You are on page 1of 224

Characterization and Modeling of Structural and SelfMonitoring Behavior of Fiber Reinforced Polymer

Concrete

A Thesis
Presented to
the Faculty of the Department of Civil and Environmental Engineering
University of Houston

In Partial Fulfillment
of the Requirements for the Degree
Master of Science in Civil Engineering

by
Kallol Sett

May, 2003

Characterization and Modeling of Structural and Self-Monitoring


Behavior of Fiber Reinforced Polymer Concrete

Kallol Sett

Approved:
Chairman of the Committee
C. Vipulanandan, Professor and Chairman,
Civil and Environmental Engineering

Committee Members:
Y. L. Mo, Professor,
Civil and Environmental Engineering

K. W. White, Professor,
Mechanical Engineering

E. Joe Charlson, Associate Dean,


Cullen College of Engineering

C. Vipulanandan, Professor and Chairman,


Civil and Environmental Engineering

To my Parents

ACKNOWLEDGEMENTS

I would like to express my sincere gratitude to my advisor Dr. C. Vipulanandan for


his valuable guidance, encouragement and kindness throughout my research.
I would also like to extend my thanks to Dr. Y. L. Mo and Dr. K. W. White for
serving on my committee and taking time to critique my work.
The help extended by the technicians, Martin Kowis, Joe Thomas and Robert Van
Vickle is also gratefully acknowledged.
I would like to acknowledge the members of the GEM group, my friends in the
department for their help and companionship.
The financial support provided by the Department of Civil and Environmental
Engineering at the University of Houston is also gratefully acknowledged.

Characterization and Modeling of Structural and SelfMonitoring Behavior of Fiber Reinforced Polymer
Concrete

An Abstract
of a
Thesis
Presented to
the Faculty of the Department of Civil and Environmental Engineering
University of Houston

In Partial Fulfillment
of the Requirements for the Degree
Master of Science
in Civil Engineering

by
Kallol Sett

May, 2003

vi

ABSTRACT

Polymer Concrete (PC) is being used because of its rapid setting properties, high
strength to density ratio and durability for civil infrastructure repairs to construction of
sewer pipes and manholes. But there is increasing interest in using self-monitoring
materials for these applications and hence developing smart PC was investigated.
In this study the effect of adding glass and carbon fibers on the mechanical and
electrical properties of polyester polymer concrete was investigated. Strength, stiffness
and stress-strain models were used to predict the behavior of PC with and without fibers.
It had been shown that pulse velocity method predict the static moduli more closely than
the impact resonance method.
The self-monitoring property of fiber reinforced polymer concrete was studied
experimentally under both uniaxial compression and uniaxial tension and characterized in
terms of piezoresistivity coefficients and gage factors. The conductivity increased and the
sensitivity of self-monitoring property decreased with increased fiber content. The
electrical response of the bulk sensor composite was on an average 10 to 100 times
higher than the strain in the material. An electro-mechanical constitutive model was
developed by combining the principle of percolation theory and continuum mechanics to
model the self-monitoring behavior of fiber reinforced polymer concrete. Structural
application of the self-monitoring composite as a bulk sensor in beam under four-point
loading and sandwich pipe configuration under parallel plate loading were tested and
modeled.

vii

TABLE OF CONTENTS
Acknowledgements... v
Abstract....vii
Table of Contents....viii
List of Figures......... xii
List of Tables. .... xx

Chapter 1: INTRODUCTION
1.1 Objective... 2
1.2 Organization.. 3

Chapter 2: LITERATURE REVIEW


2.1 Introduction4
2.2 Polymer Concrete (PC)..... 4
2.3 Fiber Reinforced Polymeric and Cement Composites.. 5
2.4 Self-Monitoring Structural Material. 7
2.5 Electrical Resistivity of Multiphase Conductive-Insulative Solids.. 7
2.5.1 Microstructure.. 8
2.5.2 Percolation Theory.. 9
2.5.3 Effective Medium Theory.. 11
2.6 Piezoresistivity of Multiphase Conductive-Insulative Solids. 11

viii

2.6.1 Microstructure 13
2.6.2 Extended Percolation Theory. 15
2.6.3 Extended Effective Medium Theory.. 18
2.7 Summary..19

Chapter 3: STRUCTURAL BEHAVIOR OF FIBER REINFORCED


POLYMER CONCRETE
3.1 Introduction. 21
3.2 Specimen Preparation. 22
3.3 Testing Program.. 24
3.4 Test Results and Discussions
3.4.1 Destructive Tests
3.4.1.1 Plane Polymer Concrete. 27
3.4.1.2 Polymer Concrete with Glass Fibers.. 33
3.4.1.3 Polymer Concrete with Carbon Fibers... 34
3.4.2 Non-Destructive Tests... 37
3.5 Strength Model.... 51
3.6 Stress-Strain Model..... 52
3.7 Summary..... 55

ix

Chapter 4: SELF-MONITORING BEHAVIOR OF FIBER


REINFORCED POLYMER CONCRETE
4.1 Introduction. 59
4.2 Specimen Design and Instrumentation... 58
4.3 Testing Program.. 60
4.4 Test Results and Discussions.. 65
4.4.1 Compressive Behavior (Uniaxial).. 70
4.4.2 Tensile Behavior (Uniaxial)... 79
4.5 Summary. 81

Chapter 5: MODELING OF SELF-MONITORING BEHAVIOR OF


FIBER REINFORCED POLYMER CONCRETE
5.1 Introduction. 91
5.2 Analytical Formulation... 91
5.3 Resistivity at Unstressed State...102
5.4 Piezoresistivity under Uniaxial Compression....103
5.5 Piezoresistivity under Uniaxial Tension....109
5.6 Parametric Study....114
5.6.1 Compressive Behavior..114
5.6.2 Tensile Behavior...119
5.5 Summary... 123

Chapter 6: APPLICATION OF SELF-MONITORING BEHAVIOR IN


STRUCTURAL ELEMENTS
6.1 Introduction... 128
6.2 CFRPC Beam under Four-Point Bending Test
6.2.1 Specimen Design and Instrumentation 128
6.2.2 Testing Program... 129
6.2.3 Test results and Discussions 129
6.2.4 Modeling...133
6.3 PVC-CFRPC-PVC Sandwich Ring
6.3.1 Specimen Design and Instrumentation.... 149
6.3.2 Testing Program... 150
6.3.3 Test Results and Discussions... 150
6.4 Summary... 156

Chapter 7: CONCLUSIONS AND RECOMMENDATIONS


7.1 Conclusions... 159
7.2 Recommendations. 165

REFERENCES..... 168
APPENDIX.....175

xi

LIST OF FIGURES
Chapter 2
Figure 2.1

Contact Resistance....13

Chapter 3
Figure 3.1

Deformation during (a) shear wave and (b) constrained P-wave.26

Figure 3.2

Compressive stress-strain relationships of polyester PC systems....29

Figure 3.3

Variations of properties with polymer content for polyester PC systems29

Figure 3.4

Behavior models for optimum PC system (a) series (b) series-parallel...31

Figure 3.5 Behavior models for excess PC system (a) series (b) series-parallel31
Figure 3.6 Comparison of predicted and experimental compressive modulus of PC.32
Figure 3.7 Comparison of tensile and compressive behavior of 14% PC system..32
Figure 3.8 Compressive stress-strain relationships of GFRPC systems.33
Figure 3.9 Tensile stress-strain properties of optimum combination of PC, GFRPC and
CFRPC system..35
Figure 3.10 Compressive stress-strain relationships of CFRPC systems35
Figure 3.11 Variations of compressive properties with fiber content for GFRPC and
CFRPC..36
Figure 3.12 Direct tensile stress-strain properties of CFRPC..37
Figure 3.13 Variations of tensile properties with fiber content for CFRPC system38
Figure 3.14 Typical frequency spectrum (X-f) for 6% GFRPC cylinders (a) Longitudinal
mode (b) Transverse mode and (c) Torsional mode.41

xii

Figure 3.15 Typical time decay (X-t) relationships for 6% GFRPC cylinders
(a) Longitudinal mode (b) Transverse mode and (c) Torsional mode......42
Figure 3.16 Damping ratio of 6% GFRPC cylindrical specimen in longitudinal mode..44
Figure 3.17 Comparison of Youngs moduli obtained from pulse velocity and impact
resonance tests for prism specimens.49
Figure 3.18 Comparison of shear moduli obtained from pulse velocity and impact
resonance tests for prism specimens.49
Figure 3.19 Comparison of Youngs moduli obtained from pulse velocity and impact
resonance tests for cylindrical specimens.50
Figure 3.20 Comparison of shear moduli obtained from pulse velocity and impact
resonance tests for cylindrical specimens.50
Figure 3.21 Experimental and predicted tensile strength of CFRPC and GFRPC..51
Figure 3.22 Material constant p for 2.5% CFRPC in tension..54
Figure 3.23 Predicted and measured compressive stress-strain relationships for CFRPC
systems with (a) 4% fibers (b) 6% fibers..56
Figure 3.24 Variations of toughness with material constant p.....57

