tensile strain of paste at failure is -250 (which can be calculated by dividing the tensile
strength by the elastic modulus) while all three composites failed at a much lower
shrinkage strain of the paste, which is due to the existence of stress concentration and
localization in a composite system induced by the heterogeneity. Therefore, it can be
said that the concrete would fail at a lower magnitude of shrinkage strain of the paste
because of the stress concentration and localization while the equivalent (apparent)
shrinkage strain of a composite is smaller than the maximum tensile strain of the paste
phase.
Paste-1 (Elastic)
Paste-2 (Elastic)
Paste-3 (Elastic)
Paste-1 (Cracked)
Paste-2 (Cracked)
Paste-3 (Cracked)

Reaction Force (N)

20000

1.6

15000

1.2

10000

0.8

5000

0.4

0
0

40

80

120

Average Cross Sectional Stress (MPa)

25000

160

Shrinkage Strain of Paste ()

Figure 5.11 Shrinkage strain of paste vs. load (horizontally-restrained boundary


condition)

128

Figures 5.12 and 13 show the cracked images of each case. The results show that
composites demonstrated localized through-cracking in the case of the restrained
boundary condition while distributed cracking occurred in the case of the free boundary
condition. It appears that the change of material properties also changed the paths of the
cracking in the case of the restrained boundary condition (Figure 5.12) while the amount
of microcracks varied with the change of the material properties of the paste in the case
of the free boundary condition (Figure 5.13). The amount of microcracking will be
further discussed later in this chapter.

129

(a) Paste-1 (paste= - 140 )

(b) Paste-2 (paste= - 125 )

(c) Paste-3 (paste= - 120 )

Figure 5.12 Cracks in a composite (one directionally restrained boundary condition)

130

(a) Paste-1 (paste= - 250 )

(b) Paste-2 (paste= - 250 )

(c) Paste-3 (paste= - 250 )

Figure 5.13 Cracks in a composite (free boundary condition)

131

Image analyses were performed using the color contour images of the stress and
energy distributions. To observe the change of the stress and energy developments, two
different simulations were performed: an elastic condition with no cracking and cracking
simulations (Figure 5.14). It was observed in the cracking simulations that the level of
stress successively increased by increasing the paste shrinkage and that the initial cracks
occurred at the boundaries between the paste and the aggregates where the highest
stresses developed due to stress concentration and localization. From Figure 5.14, it can
be seen that the higher stresses are localized at the boundaries between the paste and the
aggregates and these higher stress levels (brighter yellow color) developed in a composite
decrease after cracking.

132

(Pa)
6E6

(a) Elastic condition with no cracking (paste = -140 )

(b) Cracking simulation (paste = -140 )


Figure 5.14 Color contour image of stress distribution in a composite
(One directionally restrained boundary condition)

133

Figure 5.15 shows the cumulative area fraction of the composite that carries a
certain level of residual stresses when the specimen is externally restrained (Figure 5.15
(a)) or unrestrained (Figure 5.15 (b)). For example, in Figure 5.15 (a), the elastic curve
(hollow circles) shows that 40 percent of the composite surface area has a residual stress
level that is higher than 2 MPa stress. By comparing the elastic and cracked curves, it is
apparent that cracking releases stresses in the material.
Cracking occurs where tensile stresses exceed the paste strength. This leads to
through-cracking only after the stress in a small percentage (0.5 %) of the total area
exceeded the strength when the specimen is externally restrained (Figure 5.15 (a)).
However, the specimen without restraint (i.e., free boundary conditions) did not exhibit a
through crack even though the stress in a considerable fraction of the area exceeded the
tensile strength (Figure 5.15 (b)). In this case, dispersed microcracking occurred at areas
with high residual stresses, without significant coalescence of microcracks (i.e., stable
crack growth).
A final observation can be made based on the results presented in Figure 5.15 (a).
In the restrained sample, a through cracking splits the specimen into two parts. Each of
these parts behaves independently and similarly to a specimen with free boundary
condition (compare the stress level between the cracked and the free boundary in Figure
5.15 (a)).

Cumulative area fraction of stress level (%)

134

100

Horizontally Restrained Boundary


Paste-1
Elastic
Cracked
Free boundary

80

paste = -140
60

Average Sectional
Stress (Elastic, 1.7 MPa)
40

Tensile Strength
of paste-1 (5 MPa)
20

0
0

Stress (MPa)

Cumulative area fraction of stress level (%)

(a) Restrained boundary condition (paste = -140 )


100

Free Boundary
Paste-1
80

Elastic
Cracked

paste = -250
60

Tensile Strength
of paste-1 (5MPa)

40

20

0
0

Strain Energy Density (N/m2)


(b) Free boundary condition (paste = -250 )
Figure 5.15 Cumulative area fraction vs. stress

135

Strain energy was used for qualifying and quantifying the cracking behavior of
the cement composite. The amount of strain energy that was stored and released before
and after cracking was detected. Figure 5.16 shows the distribution of the energy density
level in a composite system. As shown in Figure 5.16, the area fractions at lower levels
of strain energy density increased (the area fractions at higher levels of strain energy
density decreased) after cracking. In the case of the restrained boundary condition, the
elastic curve exhibits two peaks, one at higher strain energy density (around 150 N/m2)
which corresponds to higher stress level, and the other at lower energy density (around 25
N/m2) which corresponds to a lower stress level. Cracking causes release of energy in the
high strain energy density (high stress) areas as shown by the elimination of the high
energy peak in Figure 5.16 (a). In the case of free boundary condition, the occurrence of
microcracking did not change the distribution of strain energy density considerably
(Figure 5.16 (b)).

136

25
Paste - 1
Elastic
Cracked

Area Fraction (%)

20

15

10

0
0

100

200

300

400

Strain Energy Density (N/m2)

(a) Restrained boundary condition (paste = -140 )

Area Fraction (%)

16
Paste - 1
Elastic
Cracked

12

0
0

200

400

600

800
2

Strain Energy Density (N/m )

(b) Free boundary condition (paste = -250 )


Figure 5.16 Strain energy density versus area fraction

137

To clearly observe the magnitude of the stored energy for each energy density
level, a normalized stored energy value was used. The normalized stored energy is an
energy value which is calculated by multiplying each energy density level with its area
fraction, which represents the magnitude of the stored energy in a composite for each
energy density level per unit area. Figure 5.17 shows the change of the normalized stored
energy for the energy density level in a composite due to cracking. As can be seen,
portions of the higher strain energy density levels mainly decreased after cracking while
portions of the lower energy density levels increased in the case of the restrained
boundary condition (Figure 5.17 (a)). A similar tendency was also observed in the case
of the free boundary condition (Figure 5.17 (b)). One interesting finding was that the
curve had a bimodal shape. It was found in the simulations that the portion of the first
peak represents the stored energy mainly in the aggregate phases while the portion of the
second peak is the stored energy in the paste phase. Therefore, from Figure 5.17, it can
be seen that the energy mainly decreased in the second peak, which means that the paste
phase released energy due to cracking.
The total area below each curve indicates the normalized total strain energy stored
in a composite. Therefore, it can be expected that the composite after cracking had a
smaller amount of stored energy due to the cracking and the difference of the total stored
energy is the released energy.

138

Normalized Stored Energy (N*m)

Paste-1
Elastic
Cracked

100

200

300

400

Strain Energy Density (N/m2)

(a) Restrained boundary condition (paste = -140 )

Normalized Stored Energy (N*m)

Paste-1
Elastic
Cracked

200

400

600

800

1000

Strain Energy Density (N/m2)

(b) Free boundary condition (paste = -250 )


Figure 5.17 Strain energy density level versus normalized stored energy

139

Figure 5.18 shows the changes of the cumulative normalized stored energy for the
elastic condition and the cracking condition. It can be clearly seen that the energies at
higher levels of a strain energy density were mainly released by cracking. The difference
between the elastic and the cracking conditions represents the released energy by
cracking. When the composite made using paste-1 was restrained it released about 45
percent of its total stored energy at failure (paste = -140 ) (Figure 5.18 (a)). In the case
of the free boundary condition, initial cracking occurred when the shrinkage strain of the
paste was -150 and the microcracks increased until the simulation was stopped at -300
. Figure 5.18 (b), free boundary condition, shows about 33 percent release of the total
stored energy at -250 of paste shrinkage. Therefore, by comparing the total stored
energy in a composite for both simulation cases (elastic and cracking), it is possible to
observe and quantify the cracking related energy changes of the composite systems.

140

Cumulation of Normalized Stored


Energy (N*m)

120

Energy Release

80

40

Paste-1
Elastic
Cracked
0

200

400

600

Strain Energy Density (N/m2)

(a) Restrained boundary condition (paste = -140 )

Cumulation of Normalized Stored


Energy (N*m)

200

Paste-1
Elastic
Cracked

160

Energy Release
120

80

40

400

800

Strain Energy Density (N/m2)

(b) Free boundary condition (paste = -250 )


Figure 5.18 Strain energy density level vs. cumulative values of normalized stored energy

141

Figure 5.19 shows the relationships between the shrinkage strain of the paste and
the cumulative normalized stored energy for a composite with different paste properties.
In the case of the restrained boundary condition, the composite model which was made
using paste-1 (which has a higher elastic modulus and tensile strength) stored more
energy and released more energy at failure due to cracking (Figure 5.19 (a)). It was also
observed that the composite failed at a higher paste shrinkage load when the material
properties of the paste were higher (-140 for paste-1, -125 for paste-2, and -120
for paste-3) where paste-1 had the highest elastic modulus and strength and paste-3 had
the lowest values.
In the case of the free boundary condition, cracks began to occur when the
shrinkage strain of the paste reached -150 for paste-1, -140 for paste-2, and -140
for paste-3. The cumulative normalized stored energy of the unrestrained specimen
increased as microcracking occurred at the external boundaries of the aggregate (until the
shrinkage strain reached -220 for paste-1, -210 for paste-2, and -200 for paste3). As the shrinkage of the paste continued to increase, the microcracks began to
coalesce, resulting in a decrease in the stored strain energy.

Cumulation of Normalized Stored Energy (N*m)

142

120

One Directionally Restrained Boundary


Paste-1 (Elastic)
Paste-2 (Elastic)
Paste-3 (Elastic)
Paste-1 (Cracked)
Paste-2 (Cracked)
Paste-3 (Cracked)

100

80

60

40

20

0
40

60

80

100

120

140

Shrinkage Strain of Paste ()

Cumulation of Normalized Stored Energy (N*m)

(a) Restrained boundary condition

300

Free Boundary
Paste-1 (Elastic)
Paste-2 (Elastic)
Paste-3 (Elastic)
Paste-1 (Cracked)
Paste-2 (Cracked)
Paste-3 (Cracked)

200

100

0
100

150

200

250

300

Shrinkage Strain of Paste ()

(b) Free boundary condition


Figure 5.19 Shrinkage strain of paste vs. cumulation of normalized stored energy

143

Although the analytical approach using the energy of a composite (which is stored
and released) is informative and important for cracking analysis, this approach cannot be
used directly for quantifying the number of microcracks that may have formed. For
example, a composite with paste-3 released lower energy (Figure 5.19 (b)) than the cases
of paste-1 and paste-2 while there was a large number of microcracks in the case of the
free boundary condition (Figure 5.13).
To observe the relationship between the number of microcracks and the energy
approach, the percentage of the energy that was released due to cracking and the area
fraction of cracked zones were plotted (Figure 5.20) in the case of free boundary
condition. As shown in Figure 5.20 (a), lower percentage of stored energy was released
in the case of paste-1 while the magnitude of the stored energy and the released energy is
higher (Figure 5.19 (b)). A similar tendency was observed in the case of the area fraction
of cracked zones as shown in Figure 5.20 (b), where a smaller amount of microcracks
developed in the case of paste-1 at the same magnitude of paste shrinkage. This result
indicates that a composite with a paste phase that has a higher elastic modulus and tensile
strength will have a smaller amount of microcracks and release a lower percentage of its
stored energy, although the magnitude of the stored energy and the released energy is
higher. From this observation, it can be noticed that the percent released energy would
have a relationship with the number of microcracks (area fraction of cracked zones).