Chapter 4
Figure 4.1

Instrumentation of CFRPC specimen for compressive electrical property


measurement (a) with brass electrodes (b) with 3-conductor cables61

Figure 4.2

Variations of resistivity of CFRPC system with fiber weigh fraction.62

Figure 4.3

Uniaxial compression test (a) the complete set-up and (b) the specimen....63

Figure 4.4

Uniaxial tension test (a) the complete set-up and (b) the specimen.....64

xiii

Figure 4.5 Change in resistance of 3% CFRPC system under cyclic tensile stress with
all the cycles subjected to the same stress level........66
Figure 4.6

Increase in unstressed resistance of 3% CFRPC system with cycles under


cyclic tensile stress with all the cycles subjected to 25% of failure
strength..66

Figure 4.7 Change in resistance for 3% CFRPC system under cyclic tensile stress with
first cycle subjected to much higher stress than the following cycles..67
Figure 4.8

Typical stress/strain versus change in resistivity relationship of CFRPC in


uniaxial compression....70

Figure 4.9 Compressive stress-strain relationship for 3% CFRPC system in uniaxial


compression (a) overall behavior and (b) initial behavior73
Figure 4.10 Compressive stress-change in resistivity relationship for 3% CFRPC system
in uniaxial compression (a) overall behavior and (b) initial behavior..74
Figure 4.11 Compressive strain-change in resistivity relationship for 3% CFRPC system
in uniaxial compression (a) overall behavior and (b) initial behavior..75
Figure 4.12 Compressive stress-strain relationship for 6% CFRPC system in uniaxial
compression (a) overall behavior and (b) initial behavior76
Figure 4.13 Compressive stress-change in resistivity relationship for 6% CFRPC system
in uniaxial compression (a) overall behavior and (b) initial behavior..77
Figure 4.14 Compressive strain-change in resistivity relationship for 6% CFRPC system
in uniaxial compression (a) overall behavior and (b) initial behavior..78
Figure 4.15 Tensile stress-strain relationship for 3% CFRPC system in uniaxial tension
(a) overall behavior and (b) initial behavior.82

xiv

Figure 4.16 Tensile stress-change in resistivity relationship for 3% CFRPC system in


uniaxial tension (a) overall behavior and (b) initial behavior.83
Figure 4.17 Tensile strain-change in resistivity relationship for 3% CFRPC system in
uniaxial tension (a) overall behavior and (b) initial behavior.84
Figure 4.18 Tensile stress-strain relationship for 6% CFRPC system in uniaxial tension
(a) overall behavior and (b) initial behavior.85
Figure 4.19 Tensile stress-change in resistivity relationship for 6% CFRPC system in
uniaxial tension (a) overall behavior and (b) initial behavior.86
Figure 4.20 Tensile strain-change in resistivity relationship for 6% CFRPC system in
uniaxial tension (a) overall behavior and (b) initial behavior.87
Figure 4.21 Variations of secant piezoresistivity coefficient of different CFRPC systems
with stress (a) overall behavior and (b) initial behavior...88
Figure 4.22 Variations of secant gage factor of different CFRPC systems with stress
(a) overall behavior and (b) initial behavior.....89

Chapter 5
Figure 5.1

Idealized microstructure of carbon fiber reinforced polymer concrete....92

Figure 5.2

Effect of strain on microstructure (a) percolating before straining and


(b) not percolating after straining.....96

Figure 5.3

Prediction of longitudinal resistivity of CFRPC systems with varying fiber


content.....103

Figure 5.4

Model prediction of piezoresistivity under uniaxial compression of 3%


CFRPC system (a) overall behavior and (b) initial behavior......107

xv

Figure 5.5 Model prediction of piezoresistivity under uniaxial compression of 6%


CFRPC system (a) overall behavior and (b) initial behavior.108
Figure 5.6

Model predictions of piezoresistivity under uniaxial tension of 3% CFRPC


system (a) overall behavior and (b) initial behavior...112

Figure 5.7 Model predictions of piezoresistivity under uniaxial tension of 6% CFRPC


system (a) overall behavior and (b) initial behavior...113
Figure 5.8

Predicted effect of Youngs modulus of fiber on compressive piezoresistive


behavior of CFRPC system (a) overall behavior and
(b) initial behavior...115

Figure 5.9

Predicted effect of Poissons ratio of fiber on compressive piezoresistive


behavior of CFRPC system (a) overall behavior and
(b) initial behavior..116

Figure 5.10 Predicted effect of fiber weight fraction on compressive piezoresistive


behavior of CFRPC system (a) overall behavior and
(b) initial behavior...118
Figure 5.11 Variations of mechanical properties of CFRPC with fiber content119
Figure 5.12 Predicted effect of Poissons ratio of composite on compressive
piezoresistive behavior of CFRPC system (a) overall behavior and
(b) initial behavior...120
Figure 5.13 Predicted effect of Youngs modulus of fiber on tensile piezoresistive
behavior of CFRPC system (a) overall behavior and
(b) initial behavior...121

xvi

Figure 5.14 Predicted effect of Poissons ratio of fiber on tensile piezoresistive behavior
of CFRPC system (a) overall behavior and
(b) initial behavior...122
Figure 5.15 Predicted effect of fiber weight fraction on tensile piezoresistive behavior of
CFRPC system (a) overall behavior and
(b) initial behavior...124
Figure 5.16 Predicted effect of Poissons ratio of composite on tensile piezoresistive
behavior of CFRPC system (a) overall behavior and
(b) initial behavior...125

Chapter 6
Figure 6.1

Four-point bending test (a) the complete set-up and (b) the specimen..130

Figure 6.2

Mechanical and electrical response of 3% CFRPC beam under four-point


bending test (a) load versus change in resistance relationship and (b) midspan deflection versus change in resistance relationship....134

Figure 6.3

Variation of piezoresistance coefficient in bending of 3% CFRPC beam


specimen under four-point bending test with load......135

Figure 6.4

Variation of gage factor in bending of 3% CFRPC beam specimen under


four-point bending test with load....135

Figure 6.5

Variation of secant stiffness in bending of 3% CFRPC beam specimen


under four-point bending test with load..136

xvii

Figure 6.6

Mechanical and electrical response of 6% CFRPC beam under four-point


bending test (a) load versus change in resistance relationship and (b) midspan deflection versus change in resistance relationship....137

Figure 6.7

Variation of piezoresistance coefficient in bending of 6% CFRPC beam


specimen under four-point loading with load.138

Figure 6.8

Variation of gage factor in bending of 6% CFRPC beam specimen under


four-point loading with load...138

Figure 6.9

Variation of secant stiffness in bending of 6% CFRPC beam specimen


under four-point bending test with load......139

Figure 6.10 Idealized beam specimen....140


Figure 6.11 Beam under four-point bending test (a) the stresses at different locations (b)
shear force diagram and (c) bending moment diagram..142
Figure 6.12 Predicted variations of change in resistance at idealized locations R1 and R3
with load for 3% and 6% CFRPC beam specimen.146
Figure 6.13 Predicted variations of change in resistance at idealized location R2 with
load for 3% and 6% CFRPC beam specimen.146
Figure 6.14 Predicted variations of change in resistance at idealized locations R4 and R6
with load for 3% and 6% CFRPC beam specimen.147
Figure 6.15 Predicted variations of change in resistance at idealized location R5 with
load for 3% and 6% CFRPC beam specimen.147
Figure 6.16 Experimental and model prediction of piezoresistance of 3% CFRPC system
under four-point bending test (a) overall behavior and (b) initial
behavior...148

xviii

Figure 6.17 Experimental and model prediction of piezoresistance of 6% CFRPC system


under four-point bending test (a) overall behavior and (b) initial
behavior...148
Figure 6.18 PVC-CFRPC (6% Fiber)-PVC sandwich ring under diametrically opposite
compressive loading (a) test set-up for simultaneous mechanical and
electrical measurements and (b) the specimen....151
Figure 6.19 Calibration of strain gage readings with dial gage readings.......152
Figure 6.20 Load versus change in resistance behavior of PVC-CFRPC (6%)-PVC
sandwich ring under parallel plate loading.........154
Figure 6.21 Change in diameter versus change in resistance behavior of PVC-CFRPC
(6%)-PVC sandwich ring under parallel plate loading...154
Figure 6.22 Variation of piezoresistance coefficient with load for PVC-CFRPC (6%)PVC sandwich ring under parallel plate loading....155
Figure 6.23 Variation of gage factor with load for PVC-CFRPC (6%)-PVC sandwich
ring under parallel plate loading.155
Figure 6.20 Variation of ring stiffness with load for PVC-CFRPC (6%)-PVC sandwich
ring under parallel plate loading.156