144

100

Free Boundary Condition


Paste-1
Paste-2
Paste-3

Released Energy (%)

80

60

40

20

0
100

150

200

250

300

Shrinkage Strain of Paste ()

Area Fraction of Cracked Zones (%)

(a) Percent released energy versus paste strain


Free Boundary Condition
Paste-1
Paste-2
Paste-3

12

0
100

150

200

250

300

Shrinkage Strain of Paste ()

(b) Area fraction of cracked zones versus paste strain


Figure 5.20 Relationships among the percentage of the released energy, area fraction of
cracked zones, and the magnitude of shrinkage strain of paste (free boundary condition)

145

Figure 5.21 shows the relationship between the area fraction of cracks and the
percentage of the released energy. The area fraction of cracks can be seen to have a
single relationship with the percent-released energy regardless of the material properties
of the paste phase in a composite. This result indicates that the degree of microcracking
(percentage of microcracking) can be determined regardless of the change in the paste
properties if the percentage of the released energy and its relationship with the degree of

Area Fraction of Cracked Zones (%)

microcracking are known.

Free Boundary Condition


Paste-1
Paste-2
Paste-3

12

0
0

20

40

60

80

100

Released Energy (%)


Figure 5.21 Relationship between the area fraction of cracked zones and the percentage
of the released energy (free boundary condition)

146

5.8 Summary
This chapter presented the application of a new approach for assessing the
potential of microcracking in a concrete composite undergoing uniform paste shrinkage.
The use of the NIST object-oriented finite element code enabled scanned twodimensional concrete images to be used to develop a finite element mesh to analyze
composite materials on the meso-scale. The simulation results were analyzed using color
contour images of stress and stored energy.
It was shown that the external boundary condition plays a significant role in
governing the cracking behavior of a composite.

Externally-restrained composites

exhibited localized sudden through-cracking while unrestrained composites showed


distributed cracks with stable growth. It was also found that a composite with a stiffer
(higher elastic modulus) cement paste increased the brittleness of the composite, although
it delayed the paste strain that was required for initial cracking.
It was found, in the case of free boundary condition, that there exists a single
relationship between the percentage of released energy and the area fraction of cracked
zones regardless of the variation of the material properties. This result indicates the
possible analytical approach for quantifying the degree of microcracking of a composite
using the parameter of the percent-released energy at cracking.
Further observations are required to observe the influence of the aggregate
volume fraction, the aggregate shape, the aggregate size distribution, and the bond
condition between the cement paste and the aggregates to clearly understand the cracking
behavior of concrete composites.

147

CHAPTER 6: PARAMETRIC STUDY OF SHRINKAGE CRACKING IN A


CONCRETE COMPOSITE

6.1 Introduction
The previous chapter demonstrated the applicability of performing a meso-scale
finite element simulation using the OOF approach developed by NIST since it enables the
use of actual 2-D images to generate the finite element mesh for the composite. It can be
expected that the mixture proportions (namely the volume fraction of the aggregate, the
size distribution of the aggregates, and the shape of the aggregate (shape of aggregate is
not investigated in this study)) will influence the behavior of concrete composites.
This chapter investigates the role of aggregate volume fraction, aggregate size,
and bond between the aggregate and paste using the OOF simulation approach. The
simulation results will be compared to both analytical approaches and experimental
results. A discussion will be provided to describe how the OOF simulation technique
could be used to estimate microcracking and cracking of concrete.

148

6.2 Importance of This Research


No design tool currently exists which clearly considers the influence of material
heterogeneity on the cracking behavior of concrete. Conventional design and analysis
tools assume that the concrete behaves as a homogeneous material considering only the
continuum response of concrete while ignoring microcracking. If the microcracking is
not considered, the long-term durability of the concrete structure may not able to be
accurately described.
The OOF simulation technique was performed by considering the material at the
meso-scale (i.e., mm) to enable the investigation of the microcracking in concrete. In
order to develop a better design tool for evaluating the behavior of concrete, the key
parameters need to be evaluated.

In this research, five different parameters were

considered including: 1) the external boundary condition, 2) the material properties


(elastic modulus, strength), 3) the volume fraction of aggregate, 4) the size distribution of
the aggregate, and 5) the bond condition between the paste and the aggregates. While
previous chapter dealt with the material properties (paste property changes), this chapter
will discuss the other parameters (the volume fraction of the aggregate, the size
distribution of the aggregate, and the bond condition between the paste and the
aggregates) considering two different boundary conditions (one directionally restrained
boundary condition and free boundary condition).

149

6.3 One Directionally Restrained Boundary Condition

6.3.1 Volume fraction of aggregate

6.3.1.1 Theoretical Observation (Conventional Approach)

The material properties of a composite are governed by the properties of the


constituent materials. In the case of concrete (or mortar), tensile strength is mainly
governed by the paste phase, which typically has lower tensile strength than that of the
aggregates. Chapter 4 showed the conventional approach for calculating the equivalent
elastic modulus of a composite using the Hansens model (Eq. 4.4). Figure 6.1 shows the
equivalent elastic modulus (as determined using Hansens model) for a mortar that is
composed of paste -1 and siliceous quartz aggregate as described in Tables 5.1 and 5.2.

Equivalent Elastic Modulus (GPa)

150

50

EAgg. (50 GPa)


40

30

20

Epaste (20 GPa)


10

0
0

20

40

60

80

100

Volume Fraction of Aggregate (%)


Figure 6.1 Equivalent elastic modulus of a composite with the volume fraction
of aggregate

The equivalent shrinkage strain of a concrete composite can be estimated using


Picketts equation (Eq. 4.1 and 4.2). Figure 6.2 shows the equivalent shrinkage strains
that may be experienced in a concrete when the uniform shrinkage strain of paste is -225
. For the calculations, the material properties of the mortar consisted of paste 1 and
siliceous quartz aggregates as described in Tables 5.1 and 5.2.

151

Equivalent Strain ()

250

paste (225 )

200

150

100

50

0
0

20

40

60

80

100

Volume Fraction of Aggregate (%)


Figure 6.2 Equivalent shrinkage strain of a concrete as a function of the volume fraction
of aggregate

Therefore, if the concrete specimen is restrained in one direction, the average


sectional stress (equivalent stress) experienced in the concrete can be calculated by
multiplying the equivalent elastic modulus by the equivalent shrinkage strain as
illustrated in Figure 6.3 (assuming no creep). It can be seen that the equivalent stress that
develops in the concrete (the product of strain and elastic modulus) decreases with
increasing aggregate volume fraction.

152

Equivalent Stress (MPa)

0
0

20

40

60

80

100

Volume Fraction of Aggregate (%)


Figure 6.3 Equivalent stress of a concrete (paste = 225 )

While this information is useful for evaluating the possible stress development in
a concrete specimen, it does not provide information about the cracking potential of
concrete because the strength of concrete also varies as a factor of the volume fraction of
the aggregate.
The tensile strength of a composite can be approximately evaluated using the
series law (Reuss model) as illustrated in Figure 6.4 (a) (Daniel, 1994). If a composite is
uniformly layered and the layers are perfectly bonded, the tensile strength of the
composite can be calculated by Eq. 6.1.
d E paste

f t 'composite = E concrete 1 +
1 paste ult

s E Agg

Eq. 6.1

153

where, Econcrete is the equivalent elastic modulus of concrete (Eq. 4.4) and paste-ult is the
shrinkage strain of the paste phase at the failure stress. The parameters d and s are
indicated in Figure 6.1 (a). If the aggregates are assumed to have a single size (diameter
is d) and are to be placed following the square array rule (Figure 6.4 (b)), Eq. 6.1 can be
modified, replacing d/s term by the volume fraction of aggregate (Eq. 6.2)

4 V Agg
f t 'composite = E concrete 1 +

E paste

1 paste ult
E

Agg

Eq. 6.2

Eq. 6.2 is valid when the volume fraction of aggregate is lower than 78.5 %
because the maximum volume fraction of aggregate in the square array is 78.5 %.

Pas te
Aggr e gate

(a)

(b)

Figure 6.4 Series law for the calculation of the tensile strength of a concrete composite:
(a) layer array (b) square array (particle array)

154

Figure 6.5 shows an example of the variations of the strength and the equivalent
stress of a concrete composite depending on the volume fraction of aggregate. As shown
in Figure 6.5, the strength of a composite decreased when the volume fraction of
aggregate increased. It should be noted, however, that the strength does not decrease as
rapidly as residual stress.

The equivalent stress and strength curves meet at

approximately 9 percent of the volume fraction of aggregate when the paste strain was
about -225 , as shown in Figure 6.5. From this investigation, theoretically, it can be
said that a volume fraction of aggregate at a converging point is a critical volume
fraction of aggregate with which concrete will have a higher cracking potential.

Equivalent Stress (MPa)

Strength (Composite)
4

paste = 225
3

paste = 150
2

paste = 100
1

0
0

20

40

60

80

Volume Fraction of Aggregate (%)


Figure 6.5 Strength and equivalent stress of a composite with the variation of the volume
fraction of aggregate (paste 1 in Table 5.1)

155

Therefore, it can be expected that the cracking potential of a composite will


decrease with a higher or lower volume fraction of aggregate than the critical volume
fraction of aggregate. In a practical point of view, higher volume fraction of aggregate is
recommendable because the cement content for a concrete mixture can de reduced by
increasing aggregate volume (economical).
While Figure 6.5 shows the converging point at about 9 percent, it was also
observed that the volume fraction of aggregate at converging point increases with lower
material properties (especially the elastic modulus) of paste phase. Further investigations
were performed using different material properties of the paste phase (paste 2 and paste
3 in Table 5.1) as shown in Figure 6.6. It can be seen that each composite had a
converging point at different paste shrinkage (-225 for paste-1, -200 for paste-2,
and -180 for paste-3). The composite with paste-1 has the highest paste strain which
corresponds to the highest strength of a composite. This is mainly due to the fact that
paste-1 has the highest paste strength than the others resulting in the increase of the
composite strength. It was previously observed that higher elastic modulus of paste
phase will increase the stress leading to the increase of cracking potential. Therefore, if
the strengths of all paste phases are same, a concrete with higher elastic modulus will
have higher cracking potential (cracking at lower paste shrinkage) while Figure 6.6
shows reverse results due to the change of strength.
While this theoretical approach does not consider the stress concentration and
localization in a composite system, it does illustrate that increasing the volume of
aggregate reduces the potential for cracking.

156

Difference Between Strength


and Stress (MPa)

0.5

Paste - 1 at

0.4

paste = 225

0.3

Paste - 2 at

paste = 201
0.2

Paste - 3 at
0.1

paste = 180

0
0

20

40

60

80

Volume Fraction of Aggregate (%)


Figure 6.6 Cracking potential of composites based on the volume fraction of aggregate
and material properties

157

6.3.1.2 OOF Simulations


Seven different mortar images were prepared and used for this investigation. It
should be noted that two images are artificially created images (Figures 6.7 (c) and (d))
where aggregates have been removed from an actual mortar image using photo editing
tools to vary the volume fraction of aggregate. Each image has a different volume
fraction of aggregate. Figure 6.7 shows the images used for this study. The first three
images were obtained from mortar specimens that contained sieved fine aggregates.
Figure 6.7 (a) is a pre-processed image from a mortar specimen, which consisted of fine
aggregates that were larger than 4.75 mm. Figure 6.7 (b) is a mortar image with 3.35
mm ~ 4 mm fine aggregates, and Figure 6.7 (c) is a mortar image with 1.18 ~2.36 mm
fine aggregates. These three mortar specimens had an aggregate volume of 55% (the
images were observed to contain 61.8, 59.3, and 55.7 % of aggregates, respectively).
Figure 6.7 (d) and (e) are artificial images obtained by deleting aggregate particles in the
image of Figure 6.7 (c). Figures 6.7 (d) and (e) were prepared to observe the cracking
behavior at a low volume fraction of aggregate. Figures 6.7 (f) and (g) are the preprocessed images from mortar specimens with 45% and 15% volume fraction of fine
aggregates (the images were observed to contain 42.6 and 23.4 % of aggregates,
respectively). The original sizes of images are 5 cm wide and 2.5 cm high for Figure 6.7
(a) and (b), and 2.54 cm side and 1.27 cm high for the other images.
Simulations were performed increasing the shrinkage strain of paste under the
horizontally restrained boundary condition. The material properties of paste-1 (Table
5.1) and the fine aggregate (Table 5.2) were used for the paste phase and aggregate phase,
respectively.

2.5 cm

158

5 cm

2.5 cm

(a) Vagg = 61.8 %, original size of image (5 cm x 2.5 cm)

5 cm

(b) Vagg = 59.3 %, original size of image (5 cm x 2.5 cm)


Figure 6.7 Images with different volume fraction of aggregate (continued)

1.27 cm

159

2.54 cm

1.27 cm

(c) Vagg = 55.7 %, original size of image (2.5 cm x 1.27 cm)

2.54 cm

(d) Vagg = 26.6 %, original size of image (2.5 cm x 1.27 cm)


Figure 6.7 Images with different volume fraction of aggregate (continued)

1.27 cm

160

2.54 cm

1.27 cm

(e) Vagg = 13.0 %, original size of image (2.5 cm x 1.27 cm)

2.54 cm

(f) Vagg = 42.6 %, original size of image (2.5 cm x 1.27 cm)


Figure 6.7 Images with different volume fraction of aggregate (continued)

1.27 cm

161

2.54 cm

(g) Vagg = 23.4 %, original size of image (2.5 cm x 1.27 cm)


Figure 6.7 Images with different volume fraction of aggregate

As a first step in performing the OOF simulations, a perfect-bond condition was


assumed to exist between the aggregate and paste phase (with which bond phase has the
same material properties as paste phase) and a brittle failure criterion applied to both the
paste and aggregate.

The details of OOF simulation procedures can be found in

Chapter 5.
Figure 6.8 shows the image of each mortar composite obtained from OOF
simulation with the cracks that would exist in the material at the time of failure. Each
mortar composite failed at a different magnitude of paste shrinkage strain.

162

(a) Vagg = 61.8 % (failed at paste= - 145 )

(b) Vagg = 59.3 % (failed at paste= - 135 )


Figure 6.8 Images of composites with cracks (one directionally restrained
boundary condition continued)

163

(b) Vagg = 55.7 % (failed at paste= - 140 )

(d) Vagg = 26.6 % (failed at paste= - 135 )


Figure 6.8 Images of composites with cracks (one directionally restrained
boundary condition continued)

164

(e) Vagg = 13.0 % (failed at paste= - 125 )

(f) Vagg = 42.6 % (failed at paste= - 125 )


Figure 6.8 Images of composites with cracks (one directionally restrained
boundary condition continued)

165

(g) Vagg = 23.4 % (failed at paste= - 120 )


Figure 6.8 Images of composites with cracks (one directionally restrained
boundary condition)

Results of simulations performed on perfectly restrained specimens are shown in


Figure 6.9. Figure 6.9 (a) shows that the composite with a higher aggregate volume had a
greater increase of the average section stress, and failed at a lower mortar strain (lower
equivalent strain of a composite). While each composite shows the noticeable change of
the mortar shrinkage at failure, the shrinkage strain of the paste at failure did not change
significantly (-120 ~ -145 ) as a function of the volume fraction of aggregate (13.0 ~
61.8 %) (Figure 6.9 (b)).

This result indicates that, if the cracking potential of a

composite is evaluated by the equivalent values (such as equivalent elastic modulus and
strain of a composite), it seems a composite with a higher aggregate volume to have
higher cracking potential.

It is believed, however, that the potential for shrinkage

cracking in actual concrete decreases by increasing the volume fraction of aggregate.


This may be explained by the fact that, in many field applications, the concrete behavior

166

is considered on the homogeneous scale. As such, a concrete with a higher aggregate


volume would exhibit less overall length change of the composite although the shrinkage
of the paste may be the same as shown in Figure 6.9 (b).

167

Average Sectional Stress (MPa)

3
Vagg = 61.8 %
Vagg = 59.3 %
Vagg = 55.7 %
Vagg = 26.6 %
Vagg = 13.0 %
Vagg = 42.6 %
Vagg = 23.4 %

0
0

40

80

120

Shrinkage Strain of Mortar ()

(a)

Average Sectional Stress (MPa)

3
Vagg = 61.8 %
Vagg = 59.3 %
Vagg = 55.7 %
Vagg = 26.6 %
Vagg = 13.0 %
Vagg = 42.6 %
Vagg = 23.4 %

0
0

40

80

120

160

Shrinkage Strain of Paste ()

(b)
Figure 6.9 Average sectional stresses of composites vs. (a) shrinkage strain of mortar and
(b) shrinkage strain of paste (one directionally restrained boundary condition)

168

In addition to observing the strain in the paste at failure, the change of stored
energy in each composite was also recorded. Figure 6.10 (a) and (b) shows the stored
strain energy of the composites as a function of the mortar shrinkage and as a function of
the paste shrinkage. It can be observed that, as the aggregate volume increases, higher
energy is stored in a composite with lower mortar shrinkage (Figure 6.10 (a)). However,
it can be seen that, as the aggregate volume increases, the stored energy is reduced with
the same paste strain (Figure 6.10 (b)). Once the shrinkage of the paste reaches a critical
value, for each mixture, a sudden decrease in stored energy is observed. This sudden
decrease in energy is consistent with the observation of a through cracking. It can be
noted that composites did not release all of their stored energy upon failure (i.e., through
cracking) (Figure 6.10 (a) and (b)). As discussed in Chapter 5, after the specimen fails,
the internal restraint provided by the aggregate still exists while the energy that is stored
due to the external restraint is released by cracking. Figure 6.10 (c) shows that, as a
higher volume fraction of aggregate is used, the amount energy released by the
development of a crack would be reduced.
Figure 6.10 (d) shows the relationship between the percent released energy (i.e.,
the ratio of the energy released on cracking and the energy in the specimen without
permitting cracking) and the area fraction of cracks after failure (i.e., area fraction of
cracked zones). The area fraction of the cracked zone is primarily a measure of the area
of the localized through cracks. Some distributed microcracks may also be included. It
should be noted, however, that the area of microcracks was extremely small for these
simulations. Therefore, we can think that the area fraction of cracked zone is similar to
the width of the localized crack times the length of the localized crack.

A linear

169

relationship exists between the percent energy released and the area of the cracks. This
implies that a brittle specimen (i.e., a specimen which releases the majority of its energy
during the formation of a localized crack) would have a larger crack width. This can be
observed in experimental observations as well (Shah et al., 2004; Kim and Weiss 2003).
From this result, it can be expected that a concrete composite with a higher volume
fraction of aggregate releases less energy when through cracking occurrs. This is similar
to the observation of Chariton et al. (2002) (Figure 6.10 (e)), where the energy released
due to distributed cracking due to moisture loss and cracking around the aggregates
needed to be seperated from the energy consumed during localized cracking.

Stored Strain Energy (N*m)

160
Vagg = 61.8 %
Vagg = 59.3 %
Vagg = 55.7 %
Vagg = 26.6 %
Vagg = 13.0 %
Vagg = 42.6 %
Vagg = 23.4 %

140
120
100
80
60
40
20
0
0

20

40

60

80

100

120

Shrinkage Strain of Mortar ()

Figure 6.10 (a) stored strain energy vs. shrinkage strain of mortar

170

Stored Strain Energy (N*m)

160
Vagg = 61.8 %
Vagg = 59.3 %
Vagg = 55.7 %
Vagg = 26.6 %
Vagg = 13.0 %
Vagg = 42.6 %
Vagg = 23.4 %

140
120
100
80
60
40
20
0
40

60

80

100

120

140

160

Shrinkage Strain of Paste ()

Released Energy Due to Cracking (N*m)

Figure 6.10 (b) stored strain energy vs. shrinkage strain of paste
160
(a) Vagg = 61.8 %
(b) Vagg = 59.3 %
(c) Vagg = 55.7 %
(d) Vagg = 26.6 %
(e) Vagg = 13.0 %
(f ) Vagg = 42.6 %
(g) Vagg = 23.4 %

(e)

120

(d)

(g)

80

(f)
(a)

(c)

40

(b)

0
0

20

40

60

80

Volume Fraction of Aggregate (%)

Figure 6.10 (c) energy release at failure vs. volume fraction of aggregate

171

Area Fraction of Cracked Zone (%)

5
(a) Vagg = 61.8 %
(b) Vagg = 59.3 %
(c) Vagg = 55.7 %
(d) Vagg = 26.6 %
(e) Vagg = 13.0 %
(f ) Vagg = 42.6 %
(g) Vagg = 23.4 %

(g)

(c)

(e)

(d)

(f)
(a)

(b)

0
0

20

40

60

80

100

Released Energy (%)

Figure 6.10 (d) area fraction of cracked zone vs. percent-released energy

Figure 6.10 (e) Acoustic energy development as a function of time (Chariton et al., 2002)

172

6.3.2 Size Distribution of Aggregate


It was observed from the simulations that the specimen (a), (b), and (c) in Figure
6.7 (Vagg = 61.8, 59.3, and 55.7 percent, respectively) showed similar mechanical
behavior (Figures 6.9 and 6.10). It should be remembered that specimen (a) consisted of
aggregates larger than 4.75 mm, specimen (b) consisted of aggregates between 3.35 mm
and 4.00 mm, and specimen (c) consisted of aggregates between 1.18 mm and 2.36 mm.
Therefore, it could be said that the size of aggregate does not appear to have a significant
influence on the cracking behavior of concrete. However, it was observed from the
specimen with a well graded aggregates distribution (i.e., specimen (f) and (g) in Figure
6.7) (Vagg = 42.6 and 23.4 percent) that they showed through-cracking near the larger
aggregates (Figures 6.8 (f) and (g)). This result indicates that there exists an effect of
aggregate distribution on the cracking behavior of a composite. To better understand this
behavior, additional simulations were performed.

These simulations consisted of

aggregates placed in the center of a sample. Five different composites were simulated
and each sample contained a different number of aggregates although the volume fraction
of the aggregate was maintained constant (Figure 6.11). Table 6.1 and Figure 6.11
describes the geometry of the aggregates that were used in each composite. The material
properties used in these simulations are the same as those described in Table 5.1 and 5.2
(paste 1 and the aggregate (siliceous quartz)).

Each composite was horizontally

restrained and the simulations were performed by increasing the shrinkage strain of paste
until the composites failed.

173

Table 6.1 Information on the volume of aggregate in each composite


shown in Figure 6.11
Composite
1EA
2EA
3EA
4EA
5EA

Agg.
Radius (m)
0.075
0.053033
0.043301
0.0375
0.033541

Agg.
Area (m2)
0.017671
0.008836
0.00589
0.004418
0.003534

1EA

2EA

3EA

4EA

Total Agg.
Area (m2)
0.017671
0.017671
0.017671
0.017671
0.017671

5EA

Figure 6.11 Composites with different number of aggregates maintaining the volume
fraction of aggregate

The simulations showed that the size of the aggregate did not have a significant
role on the stresses that were developed or on the cracking potential. The shrinkage
strain of the paste at failure was observed to decrease in a composite with smaller

174

aggregates (Figure 6.12 (a)). This, however, may be due to the influence of the size and
number of the aggregates, where a composite with five small size aggregates has the
largest aggregate chord length along the center cross section (Figure 6.12 (b)).

175

4
1EA
2EA
3EA
4EA
5EA

Stress (MPa)

0
40

80

120

160

200

Shrinkage Strain of Paste ()

(a)

Aggregate Chord Length at the


Critical Section (m)

1.6

1.2

0.8

0.4

0
1EA

2 EA

3EA

4 EA

5EA

Specimen

(b)
Figure 6.12 Stress development in composites: (a) stress development and (b) aggregate
chord length

176

However, these results do not explain why the concrete composites with
distributed aggregates particles would show through crackings around the larger
aggregate (as shown in Figures 6.8 (f) and (g)). To better explain this observation, it is
necessary to understand how the stresses develop in a composite system when it is
restrained from shrinking freely. Figure 6.13 illustrates the sectional stresses that develop
in a composite system when the paste shrinks. It should be noted here that the (average)
sectional stress should be the same for any vertical section that would be taken in the
composite. Due to the existence of aggregates, more stress will develop locally (at the
surface of the aggregate) in the paste phase as discussed in Chapter 4. As such, it can be
imagined that in regions of the specimen, where coarser (larger) aggregate exists, a
greater potential exists for larger stress concentration. As such, sections with a greater
aggregate sizes would have a high potential to become the critical section where failure
would occur.