xix

LIST OF TABLES
Chapter 3
Table 3.1

Compositions of PC, GFRPC and CFRPC..... 22

Table 3.2

Static Test Results of PC, GFRPC and CFRPC..... 39

Table 3.3

Non-Destructive Test Results of Prismatic Specimens.. 43

Table 3.4

Non-Destructive Test Results of Cylindrical Specimens... 45

Table 3.5

Damping Ratios of PC, GFRPC and CFRPC. 47

Chapter 4
Table 4.1

Results of Simultaneous Mechanical and Electrical Measurements of


CFRPC under Uniaxial Compression...72

Table 4.2

Results of Simultaneous Mechanical and Electrical Measurements of


CFRPC under Uniaxial Tension...... 81

xx

CHAPTER 1
INTRODUCTION
1.
Cement based materials are the most used and most researched construction
materials today because these materials are abundant, inexpensive and easy to produce
and handle. But the most important challenge encountered by the research community
investigating cementitious materials is the poor tensile strength and low fatigue toughness
of these materials. These undesirable attributes lead to an easy nucleation and
propagation of cracks restricting the range of their use. This potential drawback of
cementitious composites has leaded the researchers to study other binders as a
replacement of cement.

One popular alternate binder is polymer. Due to its rapid setting, high strength to
density ratio and ability to withstand corrosive environment, polymeric composites are
increasingly being used as an alternate to cementitious composites in many civil
engineering applications ranging from retrofits, rebars underground storage tanks and
repair applications. In order to minimize material cost, it is imperative to use the least
possible amount of polymer in polymeric composites and lead the way for Polymer
Concrete (PC), which is a mixture of aggregate and monomer, which hardens through
polymerization of monomer and Fiber Reinforced Polymer Concrete (FRPC), where
further improvement in mechanical properties can be achieved by using fibers as matrix
reinforcement. Ohama and Nishimura studied the effect of steel fibers and reported
increases in compressive, flexural and impact strength. Glass and carbon fibers are also
becoming popular as matrix reinforcements because of their superior functional

properties like finishability, weatherability and long-term chemical stability in alkaline


and other chemically aggressive environment. Vipulanandan et. al. studied the effect of
glass fibers on the compressive and flexural behavior of PC systems and reported
strength enhancement. But very few studies have been carried out to investigate the effect
of carbon fibers on the PC system.

Also in relation to functional properties, carbon fibers exhibit some exceptional


characteristics compared to other fiber types. One such unique property of carbon fibers
is electrical conductivity, in contrast to glass fibers, which are not conducting. Steel fibers
are conducting, but their typical diameter (> 60 m) is much larger than the diameter of a
typical carbon fiber (10 m). Electrical conductivity of these short fibers, when coupled
with practically insulating polymer matrix exhibit piezoresistivity i.e. the composite
changes its resistivity under applied stress. This piezoresistive property of carbon fiber
reinforced polymer concrete (CFRPC) can be utilized in health monitoring of structures,
which is made with this bulk sensor.

This unique combination of mechanical and functional properties make fiber


reinforced polymer concrete (FRPC) an ideal candidate for next generation construction
material.

1.1 Objective
The main objective of this research is to characterize a high performance selfmonitoring material for civil engineering applications. The specific objectives of this
research are to:
2

1. Investigate the effect of glass and carbon fibers on the compressive and tensile
behavior of PC system and optimize each PC systems.
2. Determine the sensitivity of non-destructive test methods to characterize the
properties of PC with and without fibers.
3. Characterize experimentally the self-monitoring behavior of CFRPC under
uniaxial compression and uniaxial tension.
4. Develop analytical model for predicting the self-monitoring behavior of CFRPC
under uniaxial compression and uniaxial tension.
5. Study the applicability of this CFRPC in sensing load on the structural elements:
CFRPC beam and PVC-CFRPC-PVC sandwich pipe configuration.

1.2 Organization
Chapter 2 summarizes the background and literature review related to fiber
reinforced polymer concrete and piezoresistivity of heterogeneous solids. Structural
behavior of polyester polymer concrete with glass and carbon fibers has been presented
in Chapter 3. Chapter 4 discusses the experimental characterization of self-monitoring
behavior of carbon fiber reinforced polymer concrete under uni-axial compression and
uni-axial tension. Chapter 5 describes the analytical formulation and modeling of change
in resistivity (self-monitoring property) under uni-axial compression and uni-axial
tension. Chapter 6 describes the application of self-monitoring property of CFRPC in
structural element as a bulk sensor to structural damage. Finally, the conclusions and
recommendations of this research have been summarized in Chapter 7.

CHAPTER 2
LITERATURE REVIEW
2.1 Introduction
The main purpose of this chapter is to provide a comprehensive review on the
topics that are closely related to the proposed research. Keeping with the objectives of
characterization of structural and self-monitoring behavior of fiber reinforced polymer
concrete, this chapter summarizes the past and present research on polymer concrete,
followed by current trends on fiber reinforced polymeric and cement composite. Also
self-monitoring behavior of structural material will be reviewed and theoretical work
carried out to-date on predicting the self-monitoring behavior of multiphase conductiveinsulative solids will be summarized in detail for better understanding of the
piezoresistive materials.

2.2 Polymer Concrete (PC)


Polymer Concrete (PC) is a composite material in which aggregates are bound
together in a dense matrix with a polymer binder. The composite does not contain a
hydrated cement paste, although cement can be used as an aggregate or filler (ACI
committee 548). PC was developed in 1950s as a synthetic marble (Fowler, 1989). In the
late 1950s, PC began to be used to make building cladding (Prusinski, 1978). As a
construction material it first achieved widespread recognition in the early 1970s as a
repair material, primarily in pavements and bridges (Fontana and Bartholomew, 1981).
PC has also been used in compressive layer of beam (Lott et. al. 1973), structural and
decorative panels (Prusinski, 1978), linings in carbon-steel pipies for geothermal

applications (Kukacka, 1978), heavy duty industrial floors (Prusinski, 1978), bridge
overlay (Fontana and Bartholomew, 1981), overlays on spillways of hydraulic structures
(Scanlon, 1981), insulator as a substitute to porcelin in electric power industry (Becerra,
1981), machine bases and railroad tie (Fowler, 1990), sewer pipes and manhole
(Bloomfield, 1997).

Depending on the type of formulation, PC composites possess a unique combination


of properties (Fontana and Bartholomew, 1981). These include the following:

1. Rapid curing between temperatures from 180C to +400C (ambient)


2. High tensile, flexural and compressive strengths.
3. Good adhesion to most surfaces.
4. Good long-term durability with respect to cycles of freezing and thawing.
5. Low permeability to water and aggressive solutions and
6. Good chemical resistance.

Polymer concrete is mainly comprised of polymer binder and aggregates. The


polymer binder is obtained by polymerization of a monomer. The most popular
monomers used in polymer concrete are methyl methacrylate (MMA) and polyester.

2.3 Fiber Reinforced Polymeric and Cement Composites


In order to minimize the material cost, it is imperative to use the least amount of
polymer in PC formulations to achieve desired properties depending on its applications.
This has lead to the of fiber addition in the PC system. Steel fibers are one of the most
5

popular fibers for matrix reinforcement. Ohama and Nishimura (1979) studied the effect
of steel fibers in PC systems and reported increases in compressive, flexural and impact
strength. Glass fibers are also very popular because of their close proximity (chemically)
to fine aggregates (sand) used in PC systems. Mebarkia and Vipulanandan (1993) studied
the effect of glass fibers in PC systems and reported compressive strength enhancement.
Carbon fibers are increasingly become popular among the researchers because of their
high tensile strength, high tensile modulus and some unique characteristics like electrical
conductivity and thermal insulation. The application of carbon fiber reinforced composite
is very wide because of its exceptional functional characteristics. Over the last few
decades, carbon fibers have been used in polymeric composites and cement composites to
impart superior functional properties to those composites depending on the applications.
Its applications include strain sensing (Shui and Chung, 1996), monitoring of integrity of
concrete structures (Reza, Baston, Yamamuro and Lee, 2001), thermal insulation (Chung,
2000), radio wave reflection (Chung, 2000), electromagnetic interference (EMI) shielding
(Strumpler and Glatz-Reichenbach, 1999) and electromagnetic discharge (Ueda and
Taya, 1986). In addition to superior mechanical properties, fiber reinforced composites
have other superior functional properties. Cement concrete researchers used this superior
functional property of steel fibers in developing high strength multifunctional cement
concrete. Yehia and Tuan (2002) used the electrical properties of steel fiber reinforced
cement concrete in bridge-deck deicing.