Figure 6.13 Sectional stress development in a composite under the restrained


boundary condition

177

In an effort to determine the volume fraction of aggregate at each section,


ImagePro PlusTM software was used. Figure 6.14 shows the ratio of the total chord
length of the aggregates (Lagg) to the total length of the specimen height (Length) (i.e.,
sectional ratio) along the entire specimens in the case of mortars with well graded
aggregates (Vagg = 42.6 and 23.4 %). As shown in Figure 6.14, the major of throughcracking occurred near the larger aggregate where the sectional ratio of the aggregate
chord length was high (even though it was not always the highest). It should be noted
that further work may be needed to determine whether this is driven by the aggregate size
or solely the clump of aggregates.

178

1
0.9
0.8
Lagg/Length

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0

0.5

1.5

2.5

Dist ance (cm)

(a) Vagg = 42.6 %


Figure 6.14 The sectional ratio of the total chord length of the aggregates to the height of
the concrete (continued)

179

1
0.9
0.8
Lagg/Length

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0

0.5

1.5

2.5

Distance (cm)

(b) Vagg = 23.4 %


Figure 6.14 The sectional ratio of the total chord length of the aggregates to the height of
the concrete

180

While larger aggregates would seem to increase the cracking potential of a


concrete, the use of larger aggregates are frequently preferred in the field due to the
potential benefit for decreasing the paste content of the concrete. This work does not
contradict that observation, rather it points to the need for uniformity in the aggregate
distribution as clumping of larger aggregates appears to result in stress localization,.

6.3.3 Bond Condition


A cracking analysis was performed assuming that the paste and aggregates were
perfectly bonded. The perfect bond condition consists of a specimen where the material
properties of the bond zone were the same as those of the bulk paste. It is typically
believed, however, that a bond phase exists (interfacial transition zone (ITZ)) which
contains higher porosity, resulting in a lower elastic modulus and lower strength than
those of bulk paste.

Scrivener and Gariner (1988) observed that the mechanical

properties, such as strength and stiffness, decrease about three times at the ITZ, and the
ITZ has a width of about 50 m. As described in Chapter 5, the scanned image had a 600
dpi resolution so one pixel is 42 m wide. Therefore, by ordering an additional one-pixel
phase around the aggregate phases through the pre-processing (Chapter 5), simulations
considering the effect of bond phase can be performed.
To investigate the effect of the bond phase on the cracking behavior, four
different material properties (25%, 50%, 75%, and 100% of the elastic modulus and
tensile strength of bulk paste) were used to simulate the properties of the bond phase.
Poissons ratio was assumed to be constant (0.18) for these simulation.

For this

181

investigation, a single mortar image (Vagg = 55.7 %) was used and the material properties
of the bulk paste and aggregates were not changed (for bulk paste, paste-1 in Table 5.1
and for aggregate, in Table 5.2).
Figure 6.15 shows the images of the cracked composites at failure when the bond
condition was varied in the case of one directionally restrained boundary condition. The
composite with the lower material properties of the bond phase failed at a lower
magnitude of paste shrinkage strain (the 25 % bond case failed at paste = -85 , the 50%
bond case at paste = -100 , the 75% bond case at paste = -120 , and the 100 % bond
case at paste = -140 ).

182

(a) material properties of bond phase (25% of paste phase) (failed at paste = -85 )

(b) material properties of bond phase (50% of paste phase) (failed at paste = -100 )
Figure 6.15 Images of cracked composites at failure (one directionally restrained
boundary condition - continued)

183

(c) material properties of bond phase (75% of paste phase) (failed at paste = -120 )

(d) material properties of bond phase (100% of paste phase) (failed at paste = -140 )
Figure 6.15 Images of cracked composites at failure (one directionally restrained
boundary condition)

Figure 6.16 (a) shows the stored strain energy in a composite with different bond
conditions.

As expected, the case with the bond phase which has 25 % of paste

properties shows the lowest stored energy at the same magnitude of paste shrinkage
strain. The specimen with a weaker bond failed at a much lower paste shrinkage strain

184

than the specimen with stronger bond. It was also observed that more energy was
released from a concrete which has the higher material properties of the bond phase
(Figure 6.16 (b)).

185

Bond Phase
25 %
50 %
75 %
100 %

Stored Strain Energy (N*m)

100

80

60

40

20

0
40

60

80

100

120

140

Shrinkage Strain of Paste ()

Energy Release Due to Cracking (N*m)

(a)
50

40

30

20

10
20

40

60

80

100

Ratio of bond material properties


to paste material properties (%)

(b)
Figure 6.16 Cracking behavior of a mortar composite (Vagg = 55.7 %) with different bond
conditions (one directionally restrained boundary condition): (a) stored strain energy vs.
shrinkage strain of paste and (b) energy release due to cracking vs. bond condition

186

From this investigation, it can be said that the bond condition has an important
role on the cracking potential of a composite in the case of the restrained boundary
condition. As such future research is needed to better quantify the properties of bond that
should be used in the simulations.
All of the simulations were performed assuming that each phase (paste, aggregate,
and bond phases) act as a perfectly brittle material. However, cement paste is a viscoelastic material. It is expected that the visco-elastic behavior at the interface between the
paste and the aggregates will alter the behavior of the composite. To observe how this
cracking behavior changes, a perfectly plastic failure criterion was applied to the bond
phase and additional simulations were performed. Similar to the trials described in
previous session, the mortar composite with a 55.7 % volume fraction of aggregate was
used for the simulations and the material properties of paste-1 and fine aggregate were
used (in Tables 5.1 and 5.2, respectively). The bond phase had the same material
properties of those of the paste phase up to the maximum tensile stress, but the perfectly
plastic condition was applied when the stress exceeded the maximum tensile stress.
Figure 6.17 shows the cracked mortar composite at failure (which failed when
paste = -190 ) in the case of one directionally restrained boundary condition.
Compared to the perfectly brittle case (Figure 6.15 (d)), which failed at -140 of paste
shrinkage strain, failure occurred at a higher paste shrinkage strain and more distributed
cracks (microcracks) were observed. It can be expected that, if lower material properties
are applied to the bond phase with the perfectly plastic failure criterion, there will be
more distributed cracks (microcracks) at the interfaces.

187

In the case of the perfectly plastic failure criterion for the bond phase, distributed
crackings occurred at the phase boundary, and through-cracking followed when the
elements in the bulk paste phase exceeded the maximum tensile stress. Compared to the
perfectly brittle case (Figure 6.15 (d)), the through-cracking zones appear to be wider
(area fraction of cracked zones: 3.37 % (c.f. perfectly brittle: 1.89%)), but it is because of
the perfectly plastic failure condition at the bond phase, which does not release energy at
the cracking points, resulting in a transfer of the excessive stresses to the nearby paste
phases and eventual failure.

Figure 6.17 Image of the cracked mortar composite (Vagg = 55.7 %) at failure
(paste = -190 ) (one directionally restrained boundary condition with perfectly plastic
failure criterion for bond phase)

188

Figure 6.18 (a) shows the stored energy in a composite with different failure
criteria for the bond phase. The specimen with the perfectly plastic failure criterion for
the bond phase failed at a higher shrinkage strain of paste than the perfectly brittle failure
criterion. Figure 6.18 (b) shows the energy release due to cracking. Even though the
magnitude of the energy release for the case of the perfectly plastic failure criterion for
the bond phase is not appreciable, successive increases of energy release were observed
before failure (Figure 6.18 (c)). These release of energy can be attributed to more
cracking at the interface between aggregate and paste.

200
Bond Phase
Perfectly Plastic
Perfectly Brittle

Stored Strain Energy (N*m)

180
160
140
120
100
80
60
40
20
0
40

80

120

160

200

Shrinkage Strain of Paste ()

(a) stored strain energy vs. shrinkage strain of paste


Figure 6.18 Cracking behavior of a mortar composite with different failure criterion on
the bond phase (one directionally restrained boundary condition - contiuned)

Energy release due to cracking (N*m)

189

100

80

Bond Phase
Perfectly Plastic
Perfectly Brittle

60

40

20

0
80

120

160

200

Shrinkage Strain of Paste ()

Energy release due to cracking (N*m)

(b) energy release due to cracking vs. volume fraction of aggregate


3
Bond Phase
Perfectly Plastic
Perfectly Brittle

0
80

120

160

200

Shrinkage Strain of Paste ()

(c) energy release due to cracking vs. volume fraction of aggregate (range up to 5 N.m)
Figure 6.18 Cracking behavior of a mortar composite with different failure criterion on
the bond phase (one directionally restrained boundary condition)

190

6.4 Free Boundary Condition

6.4.1 Volume fraction of aggregate


Figure 6.19 shows the images of each mortar after the application of -250 of
paste shrinkage strain for the free boundary condition simulations (paste 1 and the
aggregates shown in Tables 5.1 and 5.2). Mortar with a lower volume fraction of
aggregate can be seen to exhibit less cracking. In the case of the free boundary condition,
the aggregates provide internal restraint against paste shrinkage.

(a) Vagg = 61.8 % (at paste = -250 )


Figure 6.19 Images of composites with cracks
(free boundary condition, paste = -250 - continued)

191

(b) Vagg = 59.3 % (at paste = -250 )

(c) Vagg = 55.7 % (at paste = -250 )


Figure 6.19 Images of composites with cracks
(free boundary condition, paste = -250 - continued)

192

(d) Vagg = 42.6 % (at paste = -250 )

(e) Vagg = 23.4 % (at paste = -250 )


Figure 6.19 Images of composites with cracks (free boundary condition, paste = -250 )

Figures 6.20 (a) and (b) show the relationship between the area fraction of the
cracked zones and the paste strain. A composite with a higher volume fraction of
aggregate would have more microcracks due to the higher number of restraint points
(aggregates) and the higher bond area. With the increase of paste strain, the microcracks
begin at the bond regions and grow into the bulk paste.

193

Figure 6.21 shows the cracking behavior in concrete under free shrinkage using
the energy approach. Figure 6.21 shows that a composite with higher volume fraction of
aggregate had higher stored energy at the same mortar shrinkage and started
microcracking at lower mortar shrinkage. Figure 6.21 (b) shows the changes of the
stored energy in composites with increasing the shrinkage strain of paste, and Figure 6.21
(c) illustrates that more energy was released with a higher volume fraction of aggregate.

194

Area Fraction of Cracked Zones (%)

8
Vagg = 61.8 %
Vagg = 59.3 %
Vagg = 55.7 %
Vagg = 42.6 %
Vagg = 23.4 %

0
120

160

200

240

280

320

Shrinkage Strain of Paste ()

(a)
Area Fraction of Cracked Zones (%)

5
Vagg = 61.8 %
Vagg = 59.3 %
Vagg = 55.7 %
Vagg = 42.6 %
Vagg = 23.4 %

0
180

200

220

240

Shrinkage Strain of Paste ()

(b)
Figure 6.20 Relationship between the area fraction of cracked zones and the shrinkage
strain of paste (free boundary condition): (a) total range up to 300 and (b) range up
to 250

195

220

Vagg = 61.8 %
Vagg = 59.3 %
Vagg = 55.7 %
Vagg = 42.6 %
Vagg = 23.4 %

Stored Strain Energy (N*m)

200
180
160
140
120
100
80
60
40
20
0
0

40

80

120

160

200

Shrinkage Strain of Mortar ()

(a) stored strain energy vs. shrinkage strain of mortar


220
Vagg = 61.8 %
Vagg = 59.3 %
Vagg = 55.7 %
Vagg = 42.6 %
Vagg = 23.4 %

Stored Strain Energy (N*m)

200
180
160
140
120
100
80
60
40
20
0
50

100

150

200

250

Shrinkage Strain of Paste ()

(b) stored strain energy vs. shrinkage strain of paste


Figure 6.21 Cracking behavior of mortar composites (free boundary condition cont.)