2.4 Self-Monitoring Structural Material


A self-monitoring structural material refers to a structural material that can monitor
itself without the need for embedded, attached and remote sensors (Chung et al. 1995).
6

This is possible because the self-monitoring material is itself a sensor. Compared to


conventional structural materials, which require sensor addition in order to monitor, selfmonitoring structural materials are advantageous in their low cost and large sensing
volume. Self-monitoring of reversible stress or strain has been previously reported in
RuO2-based thick film resistors (Carcia, Suna and Childers, 1983), short graphite fiber
reinforced polymeric composite (Carmona, Canet and Delhaes, 1986), carbon fiber
reinforced cement composite (Chen and Chung, 1993), carbon filament reinforced
polymeric composite (Shui and Chung, 1996) and short carbon fiber reinforced
elastomeric composite (Taya, Kim and ono, 1998). This ability of stress or strain sensing
stems from the reversible increase or decrease in the volume electrical resistivity of these
multiphase conductive-insulative composite materials due to reversible reversible
stressing or straining.

2.5 Electrical Resistivity of Multiphase Conductive-Insulative Solids


The resistivity of multiphase solids, with one phase conductive and others
insulative, depends critically on the volume content of conducting solids. For very low
fiber fraction of conductive solids, the mean distance between conducting solids are large
and the resistance is limited by insulative solids. When a sufficient amount of conductive
solids are loaded, the conducting solids get closer and form linkages, which result in an
initial conducting path through the material. The corresponding conductive solid content
is called percolation threshold (c). In this concentration range, the resistivity can change
drastically by several orders of magnitude for small variations of conductive solid
content. At high loading of conductive solid, the increasing number of conducting paths

forms a three dimensional network. In this range the resistivity is low and less sensitive to
small changes in volume fraction of conductive solids.

2.5.1 Microstructure
At the microstructure level, this phenomenon of the multiphase composite can be
best described by the concept of connectivity. Strumpler and Glatz-Reichenbach (1999)
explained the resistivity of short fiber reinforced polymeric composite with the
connectivity concept. For a short fiber reinforced polymeric composite with very low
content of conducting fibers (when the surrounding polymer matrix insulates each
conducting fiber), the connectivity is said to be 0 (connectivity of conductor)-3
(connectivity of polymer). The first digit signifies that no fiber is connected with the
neighboring fiber in any direction, whereas the second number signifies that the polymer
matrix of a lattice is connected to the polymer matrix of the neighboring lattice in all the
three directions. For dense packing of fibers (when the fibers form a three-dimensional
network) the connectivity is 3-3. That means both fibers and surrounding matrix of a
lattice is connected to the counterpart of neighboring lattices in all the three directions.

When conductivity is 0-3, two different regimes of charge carrier transport are
possible, which are governed by the morphology of the composite material:

(I)

If the mean separation distance between the neighboring fibers is longer than
10 mm, the composite conductivity is the result of transport processes within
the polymer matrix (Strumpler and Glatz-Reichenbach, 1999).

(II)

If the mean separation distance between the two neighboring fibers is below a
certain threshold mean distance of 10 mm, the electrical field assisted
tunneling can occur between neighboring fibers as suggested by Beck.
Alternately, a model introduced by Frenkel suggests an electrical field assisted
hopping mechanism (electron-hole separation) for the charged carrier
transport.

On the other hand, when connectivity is 3-3, the conducting fibers are in close
contact, touching each other in close contact in all the three directions. The conduction of
charged carriers occurs through the continuous structure of the chain of fibers in the
system. The resistivity in this case is mainly determined by the filler material and its
microscopic contacts to adjacent fibers. A transition from insulator to conductor can be
understood as a change in connectivity from 0-3 to x-3 with x>0 and x<3. Change in
connectivity (i.e. change in composite resistivity) can be achieved by increasing the
volume fraction of fiber, by internally induced stress (shrinkage) or by externally applied
stress.

2.5.2 Percolation Theory


At the macrostructure level the resistivity of multiphase composites can be best
described with the help of percolation theory (strumpler and Glatz-Reichenbach, 1999).
The essence of percolation theory is to determine how a given set of lattices, regularly or
randomly positioned in some space, is interconnected. Inherent to the theory is the factor
that at some critical probability, called the percolation threshold, a connected network of
lattices is formed that spans the system, causing the system to percolate.
9

The percolation theory has been used by many researchers (Carmona, Canet and
Delhaes, 1986; Karasek, Meissner, Asai and Sumita, 1995; Strumpler and GlatazReichenbach, 1999; Rejon, Rosas-Zavala, Porcayo-Calderon and Castanao, 2000) to
interpret the behavior of composite conductivity. The sudden transition of composite
material from insulator to conductor is evidence of a percolation threshold. A simple
power law describes the relation between composite conductivity (c) and conductive
filler concentration close to percolation threshold.

c = o (f - crit) t(2.1)
where
c = Composite conductivity (-1cm-1)
o = Conductivity of conductive filler (-1cm-1)
f = Volume fraction of conductive filler (%)
crit = Percolation threshold (%)
t = Critical exponent

The above equation is valid at concentration above the percolation threshold (when f
> crit). The value of t is generally calculated through curve-fit with experimental data.
According to the classical percolation theory, t is close to 2 for three-dimensional
systems. Value of t also depends upon the aspect ratio of the conducting solid particles.
Due to this kind to conductivity behavior, small changes in the volume fractions of
conducting solid, f, close to the threshold where (f - crit) is small, can lead to large

10

change in the composite conductivity. But small changes in the conductive solid
concentration far beyond the threshold value have little effect on the composite
resistivity. This is in line with the micro-structural arrangement and connectivity concept
of composite material.

2.5.3 Effective Medium Theory


Effective Medium Theory (McLachlan and Blaszkiewicz, 1990) is another theory
popular among researchers to predict resistivity of a multi phase composite. The effective
medium theory replaces the inhomogeneous medium found in an actual composite with a
homogeneous effective medium. For a conductive composite, a lattice of similar
symmetry replaces the original lattice consisting of randomly distributed conductances,
so that the electrical properties are on average identical for each situation. Two cases of
effective medium theory exist, called the symmetric and the asymmetric cases. When
one component is perfect insulator, the symmetric media theory contains a conductorinsulator transition. In the asymmetric case, the surface of the particle of one component
is always completely covered by the other component. This means either 0-3 or 3-0
connectivity. This asymmetric case does not contain a transition.

2.6 Piezoresistivity of Multiphase Conductive-Insulative Solids


If volume resistivity of a material changes under applied stress, then the material is
said to be piezoresistive. While the piezoresistive behavior of multiphase conductiveinsulative solids have been mainly studied experimentally (Carcia et al. 1983; Chen and
Chung, 1993; Fu and Chung, 1995; Shui and Chung, 1996; Chung, 2001), there exists

11

very few pullished theoretical studies on piezoresistive behavior of multiphase solids


(Carmona et al. 1986; McLachlan et al. 1990; Taya et al. 1997). Strumpler et al. (1999)
gave an excellent review on the microstructural approach of explaining piezoresistive
behavior of short fiber reinforced polymeric composite. Carmona et al. (1986) proposed
an extended percolation theory involving an appropriate function, which describes the
changes in conductive solid volume fraction due to external loading to predict
piezoresistive behavior of short fiber reinforced polymeric concrete. They showed that
the sensitivity of piezoresistive effect of the composite depends on the elastic properties
of the matrix material and on whether the applied loading is hydrostatic or uniaxial.
McLachlan et al. (1990) pointed out that the approach by Carmona et al. (1986) may not
work because the expansion or movement of one phase with respect to another is not
equivalent to changing the volume fraction by actually varying the volume ratio of the
two phases and proposed extended effective medium theory to predict piezoresistive
behavior of Fe3O4 in flexible epoxy. Taya et al. (1997) predicted piezoresistivity of a
conductive short fiber/elastomer composite by coupling two models: finite deformation
and percolation model. The former model was used to obtain the relation between the
applied finite strain and the resulting change in microstructure of short fiber network,
while the composite resistivity is computed for a given microstructure by using latter
model.