Energy release due to cracking (N*m)

196

160
Vagg = 61.8 %
Vagg = 59.3 %
Vagg = 55.7 %
Vagg = 42.6 %
Vagg = 23.4 %

140
120
100
80
60
40
20
0
160

200

240

280

Shrinkage Strain of Paste ()

(c) energy release due to cracking vs. volume fraction of aggregate


Figure 6.21 Cracking behavior of mortar composites (free boundary condition)

Figure 6.22 (a) shows the relationship between the area fraction of cracked zones
and the energy that is released. Figure 6.22 (b) shows the relationship between the area
fraction of cracked zones and the percent-released energy. As previously observed in
Chapter 5, the relationship between the area fraction of cracked zones and the percentreleased energy does not change when the material properties of the phases were varied.
Therefore, it can be said that the change of the relationship between the area fraction of
cracked zones and the percent-released energy for each composite in Figure 6.22 (b) is
mainly dependent on the volume fraction of aggregate.

197

Area Fraction of Cracked Zones (%)

8
Vagg = 61.8 %
Vagg = 59.3 %
Vagg = 55.7 %
Vagg = 42.6 %
Vagg = 23.4 %

0
0

40

80

120

160

200

Released Energy (N*m)

(a)
Area Fraction of Cracked Zones (%)

8
Vagg = 61.8 %
Vagg = 59.3 %
Vagg = 55.7 %
Vagg = 42.6 %
Vagg = 23.4 %

0
0

20

40

60

80

Released Energy (%)

(b)
Figure 6.22 Relationship between the area fraction of cracked zones and the energy
release (Free boundary condition): (a) released energy and (b) percent-released energy

198

From Figure 6.20, it can be said that a concrete with higher volume fraction of
aggregate will have more microcracks in the free shrinkage condition.

While the

cracking potential or the magnitude of the microcracking of a composite will increase


with a higher volume fraction of aggregate, it should be noticed that the equivalent
shrinkage strain (global strain) of a composite will also vary with the volume fraction of
aggregate. It was observed in 6.3.1.1 that the equivalent strain of a composite decreases
with increasing the volume fraction of aggregate, which is desirable for the stability of a
structure.
Figure 6.19 (d) and (e) clearly showed that microcracks developed primarily
where coarser (larger) aggregates are located, since larger aggregates provide higher
restraint to the paste deformation in the free boundary condition.

199

6.4.2 Bond Condition


Figure 6.23 shows the images of the mortar when a free shrinkage of the paste
(paste = -250 ) is applied with a different bond conditions. A single composite image
(Vagg = 55.7 %) was used and four different material properties were applied to the bond
phase as previously described in Section 6.3.3. The simulations with bond phase which
had lower elastic modulus and strength showed a larger area fraction of cracked zones
(Figure 6.24) because of the lower strength of the bond phase resulting in the increased
cracking at the interfaces between the paste and the aggregates in a composite system.

200

(a) material properties of bond phase (25% of paste phase) (at paste = -250 )

(b) material properties of bond phase (50% of paste phase) (at paste = -250 )
Figure 6.23 Images of cracked composites at paste = -250
(free boundary condition - continued)

201

(c) material properties of bond phase (75% of paste phase) (at paste = -250 )

(d) material properties of bond phase (100% of paste phase) (at paste = -250 )
Figure 6.23 Images of cracked composites at paste = -250 (free boundary condition)

202

Area Fraction of Cracked Zones (%)

12
Bond Phase Properties
25 %
50 %
75 %
100 % ( = paste - 1)

0
50

100

150

200

250

300

Shrinkage Strain of Paste ()

Figure 6.24 Volume fraction of cracked zones depending on the ratio of bond material
properties to paste material properties

Figure 6.25 (a) shows the stored energy in a composite with different bond
conditions. The composites with lower bond properties can be seen to store less energy
as cracked at lower paste shrinkage strain. From Figure 6.25 (b), it can be seen that a
concrete with lower bond properties began microcracking at a lower paste shrinkage.
However, the rate of energy release (slope in Figure 6.25 (b)) was lower for the specimen
with poor bond. Therefore, it can be seen that, at lower values of paste shrinkage, higher
energy was released from a concrete with lower bond material properties but more energy
was released from a concrete with higher bond material properties at larger paste
shrinkage.

203

280

Stored Strain Energy (N*m)

260

Bond Phase
25 %
50 %
75 %
100 %

240
220
200
180
160
140
120
100
80
60
40
20
0
50

100

150

200

250

300

Shrinkage Strain of Paste ()

Energy release due to cracking (N*m)

(a) stored strain energy vs. shrinkage strain of paste


180
Bond phase
25 %
50 %
75 %
100 %

160
140
120
100
80
60
40
20
0
50

100

150

200

250

300

Shrinkage Strain of Paste ()

(b) energy release due to cracking vs. bond condition


Figure 6. 25 Cracking behavior of a mortar composite (Vagg = 55.7 %) with different
bond conditions (free boundary condition)

204

Figure 6.24 illustrates that the relationship between the area fraction of cracked
zones and the percent of the stored energy that was released does not change significantly
as the material properties of the bond phase are varied.

Area Fraction of Cracked Zones (%)

12
Bond Phase
25 %
50 %
75 %
100 %

0
0

20

40

60

80

Released Energy (%)


Figure 6.26 Relationship between the area fraction of cracked zones and the percent
released energy (bond phase, free boundary condition)

Figure 6.27 shows the response of the composite with the two different failure
criteria (i.e., the perfectly brittle failure criterion and perfectly plastic failure criterion) for
the bond phase. A smaller amount of cracking was observed for the material with the
brittle failure criterion (area fraction of cracked zones: perfectly plastic at 2.19 %, and
perfectly brittle at 3.14 % at -250 of paste strain) at lower paste strain, but the

205

difference decreased with higher paste shrinkage (Figure 6. 28). It should be noted that,
while the bond properties significantly influence the restrained system (Figure 6.15, 6.16,
6.17), they are less significant in the free shrinkage specimens.

Figure 6.27 Cracked image of a mortar composite (Vagg = 55.7 %) (paste = -250 ) (free
boundary condition with perfectly plastic failure criterion for bond phase)

206

Area Fraction of Cracked Zones (%)

10
Failure Criterion
(Bond Phase)
Perfectly Plastic
Perfectly Brittle

0
120

160

200

240

280

320

Shrinkage Strain of Paste ()


Figure 6.28 Area fraction of cracked zones vs. shrinkage strain of paste

Figure 6.29 shows the change in the stored energy and the energy release due to
cracking for the perfectly plastic failure criterion and the perfectly brittle failure criterion.
As can be seen, by applying the perfectly plastic failure criterion to the bond phase, a
smaller amount of energy is released because the toughness of the composite increased
and more stresses were transferred through the bond phase noticeably.

207

280
Failure Criterion
(Bond Phase)
Perfectly Plastic
Perfectly Brittle

Stored Strain Energy (N*m)

260
240
220
200
180
160
140
120
100
80
60
40
20
0
50

100

150

200

250

300

Shrinkage Strain of Paste ()

Energy release due to cracking (N*m)

(a) stored strain energy vs. shrinkage strain of paste


180
Failure Criterion
(Bond Phase)
ITZ: Perfectly Plastic
ITZ: Perfectly Brittle

160
140
120
100
80
60
40
20
0
120

160

200

240

280

Shrinkage Strain of Paste ()

(b) energy release due to cracking vs. volume fraction of aggregate


Figure 6.29 Cracking behavior of a mortar composite with different failure criterion on
the bond phase (free boundary condition)

208

Figure 6.30 shows the relationship between the area fraction of cracked zones and
the percent-released energy.

This relationship, as discussed, is independent of the

material properties of the phases in the perfectly brittle failure criterion as illustrated in
Figures 5.26 and 6.21. However, in Figure 6.30, this relationship can be seen to change
depending on the failure criterion. Therefore, it can be concluded that the relationship
between the area fraction of cracked zones and the percent-released energy is dependent
not only on the volume fraction of aggregate but also on the failure criterion of the
phases.

Area Fraction of Cracked Zones (%)

10

Failure Criterion
(Bond Phase)
Perfectly Plastic
Perfectly Brittle

0
0

20

40

60

80

Released Energy (%)

Figure 6.30 Relationship between the area fraction of cracked zones and the percent
released energy

209

From this investigation, it can be concluded that the cracking behavior and
cracking potential of a composite will be highly affected by the condition of the bond
phase. Knowledge about the mechanical properties of the bond phase are still limited
though, so further research is required to apply the proper material properties to the bond
phase to get more realistic results from simulations.

210

6.5 Discussion
By performing the series of OOF simulations, the cracking behavior of concrete
composites was observed for two different boundary conditions.

In the case of a

restrained boundary condition, it was discussed that, theoretically, there exists a critical
volume fraction of aggregate that has the highest cracking potential under the shrinkage
condition. From a practical point of view, a large enough volume fraction of aggregate
was recommended to reduce the potential for cracking. It was found that the role of
aggregate size is less important. However, cracks form at the coarser (larger) aggregates
especially when they clump together as such proper aggregate distribution would be
beneficial to reduce the through-cracking potential.
In the case of the free boundary condition, it was shown from the simulations that
a concrete with a higher volume fraction of aggregate has more distributed cracks
(microcracks).
It was observed that the bond condition between aggregate and paste has an
important role on the cracking potential and the behavior of a composite for the
externally restrained boundary conditions. Further research is expected using proper
material properties for the bond phase.
While this chapter provided a discussion of the cracking behavior of concrete, it
should be noted that all of the simulations performed were based on linear elastic material
behaviors. The material properties of cement paste change with time due to cement
hydration (aging), and the paste experiences creep under loading time. The magnitude of
the shrinkage strain of the paste is a function of time.

Therefore, to improve the

simulation of concrete composites, time should be included. To incorporate creep to the

211

meso-scale OOF simulations, a creep function is needed in which the microcracking and
bond effects are separated as the conventional creep models assume concrete to be a
homogeneous material.

212

CHAPTER 7: SUMMARY

7.1 Introduction
This study was composed of two parts. First, the study focused on developing a
methodology that would enable data from the restrained ring test to be used to quantify
the internal stress development in a concrete specimen when non-uniform shrinkage
occurs due to moisture gradients. Chapter 2 focused on the development of an analytical
procedure that could be used to compute the complex residual stress fields that develop in
the case of non-uniform drying due to time-dependent moisture diffusion. Chapter 3
provided an approach to quantify the degree of restraint that is experienced in restrained
ring specimens as the geometry of the concrete and steel ring are varied. The second part
of this study focused on adapting a finite element simulation technique for the evaluation
of microcracking and cracking behavior that is caused by the heterogeneous nature of
concrete when the paste and aggregate are considered as two separate phases. It was
shown that the cracking behavior was highly dependent on both the external restraint
provided by the boundary condition and the internal restraint that was provided by the
aggregates.

Chapter 4 illustrated that, by assuming that concrete behaves as a

homogeneous material, areas of high residual stress that increase the cracking potential in
concrete may not be considered. A meso-scale finite element simulation technique was
implemented to enable actual images of concrete to be directly used for simulations as

213

described in Chapter 5. Chapter 6 described the influence of many material properties on


cracking from the results of the series of OOF simulations.

7.2 Quantification of the Residual Stress Development in Concrete Due to Non-Uniform


Shrinkage and the Degree of Restraint Exhibited by the Restrained Ring Test When
Different Specimen Geometries are Used
Chapter 2 described the development of an analytical approach for calculating the
residual stresses that develop in a restrained concrete ring under circumferential drying.
Equation 7.1 (or Eq. 2.13) provides the ability to compute the stress that develops as a
function of two components: 1) external restraint due to the steel ring and 2) self-restraint
due to moisture gradients.

( r , ) = , rest . ring + ,diff . shr .