2.6.1 Microstructure
The resistance of the contact spots is the leading contributor to the overall
composite resistivity. Two effects contribute to the contact resistance, the constriction
resistance of a contact spot and the tunneling resistance between the conductive solids.
12

The contact spots are the region of largest resistance and thus of highest electrical losses
and may be the source of large temperature development during current flow through the
composite. Strumpler et al (1999) gave an excellent review on the microstructural
approach of predicting piezoresistivity of short fiber reinforced polymeric composite. If
the fiber-fiber contact is assumed to be hard spheres with Youngs modulus tends to
infinity, the spheres show contact to each other at only one point. This results in
singularities of the contact resistance and the current density, without any dependence on
an applied external stress. In order to avoid these singularities and in agreement to
macroscopic contact spots, Holm (1967) assumed a contact spot with finite radius.

dc

Figure 2.1 - Contact Resistance

The contact spot resistance (Rc) is given by (Holm, 1967)

Rc = f/dc..(2.2)

where f is the bulk resistivity of the fibers. The constriction resistance at the contact
spots dominates over the bulk resistance of the fibers as long as the contact spot diameter
(dc) is much smaller than the fiber diameter. For a more realistic description of the

13

contact resistance the mechanical properties of the fibers have to be taken into account.
The contact spot diameter (dc) can be calculated from the elastic properties of the fibers
with known value of applied force and fiber diameter. If we assume contact arises
between a protrusion of one fiber and a flat surface of neighboring fiber and if the
protrusion diameter at contact is d, the elastic force between fibers at contact is for small
deformation given by (Bush, 1982)

F = [E dc3/ {12(1-2)d}].(2.3)

Where E is the Youngs modulus and is the Poissons ratio. Coupling Eq. (1) and Eq.
(2) we have the contact spot resistance (Rc) as (Strumpler et al. 1999)

Rc = 0.44 f (E/(1-2)1/3 F-1/3 d-1/3(2.4)

Hence the contact resistance as well as composite resistance of the material as a


whole decreased with increasing force (F). Hence internal stresses in the polymer matrix
caused by shrinkage or externally applied stress play an important role in the conductivity
of composite. Shrinkage of carbon fiber reinforced polymer concrete during curing
induces high internal stress. From Eq. (2.4) decrease in resistivity can be expected during
curing process. Similarly, external stress can also change the resistivity of the composite
material by modifying the number and efficiency of the fiber contacts. If any material
changes its resitivity under externally applied stress, the material is called piezoresistive
material.

14

2.6.2 Extended Percolation Theory


Carmona et al. (1987) extended the percolation theory to predict the change in
composite resistivity with hydrostatic stress, considering volume-particle concentration
change under external pressure. They also predicted the piezoresistivity of multiphase
composite under uni-axial stress semi-empirically. When current flows along length L
of a material with resistivity and having cross-sectional area A, we have from Ohms
law

R = G..(2.5)

where,
R = Resistance
G = L/A, an dimensional parameter
If a stress P is applied to the sample at constant temperature (isothermal), then

dR 1 1 G
= +
dP ..(2.6)
R P G P

Hence change in resistance with stress is contributed by two factors (a)


piezoresistivity and (b) elasticity.

Let us now concentrate on piezoresistivity of a heterogeneous mixture of conducting


fibers in an insulating matrix. If is the fiber volume fraction then

15

Vc
.(2.7)
Vc + Vm

where VC = Volume occupied by conducting fibers; Vm = Volume occupied by matrix.


If a hydrostatic pressure is applied to the material, the volume concentration will vary as

1
1 Vc
1
(Vc + Vm)
=

(2.8)
P Vc P Vc + Vm
P

= (1-) (Xm - XC)..(2.9)

where Xm =

Xc =

1 Vm
= Volume compressibility of matrix.
Vm P

1 Vc
= Volume compressibility of conducting fibers.
Vc P

Hence the concentration relative variation under hydrostatic stress is directly related to
the difference in volume compressibility of the constituents.
Now rewriting the percolation power law in terms of resistivity we have,

= C ( crit)-t(2.10)

16

Due to the change in with hydrostatic stress the resistivity the of composite material
will also experience change and can be quantified as

( Crit )
1
1 C
=
t ( Crit ) 1
.(2.11)
P C P
P

and if we assume the critical volume fraction crit is stress invariant, it follows from Eq.11
that:

1
1 C
=
t (1 )( Crit ) 1 ( X m X C ) .(2.12)
P C P

The first term of Eqn.2.12 is the intrinsic piezoresistivity of the conducting fibers.
The second term arises from the heterogeneous nature of the material. It depends strongly
on the volume concentration of the conducting fibers. It also depends on the elastic
properties of the fibers and the matrix.

In the case where the intrinsic piezoresistivity of the conducting filler is negligible to the
piezoresistivity of the composite studied, Eqn.2.12 can be written as

1
= t (1 )( Crit ) 1 ( X m X C ) .(2.13)
P

17

The above equation signifies that if the matrix is more compressible than the conducting
filler, a negative piezoresistivity of the composite can be expected.

While modeling the piezoresistivity of heterogeneous composites under uniaxial sress


Caroma, Canet and Delhaes replaced the term (Xm XC) of Eqn.2.13 by a quantity a, an
unknown function of the elastic properties of the matrix and fillers as well as their
difference and nature of applied stress and came up with the relation

at
P .(2.14)
( crit )

where P is the applied uni-axial stress and a can be calculated from experimental data
with curve-fit method.

2.6.3 Extended Effective Medium Theory


McLachlan et al. (1990) extended the effective medium theory to describe the
conductivity of conductor-insulator composites by proposing a generalized effective
medium theory (GEM), which is derived as an interpolation between the symmetric and
asymmetric effective media theories and has the mathematical form of percolation
equation. With GEM equations they also successfully modeled piezoresistivity and
change of resistivity with chemical absorption in a semi-quantative way.

18

2.7 Summary
Based on the detailed review of published literature, closely related to the objective
of this research, the following can be summarized:

1. Polymer Concrete (PC) is being used in wide variety of applications ranging from
repair of civil infrastructure to construction of sewer pipe and manholes because
of its unique properties. These properties include rapid setting, high tensile,
compressive and flexural strength, good adhesive property, good long-term
durability with respect to cycles of freezing and thawing, low permeability to
water and aggressive solutions and good chemical resistance.
2. Addition of steel and glass fibers improved the structural properties of PC
systems. Carbon fibers improved the tensile properties of polymeric composites
and cement concrete. But very limited information is available in the literature on
the effect of carbon fibers on the structural behavior of PC systems.
3. Other than improving the structural properties, fibers impart unique functional
properties to composite. In terms of functional properties carbon fiber reinforced
composites are by far the most exceptional. Its application varies from strain
sensing to electromagnetic shielding.
4. In the past, carbon fibers have been used in polymeric and cement composites for
strain sensing. But any comprehensive study to characterize strain-sensing ability
of carbon fiber reinforced polymer concrete (CFRPC) is yet to be done.
5. There exists a very few theoretical study on the stress or strain sensing ability of
short fiber reinforced composites. Extended percolation theory and extended

19

effective medium theory has been used in the past to model strain sensing of these
composite materials.

20

CHAPTER 3
STRUCTURAL BEHAVIOR OF FIBER
POLYESTER POLYMER CONCRETE

REINFORCED

3.1 Introduction
In this chapter, the effect of addition of fibers on the structural behavior of PC
systems has been presented. The mechanical properties of interest in civil engineering
applications viz. compressive strength, tensile strength, compressive modulus, tensile
modulus and ductility were studied for polymer concrete with fibers and compared with
respective properties of polymer concrete system without fibers. Glass and carbon fibers
were used as matrix reinforcement and in both the cases the fiber content were varied up to
6% by weight. The maximum weight fraction of the fibers was limited to 6% based on
workability. At least three specimens were tested to obtain the mechanical properties under
each condition. Each polymer concrete systems (PC, GFRPC and CFRPC) had been
optimized based on strength, ductility and stiffness. The strength, stiffness and stress-strain
relationship of PC with and without fibers were modeled to predict the experimental
observations. The mechanical properties found in this chapter were used in Chapter 5 to
model the self-monitoring behavior of CFRPC.

Non-destructive tests are widely used for quality control and rapid evaluation of the
properties of civil engineering materials. The properties so estimated will be of great use in
quickly assessing the performance of a structure, which is made from this material. In this
research the sensitivity of two popular non-destructive methods, the impact resonance
method and the pulse velocity method, to characterize the PC systems with and without
21

fibers was studied by comparing the non-destructive properties with the respective static
properties. The impact resonance tests and pulse velocity tests were performed to determine
dynamic Youngs modulus, dynamic shear modulus and dynamic Poissons ratio. Damping
properties of the PC systems were also studied by varying the fiber and polymer contents
for their potential application in dynamic loading design.