= steel (t ) E S

2
2

ROS
R IS2
1 + ROC
2
2
2 ( ROC
ROS
)
r2

2
2
SH const E con r + R IC

( f ( ROC ) f ( R IC )) + f (r ) f ( R IC ) erfc ( A) r 2
2
2
ROC R IC

Eq. 7.1

where, ROC is the outer radius of the concrete, RIC is the outer radius of the steel (inner
radius of the concrete), Econ is the effective elastic moduli of concrete, ESteel is the elastic
modulus of the steel, Steel is the strain of steel ring (which can be measured from the
attached strain gages), SH-const is the free shrinkage of the concrete, r is the location in the

214

concrete ring where the stresses are being calculated, and erfc is the complementary error
function,
f (r ) =

R
1
2 A
erfc ( A ) A 2 + erfc ( A ) OC A +

+
2
2

4eA

erf ( A )
8

R OC
2e

Eq. 7.2

where A = (R OC r ) / , and = 2 D t where D is the diffusion coefficient of concrete and


t is the time of drying that has been experienced when the calculation is performed.
To utilize the restrained ring test method for the simulation of an actual field
condition, it is essential that an appropriate ring geometry should be determined prior to
performing the experiment (ASTM C1581-04 recommends a single geometry which
typically develops about 70 % of the degree of restraint to the concrete ring). The
geometry of the ring test that should be selected for a particular application should be
designed to satisfy the desired degree of restraint, match the actual drying conditions in
the field, and ensure proper strain measurements to enable accurate stress calculations.
An analytical procedure was presented in Chapter 3 to determine the degree of
restraint in the restrained ring test for concrete under uniform shrinkage (Eq. 3.6 or 7.3).
It was shown that the degree of restraint depends only on the geometry of the rings and
the material properties (namely, the elastic modulus and the Poissons ratio ()) of the
concrete and the steel, regardless of the drying condition.

= 1

E 'C

ES

Eq. 7.3

E 'C

ES

R
1 IS
R OS

R
1 OC
R OS

R
(1 + C ) OC

R OS

+ (1 C )

R
(1 + S ) IS

R OS

+ (1 S )

215

The degree of restraint increases when the thickness and elastic modulus of the
restraining ring (steel) are increased and the degree of restraint decreases with increasing
the thickness of the concrete ring.
In this study, the analytical solutions were compared with the results from finite
element analysis and reasonable agreements were shown. The use of the developed
solutions will enable the restrained ring test to be used for quantifying the potential
development of the residual stresses induced by non-uniform shrinkage (drying) in
concrete structures with the consideration of the degree of restraint.
Further research should be directed at studying the effect of ring height as a ring
with an increased height would induce bending to thin and tall steel rings (resulting in the
non-uniform deformation of concrete along the height direction) for which bending is
expected to be more significant.

216

7.3 Discussion of the Comparison between OOF Simulation Results in a Heterogeneous


Material and Experimental Measurements
To illustrate some of the potentially interesting research findings that can be
obtained from the use of the OOF modeling approach, some of the modeling observations
are presented and compared with experimental behavior of mortar and paste specimens.
For comparison, the OOF simulation data in Section 6.4.1 was used, and these
simulations were performed using the material properties (paste 1 and aggregate in
Tables 5.1 and 5.2).
First, the simulation results were compared with the theoretical solution provided
by Pickett (1956) and Hansen (1965). Figure 7.1 shows the relationship between the
shrinkage strain of the paste and shrinkage strain of mortar obtained from the OOF
simulations (assuming elastic-non-cracking condition) and the analytical equation based
on Picketts model (Eq. 4.1 and 4.2). As can be seen, the results of Picketts model
match well with the OOF simulation results (i.e., the results always within 6 % of the
Picketts model).

217

Shrinkage Strain of Mortar ()

300
Pickett (Theory)
Elastic (Simulation, Vagg = 61.8 %)
Elastic (Simulation, Vagg = 42.6 %)
Elastic (Simulation, Vagg = 23.4 %)
200

100

0
0

100

200

300

Shrinkage Strain of Paste ()

Figure 7.1 A comparison of the shrinkage strain of mortar and the shrinkage strain of
paste as (OOF simulation, elastic)

Figure 7.2 shows the free shrinkage strain of mortars considering both the elastic
response and the response if cracking is permitted.

The mortars begin to develop

microcracking at a free shrinkage strain in the paste of approximately -150 . These


microcracks cause the shrinkage of mortar composites to begin to plateau even though
the shrinkage in the paste continues to increase.

218

Shrinkage Strain of Mortar ()

300
Elastic (Simulation)
Cracked (Simulation, Vagg = 61.8 %)
Cracked (Simulation, Vagg = 42.6 %)
Cracked (Simulation, Vagg = 23.4 %)
200

100

Microcracking
0
0

100

200

300

Shrinkage Strain of Paste ()

Figure 7.2 Relationship between the shrinkage strain of mortar and the shrinkage strain of
paste (OOF simulation, free boundary condition)

Figure 7.3 illustrates how the area of cracking increases (a) and how the energy is
released (b) in the mortar specimens as a function of the shrinkage of the mortar (i.e.,
composite) and paste (i.e., driving force).

It shows that, at the beginning of the

simulation, the shrinkage of mortar increased without cracking. It was observed that
limited cracking began to occur when the paste shrinkage reached -150 (Figure 7.3 (a)
and (c)). However, it should be noted that, after the mortar developed a cracked area of
at least 0.2%, cracking and shrinkage can occur without a substantial increase in the
shrinkage of the composite (Figure 7.3 (b) and (d)). This illustrates that a mortar or
concrete could not continue to shrink or even expand when microcracks develop as

219

shown in Figures 7.2 and 7.3 (mortar with Vagg = 61.8 %). It should be noted, however,

Energy Release Due to Cracking (N*m)

that further work is needed to confirm this with experimental measurements.

160
Vagg = 61.8 %
Vagg = 42.6 %
Vagg = 23.4 %
120

80

40

0
0

40

80

120

160

200

Shrinkage Strain of Mortar ()

(a) energy release vs. shrinkage strain of mortar


Figure 7.3 Energy release and Area fraction of cracked zones versus shrinkage strain of
mortar and paste phase (OOF simulation, free boundary condition - continued)

Energy Release Due to Cracking (N*m)

220

160
Vagg = 61.8 %
Vagg = 42.6 %
Vagg = 23.4 %

140
120
100
80
60
40
20
0
160

200

240

280

Shrinkage Strain of Paste ()

(b) energy release vs. shrinkage strain of paste phase


Area Fraction of Cracked Zones (%)

3
Vagg = 61.8 %
Vagg = 42.6 %
Vagg = 23.4 %
2

0.2 %
0
0

40

80

120

160

200

Shrinkage Strain of Mortar ()

(c) area fraction of cracked zones vs. shrinkage strain of mortar


Figure 7.3 Energy release and Area fraction of cracked zones versus shrinkage strain of
mortar and paste phase (OOF simulation, free boundary condition - continued)

221

Area Fraction of Cracked Zones (%)

5
Vagg = 61.8 %
Vagg = 42.6 %
Vagg = 23.4 %

0
180

200

220

240

260

280

Shrinkage Strain of Paste ()

(d) area fraction of cracked zones vs. shrinkage strain of paste


Figure 7.3 Energy release and Area fraction of cracked zones versus shrinkage strain of
mortar and paste phase (OOF simulation, free boundary condition)

To compare the simulation results with the experimental behavior of mortar and
paste specimens three different water to cement ratio materials were used (0.30, 0.35, and
0.40). The mortar mixtures had 55% fine aggregate (by volume). The specimens were
prepared using the procedure described in (Pease, 2005) with a 2.5 cm wide, 2.5 cm high,
and 30 cm long specimen (Figure 7.4). Shrinkage measurements were performed using a
non-contact laser on paste and mortar specimens from the time of initial set to 24 hours
after placing (Pease et al., 2005).
In addition to measuring the length changes, the internal damage was detected
using acoustic emission.

Puri (2003) showed that the acoustic energy has a linear

relationship with the fracture energy density (i.e., the energy release due to cracking). To

222

detect the internal damage (i.e., microcracking) in mortar specimens caused by


autogenous shrinkage, passive acoustic emission testing was used for both a restrained
shrinkage test and a free shrinkage test (Figure 7.4, Pease et al., 2004). The acoustic
emission sensors were placed on the specimen at the time of initial setting (3 hours for
w/c = 0.30, 4 hours for w/c = 0.35, and 5 hours for w/c = 0.40). Further details about the
passive acoustic emission testing can be found in (Chariton and Weiss, 2002; Kim and
Weiss, 2003; Pease, 2005).

Figure 7.4 Restrained (a) and free (b) shrinkage specimen geometries with locations of
acoustic emission sensors (Pease et al., 2004)

Figure 7.5 (a) shows the autogenous shrinkage strain measured in a paste
specimen with a w/c of 0.30 and the shrinkage strain measured in a mortar specimen after

223

the concrete had hardened (it should be noted that the final set for the mixture with a w/c
of 0.3 was about 5 hours but the data were analyzed beginning at an age of 7 hours). The
shrinkage of the mortar is lower than the shrinkage strain of the paste specimen which is
consistent with Picketts general approach calculated by Eq. 4.1 and 4.2. The evolution
of the dynamic elastic modulus of pastes with 0.3 w/c was measured using ultrasonic
pulse velocity equipment (Figure 7.5 (b)) and 50 GPa was used for aggregate (siliceous
quartz Table 5.2) for the calculation of the Picketts model.

224

w/c = 0.3
Paste
Mortar (Vagg = 55%)
Pickett's Model

Shrinkage Strain ()

600

400

200

0
8

12

16

20

24

Time (hours)

(a)

Elastic Modulus of Paste (GPa)

30

20

10

w/c = 0.3

0
8

12

16

20

24

Time (hours)

(b)
Figure 7.5 Autogenous shrinkage of a paste specimen and a mortar specimen, and the
dynamic elastic modulus of paste (w/c = 0.3): (a) autogenous shrinkage strain of paste
and mortar and (b) dynamic elastic modulus of paste

225

Acoustic energy was monitored on the 0.3 w/c mortar specimen, as well as the 0.35
and 0.4 w/c mortar specimens for unrestrained boundary conditions (Figures 7.6). While
the mortar specimen with 0.3 w/c showed continuous shrinkage, the mortar specimens
with 0.35 and 0.4 w/c experienced little changes showing small amount of expansions
(Figure 7.6 (b)).

Cumulative Acoustic Energy (V*s)

226

1
Free Boundary Condition
0.30 w/c
0.35 w/c
0.40 w/c

0.8

0.6

0.4

0.2

0
8

12

16

20

24

Time (hours)

Cumulative Acoustic Energy (V*s)

(a) cumulative acoustic energy vs. time


1
Free Boundary Condition
0.8

Expand

Shrink

0.6

0.4
0.30 w/c
0.35 w/c
0.40 w/c

0.2

0
-100

100

200

300

Shrinkage of Mortar ()
(b) cumulative acoustic energy vs. shrinkage strain of mortar
Figure 7.6 Cumulative acoustic energies of mortar specimens (free boundary condition)

227

While similar behaviors were observed from both experiments and simulations in
this section, further work is required to better understanding of the microcracking
behavior of mortar and concrete. Additional experiments should be conducted to detect
the length change that occurs as a function of shrinkage in both restrained and
unrestrained specimens. Future researchers should attempt to add time-dependent and
visco-elastic material behavior to the paste.

In addition, to better describe damage

development in the model a correlation should be developed between the cumulative


acoustic energy and the energy that is released in the specimen simulations.

228

7.4 Conclusion
The purpose of this study was to develop solutions for quantifying the cracking
potential and microcracking of concrete due to non-uniform shrinkage in a homogeneous
material and uniform shrinkage in a heterogeneous structure.
Chapter 2 discussed the development of an analytical solution for calculating
residual stresses in the restrained ring test geometry when moisture gradients are present.
The restrained ring test is preferred in many quality control applications because it is
economical and simple to perform. However, insufficient attention has been paid to
problems that can arise interpreting the results from these tests. Specifically, many
people have not considered the influence of moisture gradients and have erroneously used
equations that assume uniform shrinkage occurs throughout the wall of the concrete.
This oversimplification results in the prediction of an incorrect stress field and an
incorrect age of cracking. This work has shown that Equation 2.13 (7.1) can be used to
quantify residual stresses that develop in the concrete ring when drying from the
circumference occurs. The use of Equation 2.13 (7.1) provides an approach for more
accurately computing the failure of the ring test when drying occurs from the outer
circumference.
While the restrained ring test is becoming widely used in industry, there is still
not sufficient information or attention to the degree of restraint that may be experienced
in each specimen as the geometry is varied. This work provides a procedure to match the
stiffness of a structure (i.e., degree of restraint) with the degree of restraint that may be
experienced in the ring test. This would enable the ring test to be used to quantify the
time of cracking and actual residual stress development in a ring specimen that more

229

accurately simulates field performance.