3.2 Specimen Preparation


In order to optimize the composition of PC based on workability and performance,
various resin contents were investigated. Resin content varied up to 20% by weight. For
fiber reinforced polymer concrete (FRPC) systems specimens were prepared with different
fiber content. Fiber content were varied up to 6% by weight for both GFRPC and CFRPC.
For the FRPC systems, resin contents were chosen based on workability. The resin contents
chosen for this study were 18% and 20% (by weight) for GFRPC and CFRPC respectively.
The compositions of the PC system are summarized in Table 3.1.
Table 3.1- Composition of PC, GFRPC and CFRPC
Constituent Material

Function

Polyester Resin
Binder
MEKPO
Initiator
CN
Promoter
Blasting Sand
Filler
Glass Fiber
Filler
Carbon Fiber
Filler
a
by weight of polymer
b
by weight of (polymer + blasting sand)

Relative proportions
PC
GFRPC
CFRPC
up to 20%
18%
20%
a
a
1.5%
1.5%
1.5%a
a
a
0.3%
0.3%
0.3%a
up to 86%
82%
80%
b
up to 6%
up to 6%b

The isopthalic thermosetting polymer resin (Polygrad Inc., Tampa, FL) used in this
study had a viscosity of 40 to 50 poise (4 to 5 Pa-sec.) at room temperature and a specific
22

gravity of 1.07. The sand was constituted by mixing five grades of commercially available
blasting sand of equal weight. The sand was well graded and had a coefficient of
uniformity (Cu) of 5.8 and a coefficient of concavity (Cc) of 0.9. Based on manufacturers
literature, glass fibers with 6.35 mm in length and 13 m diameter had a tensile strength of
2500 MPa, tensile modulus of 70 GPa and density of 2550 kg/m3. PAN based chopped
carbon fibers with 6.35 mm in length and 7.2 m diameter were used in this study. Carbon
fibers had a tensile strength of 3800 MPa, tensile modulus of 228 GPa and density of 1810
kg/m3. The carbon fibers were 50% stronger and more than 300% stiffer than the glass
fibers.

In preparing the PC specimens, cobalt napthanate was first added to the polyester
resin and the solution was mixed thoroughly and then methyl ethyl ketone peroxide was
added. After further mixing, sand and fibers were added slowly and mixed long enough to
obtain a uniform mix (CIGMAT PC 1-02). Teflon molds were used to cast the specimens.
Cylindrical specimens, 100 mm in length and 38 mm in diameter were used for destructive
tests and 200 mm in length and 60 mm in diameter were used for non-destructive tests. The
prism specimens used for the non-destructive tests were 300 mm X 50 mm X 50 mm. For
the direct tensile test, dog-bone shaped specimens were used with proper metal grip
(CIGMAT PC 2-02). Specimens were cured at room temperature (about 250 C) for 24
hours followed by 800 C for 24 hours in an oven.

23

3.3 Testing Program


3.3.1 Destructive Tests
At least three cylindrical specimens were tested under uniaxial compression to
obtain the mechanical properties under each condition. Specimens were loaded using a 400
kips capacity universal testing machine at a displacement rate of 0.002 inch/minute [For
detailed description of testing procedure refer Appendix (CIGMAT PC 5-02)].

Similarly direct tension tests were performed on at least three dumb-bell specimen to
obtain tensile properties under each condition. The specimens were loaded using a 10 kips
capacity screw type machine at a displacement rate of 0.002 inch/minute [For detailed
description of testing procedure refer Appendix (CIGMAT PC 2-02)].

Commercially available 12 mm strain gages having least count of 10-6 strain/strain were
used to measure strains during compressive and tensile loading. Strain gages were attached
to the specimens directly. Test data was analyzed to obtain the compressive and tensile
strengths, moduli and stress-strain relationships.

3.3.2 Non-Destructive Tests


3.3.2.1

Impact Resonance Method


Impact resonance tests were performed as per ASTM C 215. The test specimens

were made to vibrate as a whole in one of their natural frequency modes: transverse,
longitudinal or torsional. The corresponding fundamental frequencies of these modes were
obtained by proper location of the support, impact point and accelerometer and the

24

response was monitored using a dynamic signal analyzer (HP 35665 A). The accelerometer
used in this study had a range of 500g and a reference sensitivity of 10.4 mV/g; it weighed
2.5 grams and had a resonant frequency of 54 kHz. The operating range of the frequency
response was 2 Hz to 10 kHz with an error of 5%. The accelerometer was mounted on a
small steel disc and glued to the specimen surface using grease. The specimen was
supported at nodes on small wooden blocks and struck by a standard hammer. A plotter
was used to plot the output from the dynamic signal analyzer.

3.3.2.2

Pulse Velocity Method


Elastic, homogeneous, and isotropic material can be characterized by two

material constants. These two elastic moduli control the velocity of the P-wave and the
shear wave. The deformation resulting from a shear wave is represented in Fig.3.1a. The
only deformation in this case is shear deformation (), so the shear modulus (G) controls
the wave velocity. The equation relating shear wave velocity (Vs) and shear modulus is as
follows

G = Vs ...(3.1)

where is the mass density of the material. The deformation caused by a P-wave is
represented in Fig.3.1b, the deformation is axial () and the wave velocity is controlled by
constrained modulus (M). The equation relating constrained modulus and P-wave velocity
(Vp) is as follows

25

M = Vp2.. (3.2)

For an elastic, homogeneous and isotropic material

(1 ) E .(3.3)
M =

(1 + )(1 2 )

where E is the Youngs modulus. At small strain levels the PC system can be assumed to
be linearly elastic, homogeneous and isotropic and the dynamic Poissons ratio of different
PC system were calculated by combining Eqn.3.1 and Eqn.3.3 as

2(1 )
=
Vs (1 2 )

Vp

1/ 2

..(3.4)

And Youngs modulus (E) as

E=

(1 + )(1 2 ) V 2 ....(3.5)
p
(1 )

Direction of particle
motion

Material
Material

Direction of wave
propagation

(a)

Direction of particle
motion

(b)

Figure 3.1-Deformation during (a) shear wave and (b) constrained P-wave
26

Direction of wave
propagation

Pulse velocities for the PC systems were measured using a commercially available
portable V-meter. Lead zirconate titanate ceramic transducers having a natural frequency of
150 kHz were used to pass longitudinal or shear waves through the specimens.
Commercially available grease was used to provide good coupling between the specimens
and the transducers. The travel time of the ultrasonic pulse through the specimens under
direct transmission (with the transducers on opposite faces along the length) was recorded
up to an accuracy of 0.1 s. Pulse velocities (either P-wave or shear wave) were calculated
by dividing the length of the specimen by the travel time and also the velocity ratio defined
as the ratio of shear wave and P-wave velocity. This data was analyzed to obtain dynamic
shear modulus, dynamic Youngs modulus and dynamic Poissons ratio.

3.4 Test Results and Discussions


3.4.1 Destructive Tests
3.4.1.1

Polymer Concrete Without Fiber


Compressive stress-strain relationships of PC with varying polymer contents are

compared in Fig.3.2. The variations of strength, modulus, and failure strain with polymer
content are shown in Fig.3.3. Though failure strain increased with increasing polymer
content, strength remained constant while the modulus decreased after reaching a peak
value at 14% polymer content and hence 14% PC was considered the optimum system.
Since increasing polymer content in the PC system reduced the aggregate content, it
resulted in lower modulus. Behavior models have been proposed by Cohen and Isahi22 and,
Hirsch23 for optimum combination, and modified by Vipulanandan and Dharmarajan21 for
excess polymer system. Optimum PC systems can be idealized as either aggregate and

27

porous polymer distributed in a series (Fig.3.4a) or as a combination of both series and


parallel distribution (Fig.3.4b). In excess polymer systems the polymer content over and
above the optimum polymer content is the excess polymer for that system. In excess
polymer system aggregate, optimum porous polymer and excess polymer can either be
distributed in a series (Fig 3.5a) or as a combination of both series and parallel distribution
(Fig 3.5b). In the case of optimum system, if the constituents are distributed in a series, the
modular ratio can be represented as (Cohen and Isahi, 1981)

1
E PC
(3.6)
=
P
E
1 X A

1 C 2 / 3 + mX A
V

where CV = XV/( XV + XP) is the porosity of polymer matrix and XV and XP are the void
content and polymer content by volume; the ratio of polymer modulus (EP = 6.8 GPa) to
aggregate modulus (EA = 700 GPa) is denoted by m; and XA is the solid content (ratio of
aggregate volume to total volume of polymer concrete).

If the constituents are distributed in a combination of both series and parallel then the
modular ratio can be represented as (Model 1) (Cohen and Isahi, 1981)

E PC
=
EP

(1 X A )
XA

2/3
Z
mX
+
+ (1 Z )(1 X A ) 1 CV +
A
2/3
m

1 CV

28

(3.7)

70

20 % Polymer

14 % Polymer

60

Stress (MPa)

50
40

18 % Polymer

30

Poly. (20 % Polymer)


Poly. (18 % Polymer)
Poly. (14 % Polymer)

20
10
0
0

0.2

0.4

0.6
Strain (%)

0.8

1.2

Figure 3.2-Compressive stress-strain relationships of polyester PC systems


30

(b)

(a)

60

Modulus (GPa)

Strength (MPa)

80

40

20

10

20

0
0

10
20
Polymer Content (%)

30

10
20
Polymer Content (%)

30

(c)
Failure Strain (%)

0.8
0.6
0.4
0.2
0
0

10
15
20
Polymer Content (%)

25

Figure 3.3-Variation of properties with polymer content for polyester PC systems

29

where Z and (1-Z) are the relative proportions of a series and parallel elements. The seriesparallel model predicted the PC modulus reasonably well for optimum system (14%
polymer content) with Z = 0.625. For excess polymer system the effect of excess polymer
is taken into account, if distributed in a series, the modular ratio can be represented as
(Vipulanandan and Dharmarajan, 1987)

1
E PC
=
.(3.8)
P
E
1 X A'

+ mX ' + X EP
A
1 C 2 '/ 3
P

where XA and CP are the solid content and porosity considering optimum polymer content
(14%) and XEP is the excess polymer content over and above the optimum content, and, if
distributed in a series and parallel combination, the relationship is as follows, (Model 2)
(Vipulanandan and Dharmarajan, 1987)

E PC
=
EP

(1 X A' )

2/3
Z
mX
+ A' + X EP
+
' + X EP + (1 Z ) (1 X ' ) 1 C ,
2/3

A
A
P
m

1 C P '

. (3.9)

This excess polymer model predicted the compressive moduli of 18% PC and 20%
PC reasonably well with the same relative proportion of a series and parallel elements (Z =
0.625) [Fig.3.6]. By combining the optimum polymer model [Eqn.3.7] and excess polymer
model [Eqn.3.9], the theoretical peak value of the compressive modulus of the PC system

30

was obtained [Fig.3.6]. The theoretical peak was in good agreement with the experimental
data and hence, 14% PC was considered as the optimum system.

The tensile and compressive stress-strain relationships for 14% PC is compared in


Fig.3.7. The tensile-to-compressive strength ratio for 14% PC system was 0.13, the tensileto-compressive modular ratio was 0.75 and the tensile-to-compressive toughness (area
under stress-strain curve) ratio at failure was 0.01. Based on the modular ratio, PC systems
can be characterized as bi-modulus material.

Porous
Polymer

Aggregate

(a)

1- Z

(b)

Figure 3.4-Behavior models for optimum PC system (a) series (b) series-parallel

Excess
Polymer

1-Z

(a)

(b)

Figure 3.5-Behavior models for excess polymer system (a) series (b) series-parallel

31

28
Test Data 14% Polymer Content
Test Data 18% Polymer Content

Modulus, E (GPa)

Test Data 20% Polymer Content

Optimum Polymer System


(Model 1)

24

Theoretical peak
20

Excess Polymer System (Model 2)


16
10

12

14

16

18

20

22

Polymer Content (%)

Figure 3.6-Comparison of predicted and experimental compressive modulus of PC


(b) Transverse mode (c) Torsional mode

60

14% PC System
50

Stress (MPa)

40
0

0.02

0.04

30

Compressive Behavior
2

20
1

Tensile Behavior

10
0

0
0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

Strain (%)

Figure 3.7-Comparison of tensile and compressive behavior of 14% PC system

32

90
PC with 6% Glass Fiber
75

PC with 4% Glass Fiber

Stress (MPa)

60
45

PC (without Glass
Fiber)

30

18% Polymer System

15
0
0

0.2

0.4

0.6

0.8

1.2

Strain (%)

Figure 3.8-Compressive stress-strain relationships of GFRPC systems

3.4.1.2

Polymer Concrete with Glass Fibers


Typical compressive stress-strain relationships of GFRPC with varying glass

fiber contents are shown in Fig.3.8. The variations in strength, failure strain and modulus
with glass fiber content are shown in Fig 3.11 (a, c and e). Maximum fiber content in the
GFRPC system was limited to 6% w/w based on workability. The addition of 6% glass
fibers in PC systems increased the compressive strength; strain at peak stress and
compressive modulus by 45%, 50%, and 10% respectively.
Tensile properties were also improved with the addition of glass fiber, where an
85% increase in the tensile strength of the PC system was observed with the addition of 6%
glass fiber. The tensile behavior of the optimum combination of the GFRPC system (6%
glass fiber in 18% polymer) is shown in Fig.3.9.

33

Bi-modulus behavior was observed for the GFRPC system also. The tensile-tocompressive strength ratio for a 6% GFRPC system was 0.16, the tensile-to-compressive
modular ratio was 0.85, and the tensile-to-compressive toughness ratio at failure was 0.018.

3.4.1.3

Polymer Concrete with Carbon Fibers


Fig 3.10 shows the typical compressive stress-strain relationship of the CFRPC

with varying carbon fiber contents. The variation in strength, failure strain, and modulus
with carbon fiber content is shown in Fig 3.11 (b, d and f). The test results showed that the
addition of carbon fibers did not improve the compressive properties of the PC system.
High modulus PAN based carbon fibers in the polymer matrix (fiber-to-matrix modular
ratio of 33.5) were not effective in improving the compressive properties of the PC system.

The tensile stress-strain properties of the CFRPC system with varying fiber
contents are shown in Fig.3.12. The variations of the tensile properties with fiber content
for the CFRPC systems are shown in Fig.3.13 (a, b and c). The addition of 6% carbon
fibers to the PC system (20% polymer) increased the tensile strength; strain at peak stress;
and tensile modulus by 60%, 200% and 5% respectively. Based on the tensile properties
and workability, the fiber content in the CFRPC system was limited to 6 % w/w.

Bi-modulus behavior was observed for the CFRPC system also. The tensile-tocompressive strength ratio for the 6% CFRPC system was 0.18, the tensile-to-compressive
modular ratio was 0.78, and the tensile-to-compressive toughness ratio at failure was 0.029.

34

14
12

Stress, MPa

10

GFRPC 6 % Fiber

CFRPC 6 % Fiber

6
4
2

PC- 14 % Polymer

0
0

0.02

0.04

0.06
0.08
Strain, %

0.1

0.12

0.14

Figure 3.9-Tensile stress-strain properties of optimum combination of


PC, GFRPC and CFRPC system

70

PC with 6% Carbon Fiber

60

Stress (MPa)

50
40

PC with 4% Carbon Fiber

PC (without
Carbon Fiber)

30
20

20% Polymer System

10
0
0

0.2

0.4

0.6
Strain (%)

0.8

1.2

Figure 3.10-Compressive stress-strain relationships of CFRPC systems

35

PC with Glass Fibers


100

Compressive Strength ( MPa)

PC with Carbon Fibers


(20% Polymer Systems)

(18% Polymer System)


100

(a)
90

90

80

80

70

70

60

60
50

50
0

Failure Strain (%)

(b)

4
Fiber Content (%)

0.9

0.9

0.8

0.8

0.7

0.7

0.6

4
Fiber Content (%)

0.6

(c)

(d)

0.5

0.5
0

Fiber Content (%)

Fiber Content (%)

Modulus (GPa)

25

25

20

20

15

15

10

10

(e)

(f)

Fiber Content (%)

Fiber Content (%)

Figure 3.11-Variation of compressive properties with fiber content for


GFRPC and CFRPC
36

14

6% CFRPC

20% Polymer System

12

Stress (MPa)

10
8
6

0% CFRPC

4
2
0
0

0.02

0.04

0.06

0.08

0.1

0.12

0.14

Strain (%)

Figure 3.12-Direct tensile stress-strain properties of CFRPC systems

3.4.2 Non-Destructive Tests


Impact resonance tests were performed on both cylindrical and prismatic specimens
to study the effect of shape of the specimens on the measured material properties. The
typical frequency spectrum for three modes of vibration (longitudinal, transverse, and
torsional) of 6% GFRPC is shown in Fig.3.14 (a, b and c). From the measured resonant
frequencies, the dynamic Youngs modulus, dynamic torsional modulus, and dynamic
Poissons ratio were calculated as per ASTM C215 as follows:

37

16
Tensile Strength (MPa)

14
12
10
8
6
4
2

GFRPC data

0
0

3
4
Fiber Content (%)

Tensile Modulus (GPa)

(a)
16
12
8
4

GFRPC data

0
0

3
4
Fiber Content (%)

(b)
0.14
Failure Strain (%)

0.12
0.1
0.08
0.06
0.04
0.02

GFRPC data

0
0

3
4
Fiber Content (%)

(c)
Figure 3.13-Variation of tensile properties with fiber content for CFRPC systems