Equation 3.5 (7.2) provides an analytical

solution which can be used to compute the degree of restraint for both uniform and nonuniform shrinkage in the radial direction of the ring.
In addition to the advancements made in interpreting results of the ring test, this
thesis discussed the importance of considering stress localization due to the
heterogeneous nature of concrete to better describe the behavior of concrete composites.
A finite element simulation technique was implemented at the meso-scale to describe the
behavior of concrete using object oriented finite element code and image analysis
software. This approach enabled the use of actual 2-D images of concrete to be easily
meshed and input in a finite element model (Chapter 5). Chapters 5 and 6 discussed the
cracking behavior of concrete from the series of OOF simulations varying material
properties, the volume fraction of aggregate, size of aggregates, bond condition, and
restraint condition. While Chapter 6 began to investigate the role of several parameters
on the performance of concrete, this work has really just begun to lay the foundation for
using this approach. More research is needed to incorporate the visco-elastic behavior
of the paste and the bond phase to be incorporated into the simulation as well as to
extend the simulation to three dimensions.

LIST OF REFERENCES

230

LIST OF REFERENCES

ACI 209R-92, 1997, Prediction of Creep, Shrinkage, and Temperature Effects in


Concrete Structures, Manual of Concrete Practice, American Concrete Institute,
Farmington Hills, MI
ACIs Manual of Concrete Practice; 209, (under revision), Committee report on
Shrinkage and Creep, American Concrete Institute, Farmington Hills, MI
ACIs Manual of Concrete Practice; 207.2R, (1995), Effect of Restraint, Volume
Change, and Reinforcement on Cracking of Mass Concrete, American Concrete
Institute, Farmington Hills, MI
Allman, M and Lawrence, D.F., (1972), Geological laboratory techniques, ARCO
Publications. Co., Inc, New York
Altoubat, S.A., and Lange, D., (2002), Grip-Specimen Interaction in Uniaxial Restrained
Test, ACI SP 206, Concrete: Material Science to Application: A Tribute To
Surendra P. Shah, Detroit, pp. 189-204
Attiogbe, E.K., Weiss, W.J., and See, H.T., (2004), A Look at the Stress Rate Versus
Time of Cracking Relationship Observed in The Restrained Ring Test, The
Advances in Concrete Through Science and Engineering, a Rilem International
Conference, Northwestern University, Evanston, IL, (Electronic proceedings)

231

Attiogbe, E. K., See, H. T. and Miltenberger, M. A., (2003), Cracking Potential of


Concrete Under Restrained Shrinkage, Proceedings, Advances in Cement and
Concrete: Volume Changes, Cracking, and Durability, Engineering Conferences
International, Copper Mountain, Colorado, pp. 191-200.
Bazant, Z. P., ed., (1986), Fourth RILEM International Symposium on Creep and
Shrinkage of Concrete: Mathematical Modeling, Northwestern University,
August 26-29
Bazant, Z. P., and Chern, J. C., (1985), Concrete at Variable Humidity: Constitutive
Law and Mechanisms, Materials and Structures, RILEM, Paris, Vol. 18, pp.
1~20
Bazant, Z. P., and Wittmann, F. H., (1982), Creep and Shrinkage in Concrete
Structures, JOHN WILEY & SONS, New York (c), pp. 163.256
Bazant, Z. P., and Panula, L., (1978-79) Practical Prediction of Time-Dependent
Deformations of Concrete, Materials and Structures, vol. 11 pp. 307-328, 415434, vol. 12, pp. 169-183
Bazant, Z. P. and Najjar, L. J., (1971), Drying of concrete as a nonlinear diffusion
problem, Cement and Concrete Research, Vol. 1, pp.461~473
Berke, N.S., Dallaire, M.C., Hicks, M.C., and Kerkar, A., (1997), New Developments in
Shrinkage-Reducing Admixtures, Superplasticizers and Other Chemical
Admixtures in Concrete, Proceedings Fifth CANMET/ ACI International
Conference, Rome, Italy

232

Bernard, O. and Bruhwiler, E., 2002, Influence of autogenous shrinkage on early age
behavior of structural elements consisting of concretes of different ages,
Materials and Structures, Vol. 35, No. 235, pp. 550-556

Bisschop, J., and van Mier, J.G.M., (2001), Shrinkage Microcracking in Cement-Based
Materials with Low Water-Cement Ratio, RILEM International Conference on
Early Age Cracking in Cementitious Systems, EAC01, ed. K., Kovler, and A.
Bentur
Boley, B. A. and Weiner, J. H., (1988), Theory of Thermal Stress, Dover
PublicationsTimoshenko, S. P. and Goodier, J. N., (1970), Theory of Elasticity,
McGraw Hill College Div
Burrows, R. W., (1998) "The Visible and Invisible Cracking of Concrete", ACI
Monograph No. 11, ACI, Farmington Hills, MI, pp. 78
Carlson, R. W., and Reading, T.J., (1988), Model of Studying Shrinkage Cracking in
Concrete Building Walls, ACI Structural Journal, Vol. 85(4), pp. 395-404
Carlson, R. W., (1937), Drying Shrinkage of Large Concrete Members, Journal of the
American Concrete Institute, pp. 327~336
Carter, W.C., Langer, S.A., and Fuller, E.R., 2000, The OOF Manual: Version 1.0.8.6,
National Institute of Standards and Technology

Chariton, T., and Weiss, W. J., (2002) Using Acoustic Emission to Monitor Damage
Development in Mortars Restrained from Volumetric Changes, Concrete:
Material Science to Application, A Tribute to Surendra P. Shah, eds. P. Balaguru,
A. Namaan, W. Weiss, ACI SP-206, pp. 205-218

233

Chariton, T., Kim, B., Weiss, W.J., (2002), Using Passive Acoustic Energy to Quantify
Cracking in Volumetrically Restrained Cementitious Systems, 15th ASCE EMD
Conference, New York
Dally, J.W. and Riley, W.F., (1991), Experimental Stress Analysis, Third Edition,
McGraw-Hill, Inc.
Daniel, I. M., and Ishai, O., (1994), Engineering Mechanics of Composite Material,
Oxford University Press, Inc.
Darwin, D., Browning, J., and Lindquist, W.D., (2004), Control of Cracking in Bridge
Decks: Observations from the Field, Cement, Concrete, and Aggregates, ASTM,
Vol. 26, No. 2, pp. 148-154
Dela, B.F., and Stang, H., (2000), Two-dimensional analysis of crack formation around
aggregates in high-shrinkage cement paste, Engineering Fracture Mechanics,
Vol. 65, pp.149-164
Dutron, R., Le retrait des ciments et betons, Ann. Trav. Publ. de Belgique, (1934) (Based
on the interpretation of this paper as discussed in reference 3)
Foster, S.W., 2000, HIPERPAV-Guidance to avoid early-age cracking in concrete
pavements, ACI Special Publication, Vol. 192, pp. 1109-1122
Fuller, E. (2003) Personal Communication on the OOF programming efforts
Gardner N.J. and Lockman M.J., (2001), "Compliance, Relaxation and Creep Recovery
of Normal Strength Concrete," Second International Conference on Engineering
Materials, Canadian Society for Civil Engineering and the Japan Society for Civil
Engineers, San Jose

234

Grasley, Z. C. and Lange, D. A., (2004), Thermal Dilation and Internal Relative
Humidity in Hardened Cement Paste, Advances in Concrete Through Science
and Engineering, Proceedings of the International Rilem Symposium
Grasley, Z.C., Lange, D.A., and DAmbrosia, M.D., (2003), Internal Relative Humidity
and Drying Stress Gradients in Concrete, Engineering Conferences International,
Advances in Cement and Concrete IX, Copper Mountain, CO
Granger, L., Torrenti, J.M., and Acker, P., (1997), Thoughts About Drying Shrinkage:
Experimental Results and Quantification of Structural Drying Creep, Materials
and Structures. Vol. 30, pp. 588~598
Grzybowski, M., and Shah S.P., (1990), Shrinkage Cracking of Fiber Reinforced
Concrete, ACI Materials Journal, Vol. 87, No.2, pp. 138-148.
Grzybowski, M., (1989a), Determination of Crack Arresting Properties of Fiber
Reinforced Cementitious Composites, Ph.D. Thesis, Royal Institute of
Technology, Stockholm, Sweden
Grzybowski, M., and Shah, S.P., (1989b), Model to Predict Cracking in Fiber
Reinforced Concrete due to Restrained Shrinkage, Magazine of Concrete
Research, Vol. 41, No. 148, pp. 125~135
Hansen, T.C., (1965), Influence of Aggregate and Voids on Modulus of Elasticity of
Concrete, Cement Mortar, and Cement Paste, ACI Journal, Vol. 62, No. 2, pp.
193.216
He, Z., Zhou, X., Li, Z., (2004), New Experimental Method for Studying Early-Age
Cracking of Cement-Based Materials, ACI Materials Journal, Vol. 101

235

Hossain, A. B., and Weiss, W. J., (In press, 2005), "The Role of Specimen Geometry and
Boundary Conditions on Stress Development and Cracking in the Restrained Ring
Test," Cement and Concrete Research Journal
Hossain, A.B., and Weiss, W.J., (2004), Assessing Residual Stress Development and
Stress Relaxation in Restrained Concrete Ring Specimens, Journal of Cement
and Concrete Composites, Vol. 26, pp. 531-540
Hossain, A B. and Weiss, J, (2003a), Assessing residual stress development and stress
relaxation in restrained concrete ring specimens, Cement and Concrete
Composites, Vol. 25
Hossain, A.B., Pease, B.J., and Weiss, W.J., (2003b), Quantifying Early-Age Stress
Development and Cracking in Low w/c Concrete Using the Restrained Ring Test
with Acoustic Emission, Transportation Research Record, Concrete Materials
and Construction 1834, pp 24-33
Hossain, A. B., Shah, H., and Weiss, J., (2003c), The Restrained Shrinkage Behavior of
Specimens

Containing

Shrinkage

Reducing

Admixture

and

Fiber

Reinforcement as Assessed Using The Ring Test, Symposium on Assessing


Early-Age Cracking of Concrete, Cement, Concrete, and Aggregates Journal
Hossain, A. B., (2003d), Assessing Residual Stress Development and Stress Relaxation
in Restrained Concrete Ring Specimens, PhD Thesis, Purdue University, West
Lafayette
Jensen, O.M, and Hansen, P.F., (2001), Autogenous deformation and RH-Change in
Perspective, Cement and Concrete Research, Res. 31 (12), pp.1859~1865

236

John, D.A.St., Poole, A.W. and Sims, I., (1998), Concrete Petrography, A handbook of
investigative techniques, Arnold, London, UK.
Jones, R. and Kaplan, M. F., (1957), The effects of coarse aggregate on the mode of
failure of concrete in compression and flexure, Magazine of Concrete Research,
Vol. 9, No. 26, pp. 89~94
Kim, B., and Weiss, W. J., (2003), Using Acoustic Emission to Quantify Damage in
Restrained Fiber Reinforced Cement Mortars, Cement and Concrete Research,
Feb., 2003, vol. 33, no. 2, pp. 207-214
Kosmatka, S.H., Kerkhoff, B., and Panarese, W.C., 2002, Design and Control of
Concrete Mixtures, 14th ed., Portland Cement Association, Skokie, Illinois

Kovler, K., (1994), Testing System for Determining the Mechanical Behavior of Early
Age Concrete Under Restrained and Free Uniaxial Shrinkage, Materials and
Structures, RILEM, London, U.K., Vol. 27(170), pp. 324-330.
Kovler, K., Sikuler, J., and Bentur, A., (1993), Restrained Shrinkage Tests of Fiber
Reinforced Concrete Ring Specimens: Effect of Core Thermal Expansion,
Materials and Structures, RILEM, Vol. 26, pp. 231-237.
Krause, P. D., Rogalla, E. A., Sherman, M. R., McDonald, D. B., Osborn, A. E. N., and
Pfeifer, D. W., (1995), Transverse Cracking in Newly Constructed Bridge
Decks, NCHRP 380, Project 12.37
LHermite, R.G., (1960), Volume Changes of Concrete, Fourth International
Symposium on the Chemistry of Cement, Washington D.C., pp. 659-702

237

Langer, S. A., Fuller, E. R., and Carter, W. C., (2001), OOF: An Inage-Based FiniteElement Analysis of Material Microstructures, Computing in Science and
Engineering, pp. 15~23
Lim, Y.M., H.C. Wu and V.C. Li, (1999), "Development of Flexural Composite
Properties and Dry Shrinkage Behavior of High Performance Fiber Reinforced
Cementitious Composites at Early Age," J. of Materials, Vol. 96, No. 1, American
Concrete Institute, pp.20-26
Lura, P., Jensen, O.M., Ye, G., and Tanaka, K., (2005), Micro-crack detection in highperformance cementitious materials, 4th Seminar on Self-Desiccation and Its
Importance in Concrete Technology
Malhotra, V.M., and Zoldners, N.G., (1967), Comparison of Ring-Tensile Strength of
Concrete with Compressive, Flexural, and Splitting Tensile Strengths, Journal of
Materials, pp. 160~199
Mehta, P. K. and Monteiro, P. J.M., (1986), Concrete, Prentice Hall, 2nd edition
Mindess, S., Young, J.F., and Darwin, D., (1996), Concrete, Prentice Hall, 2nd edition
Moon, J.H., Rajabipour, F., Pease, B., and Weiss, J., 2005, Autogenous shrinkage,
residual stress, and cracking in cementitious composites: The influence of internal
and external restraint, Fourth international seminar on self-desiccation and its
importance in concrete technology, Maryland, USA

Moon, J.H., and Weiss, W.J., (2006, to be published) Estimating Residual Stress in the
Restrained Ring Test under Circumferential Drying, Journal of Cement and
Concrete Composites

238

Moon, J.H., Rajabipour, F., and Weiss, W.J., (2004a), Incorporating Moisture Diffusion
In The Analysis Of The Restrained Ring Test, Proceedings of the Fourth
International Conference on Concrete under Severe Conditions (Environment &
Loading), Seoul National University & Korea Concrete Institute, Seoul, Korea ,
pp. 1973.1980
Moon, J.H., Pease, B., Rajabipour, F., Weiss, W.J., (2004b), Residual Stress
Development and Fracture in Portland Cement Composites Insitu Stress and
Passive Asoustic Emission Measurements, Presentation for the 106th American
ceramic Society Meeting, Indianapolis
NCHRP Project 12.37, (1995), Transverse Cracking in Newly Constructed Bridge
Decks, National Cooperative Highway Research Program, Wiss, Janney, Elstner
Associated, Inc.
Neithalath, Narayanan, Moon, J.H., Rajabipour, F., Barde A., Weiss, J., and Attiogbe, E.,
(2005), The Role of Binder Composition in Early-Age Shrinkage Cracking:
Modeling How Moisture Gradients Influence Time Dependent Cracking, NSF
Workshop on High Performance Cementitious Composites, Chennai, India
Neville, A.M., (1996), Properties of Concrete, Longman, 4th edition
Pease, B. J., (2005), The Role of Shrinkage Reducing Admixtures on Shrinkage, Stress
Development, and Cracking, MSCE Thesis, Purdue University, West Lafayette
Pease, B. J., Shah, H.R., and Weiss, W.J., (2005), Shrinkage behavior and residual stress
development in mortar containing shrinkage reducing admixture (SRAs), ACISpecial Publication on Concrete Admixtures, Vol. 227, pp. 285-302

239

Pease, B. J., Hossain, A. B., and Weiss, W. J., (2004), Quantifying Volume Change,
Stress Development, and Cracking Due to Self-Desiccation, ACI SP-220,
Autogenous Deformation of Concrete, pp.23.39
Pease, B. J., Hossain, A. B., and Weiss, W. J., (2003a) Measurements of Early Age
Volume Change, To be Published in a Special Publication of the American
Concrete Institute
Pease, B., Neuwald, A., and Weiss, W. J., (2003b), The Influence of Aggregates on
Early Age Cracking in Cementitious Systems, Celebrating Concrete: Role of
Concrete in Sustainable Development, An International Symposium dedicated to
Professor Surendra Shah, Northwestern University, pp. 329~338
Pickett, G., (1956), Effect of Aggregate on Shrinkage of Concrete and Hypothesis
Concerning Shrinkage, Journal of the American Concrete Institute, Vol. 56, pp.
581-90
Puri, Sunil, (2003), Assessing the development of localized damage in concrete under
compressive loading using acoustic emission, MSCE Thesis, Purdue University,
West Lafayette
Rajabipour, F., Weiss, J., and Abraham, D., (2004), Insitu Electrical Conductivity
Measurements to Assess Moisture and Ionic Transport in Concrete, Advances in
Concrete Through Science and Engineering, Proceedings of the International
Rilem Symposium
RILEM (2002), Early Age Cracking in Cementitious Systems a State of the Art Report of
TC-EAS, ed A. Bentur, 2002

240

Schiel, A., Weiss, W. J., Shane, J. D., Berke, N. S., Mason, T. O., and Shah, S. P.,
(2000), Assessing the moisture profile of drying concrete using impedance
spectroscopy, Concrete Science and Engineering, Vol. 2, pp. 106 -116
Scrivener, K. L., Gariner, E. M, (1988), Microstructural gradients in cement paste
around aggregate particles, Materials Research Symposium Proc., 114, pp. 77-85
See, H. T., Attiogbe, E. K. and Miltenberger, M. A., (2003a), Shrinkage Cracking
Characteristics of Concrete Using Ring Specimens, ACI Materials Journal, V.
100, No. 3, pp. 239-245.
See, H. T., Attiogbe, E. K. and Miltenberger, M. A., (2003b), Potential for Restrained
Shrinkage Cracking of Concrete and Mortar, Proceedings of the ASTM
Symposium on Early-Age Cracking of Concrete
Shah, H., Hossain, A., and Weiss, W.J., (2004), Using the Restrained Ring Test in
Conjunction with Acoustic Emission to Quantify the Role of Steel Fibers in
Shinkage Cracking Mitigation, Conference on Fiber Composites, HighPerformance Concretes, and Smart Materials Organized by International Center
for Fiber Reinforced Concrete (ICFRC), Chennai, India, pp. 99-111
Shah, H.R., Hossain, A.B, Mazzotta, G., and Weiss, W. J., (2003), Time-Dependent
Fracture in Restrained Concrete: The Infulence of Notches and Fibers,
Proceedings of Advances in Cement and Concrete IX: Volume Change, Cracking
and Durability, Copper Mountain, Colorado
Shah, S.P., Karaguler, M.E., and Sarigaphuti, M., (1992), Effects of Shrinkage Reducing
Admixture on Restrained Shrinkage Cracking of Concrete, ACI Materials
Journal, Vol. 89, No. 3, pp. 88~90

241

Shah, S.P. and Jenq, Y. S., (1989)On the Concrete Fracture Testing Method, Fracture
Toughness and Fracture Energy, Editors: Minashi et.al., Balkema Publishers
Springenschmidt, R., Gierlinger, E., and Kernozycki, W., (1985), Thermal Stress in
Mass Concrete: A New testing Method and the Influence of Different Cements,
Proceedings of the 15th International Congress for Large Dams, Lausanne, R4, pp.
57-72
Swamy, R. N., and Stavrides, H., (1979), Influence of Fiber Reinforcement on
Restrained Shrinkage Cracking, ACI Journal, Vol. 76, No. 3, pp.443~460
Swazye, M. A., Discussion of the paper of LHermite, R.G., (1960), Volume Changes of
Concrete, Fourth International Symposium on the Chemistry of Cement,
Washington D.C., pp. 700-702
Timoshenko, S.P. and Goodier, J.N., (1970), Theory of Elasticity, McGraw Hill, 3rd
edition
Toma, G., Pigeon, M., Marchand, J., and Bercelo, L., (1999), Early-Age Autogenous
Restrained Shrinkage: Stress Build Up and Relaxation, Self-Desiccation and Its
Importance in Concrete Technology, eds. Persson, B., and Fagerlund G., pp. 6172
Weiss, W. J., (2002a) Chapter 6.1, Experimental Determination of the Time-Zero,
Early Age Cracking In Cementitious Systems RILEM State of the Art Report
TC-EAS, ed A. Bentur, 2002
Weiss, W. J., Shah, S. P., (2002b), Restrained shrinkage cracking: the role of shrinkage
reducing admixtures and specimen geometry, Materials and Structures, Vol. 35,
pp. 85~91

242

Weiss, W. J., and Shah, S. P., (2002c) Restrained Shrinkage Cracking: The Role of
Shrinkage Reducing Admixtures and Specimen Geometry, Materials and
Structures, VOL. 35 NO 246, pp. 85-91
Weiss, W.J., (2001a), Linking insitu monitoring with damage modeling for life-cycle
performance

simulations

of

the

concrete

infrastructure,

NSF

Career

Development Plan, National Science Foundation


Weiss, W. J. and Ferguson, S., (2001b), Restrained Shrinkage Testing: The Impact of
Specimen Geometry on Quality Control Testing for Material Performance
Assessment, Creep, Shrinkage and Durability Mechanics of Concrete and other
Quasi-Brittle Materials, Proceedings of the Sixth International Conference, F.J.
Ulm, Z.P. Bazant and F.H. Wittmann (eds.), Elsevier Science, pp. 645-650.
Weiss, W.J. and Shah, S.P., (2001), Restrained Shrinkage Cracking: The Role of
Shrinkage Reducing Admixtures and Specimen Geometry, RILEM International
Conference on Early-Age Cracking in Cementitious Systems (EAC01), eds. K.,
Kovler and A. Bentur, Haifa Israel, pp. 145-158
Weiss, W.J., Yang, W., and Shah, S.P., (2000), Influence of Specimen Size and
Geometry on Shrinkage Cracking, Journal of Engineering Mechanics Div.,
ASCE, 126(1), pp. 93.101
Weiss, W. J., (1999), Prediction of Early-Age Shrinkage Cracking in Concrete, Ph.D.
Dissertation, Northwestern University
Weiss, W.J., Yang, W., and Shah, S.P., (1998), Shrinkage Cracking of Restrained
Concrete Slabs, ASCE Journal of Engineering Mechanics, Vol. 124, No. 7, pp.
765-774.

243

Weiss, W. J., (1997), "Shrinkage Cracking in Restrained Concrete Slabs: Test Method,
Material Compositions, Shrinkage Reducing Admixtures, and Theoretical
Modeling", MS Thesis, Northwestern Univ., Evanston IL
West, Terry R. (1995), Geology Applied to Engineering, Prentice Hall, Inc,
Simon/Schuster Company, Englewood Cliffs, New Jersey
Wiegrink, K., Marinkunte, S., and Shah, S. P., (1996), " Shrinkage Cracking of High
Strength Concrete," ACI Materials Journal, Vol. 93, No. 5, pp. 409-415
Yang, W., Weiss, W. J., and Shah, S. P., (2000). "Prediction of Shrinkage Stress and
Displacement Fields in a Concrete Slab Restrained by an Elastic Subgrade"
Journal of Engineering Mechanics Division, ASCE, 126(1), pp. 35-42
Yang, Z., Weiss, W.J., and Olek, J., 2005, Using acoustic emission for the detection of
damage caused by tensile loading and its impact on the freeze-thaw resistance of
concrete, International Conference on Construction Materials: ConMat05,
Vancouver, Canada

APPENDIX

244

6
Paste-1

Stress (MPa)

3
Paste-2
2
Paste-3
1

0
0

50

100

150

200

250

300

Strain ()

Figure A.1 Stress-strain curves for paste-1, paste-2, and paste-3 (Table 5.1)
6
100 %

Stress (MPa)

75 %

3
50 %
2
25 %
1

0
0

50

100

150

200

250

Strain ()

Figure A.2 Stress-strain curves for bond phase

300

VITA

245

VITA

Jae Heum Moon was born on April 25, 1971, in Seoul, Korea. After receiving his
secondary and higher secondary education, he attended Hanyang University in Seoul,
Korea and received his Bachelor of Science in Civil Engineering in 1995. Immediately
after completing his Bachelors, he was employed in a construction company, SamWhan
Corporation, Korea, and had worked for three and half years as a civil engineer at a
subway construction field.

In March, 1998, he joined Hanyang University and he

received his Master of Science in Civil Engineering from Hanyang University in 2000.
He joined Purdue University, West Lafayette, Indiana in August 2002, to pursue his
Ph.D. in the school of Civil Engineering.

Вам также может понравиться