38

Table 3.2 Static Test Results of PC, GFRPC and CFRPC


Model Value
Peak
Strength
(MPa)
55
56
55
55

Peak
Strain
(%)
0.45
0.45
0.45
0.56

Modulus
(GPa)
19.44
19.5
19.5
19.9

55

0.45

0.04

0.63

50

0.55

19.6

55

0.56

0.07

0.49

52

0.54

19.6

Specimen- 1
20 % Polymer Specimen- 2

61

0.78

18.57

59

0.77

18.21

61

0.85

0.08

0.39

Specimen- 3

57

0.78

18.29

Specimen- 1
Tension Test 14 % Polymer Specimen- 2

0.055

15

7.5

0.05

14.5

0.055

0.015

0.85

Specimen- 3

0.055

15

Specimen- 1

55

0.56

19.9

Specimen- 2

50

0.55

19.6

55

0.56

0.07

0.49

Specimen- 3

52

0.54

19.6

Specimen- 1

65

0.8

18.57

Specimen- 2

63

0.78

18.2

65

0.8

0.06

0.44

Specimen- 3

68

0.81

18.5

Specimen- 1

80

0.85

21
80

0.85

0.3

0.45

13

0.105

0.095

0.55

61

0.85

0.05

0.39

60

0.9

0.05

0.37

65

0.95

0.05

0.39

7.3

0.063

0.05

0.86

9.3

0.075

0.065

0.86

Specimen- 1
14 % Polymer Specimen- 2
Specimen- 3
Specimen- 1
Compression 18 % Polymer Specimen- 2
Test
Specimen- 3
PC

0 % Fiber

Compression
Test

4 % Fiber

GFRPC
6 % Fiber

Tension Test

6 % Fiber

0 % Fiber

Compression
Test

4 % Fiber

6 % Fiber

CFRPC

0 % Fiber

Tension Test

Specimen- 2

82

0.88

21.6

Specimen- 3

58

0.38

21.8

Specimen- 1

13

0.105

18.5

Specimen- 2
Specimen- 3

12.5
13.5

0.1
0.11

18.4
18.6

Specimen- 1

61

0.78

18.57

Specimen- 2

59

0.77

18.21

Specimen- 3

57

0.78

18.29

Specimen- 1

60

0.9

17.8

Specimen- 2

55

0.79

17.1

Specimen- 3

61

0.88

17.9

Specimen- 1

65

0.95

17.5

Specimen- 2

68

0.92

17.6

Specimen- 3

65

0.98

17

Specimen- 1

7.33

0.063

13.5

Specimen- 2

7.5

0.071

13.1

Specimen- 3

7.1

0.059

13.3

9.5
9.27
9.2
10.55

0.08
0.075
0.072
0.095

14.1
14.3
13.9
14.4

0.095
0.095
0.19

14.4
14.3
14

10.6

0.095

0.065

0.85

Specimen- 1

10.6
10.7
11.5

Specimen- 2

11.3

0.195

13.8

11.5

0.19

0.07

0.82

Specimen- 3

11.5

0.19

14

Specimen- 1
2.5 % Fiber Specimen- 2
Specimen- 3
Specimen- 1
4 % Fiber

6 % Fiber

Specimen- 2
Specimen- 3

39

From transverse mode of vibration: Dynamic E = CM (n ) ....(3.10)


2

( )

From longitudinal mode of vibration: Dynamic E = DM n ' ...(3.11)

( )

From torsional mode of vibration: Dynamic G = BM n '' ..(3.12)

E
Dynamic =
1 (3.13)
2G

where n, n,, n,, are the respective resonance frequencies in transverse, longitudinal and
torsional mode and C, D, and B are shape correction factors (Ref.4). The dynamic Youngs
moduli calculated from longitudinal mode of vibration were 16 to 20% higher than the
moduli from the transverse mode of vibration for all the PC systems studied. This
reinforces the earlier observation that PC, GFRPC and CFRPC systems are bi-modulus
materials having different moduli in compression and tension, where direct compression
moduli were 15 to 25% higher than the direct tension moduli for the PC, GFRPC and

CFRPC systems studied. Though the shape of specimens did not have significant
effect on the Youngs modulus, it had substantial effect on the shear modulus, with the
prism specimens showing about 20% higher values than cylindrical specimens; also, very
large variations in the dynamic Poissons ratio were observed, with the cylindrical
specimens yielding higher values.

40

Amplitude

(a)

accelerometer
support
impact point
0

Frequency (kHz)

12.8

Amplitude

(b)

accelerometer
support
impact point
0

Frequency (kHz)

12.8

Amplitude

(c)

accelerometer
support
impact point
0

Frequency (kHz)

12.8

Figure 3.14-Typical frequency spectrum (X-f) for 6% GFRPC cylinders


(a) Longitudinal mode (b) Transverse mode and (c) Torsional mode

41

Amplitude

(a)

accelerometer
support
impact point
0

Time (millisecond)

31.2195

Amplitude

(b)

accelerometer
support
impact point
0

Time (millisecond)

31.2195

Amplitude

(c)

accelerometer
support
impact point
0

Time (millisecond)

31.2195

Figure 3.15-Typical time decay (X-t) relationships for 6% GFRPC cylinders


(a) Longitudinal mode (b) Transverse mode and (c) Torsional mode

42

Table 3.3 Non-Destructive Tests Results of Prismatic Specimens


Impact Resonance
Long.
Wave
EiL
Velocity
No. (GPa) (m/sec.)

PC

GFRPC

CFRPC

Pulse Velocity

Shear
Wave
Gi Velocity
EiTr
(GPa) (GPa) (m/sec) i

P-wave Shear
velocity wave
(Vp) velocity
(m/sec) (Vs)

EP
Gp
(GPa) (GPa)

14%
Polymer

1 21.92
2 21.90

3210 17.27
3199 17.86

9.20
9.99

1932 0.19
2027 0.10

3226
3215

2027
2020

0.17
0.17

8.92 20.93
8.91 20.92

18%
Polymer

1 20.94

3114 18.18 10.22

1980 0.02

3297

2041

0.19

9.18 21.83

2 22.47

3234 17.57 10.90

2010 0.03

3279

2069

0.17

9.39 21.95

20%
Polymer

1 21.13

3250 17.79

8.84

1993 0.20

3268

2041

0.18

8.5 20.07

2 21.79

3234 17.14 10.22

2016 0.07

3261

2055

0.17

8.98 21.02

0%
Fiber

1 20.94

3114 18.18 10.22

1980 0.02

3297

2041

0.19

9.18 21.83

2 22.47

3234 17.57 10.90

2010 0.03

3279

2069

0.17

9.39 21.95

4%
Fiber

1 20.32

3227 17.10

9.47

1987 0.07

3233

2020

0.18

8.13 19.17

2 21.77

3246 17.58

9.93

2016 0.10

3293

2013

0.2

8.55 20.54

6%
Fiber

1 22.37

3332 18.97 10.42

2091 0.07

3389

2095

0.19

9.03 21.49

2 22.48

3334 18.72

9.79

2066 0.15

3393

2097

0.19

9.08 21.62

0%
Fiber

1 21.13

3250 17.79

8.84

1993 0.20

3268

2041

0.18

8.5 20.07

2 21.79

3234 17.14 10.22

2016 0.07

3261

2055

0.17

8.98 21.02

1 21.56

3210 17.39 10.24

2034 0.05

3257

2041

0.18

8.99 20.93

2 21.06

3256 17.82 10.25

2023 0.03

3272

2041

0.18

8.44 19.95

1 20.29

3210 17.95

9.61

1992 0.06

3243

2027

0.18

8.25 19.47

2 21.08

3257 17.89

9.89

2011 0.07

3279

2076

0.17

8.74 20.37

4%
Fiber
6%
Fiber

Since resonant frequencies are not material property, resonant wave velocities
(material property) for PC, GFRPC and CFRPC systems were calculated and later
compared with P-wave and shear wave velocities obtained from the pulse velocity test to
study the shape effect of the specimens. The wave velocity (V) can be determined by

V = fn . n ...(3.14)

where wavelength, n = 2 L for longitudinal and torsional fundamental modes. Obtaining


the resonant frequencies in longitudinal and torsional modes of vibration from the impact
resonance tests, and using Eq.3.14 resonant P-wave and shear wave velocities of PC,
GFRPC, and CFRPC systems were calculated and summarized in Table 3.3 for prism

43

specimens and in Table 3.4 for cylindrical specimens. The typical X-t plots for three modes
of vibration of 6% GFRPC are shown in Fig.3.15 (a, b and c). From these plots the
damping ratio for the corresponding modes were calculated and summarized in Table 3.5
for both prismatic and cylindrical specimens. The damping ratio was calculated by
logarithmic decrement method as follows

= /2 where = ln (yn/yn+1).....(3.15)

1.6
1.4

Damping Ratio (%)

1.2
1
0.8
0.6
0.4
0.2
0
1

5
Cycles (Nos)

11

Figure 20-Damping ratio of 6% GFRPC Cylindrical specimen


in longitudinal mode

The average damping ratio was calculated from the X-t plots by considering at least
10 cycles as shown in Fig.3.16. Overall damping ratio increased with increasing polymer
content and in general, glass fiber PC had a higher damping ratio than the carbon fiber PC

44

systems. For the CFRPC systems, though the longitudinal damping ratio increased with
addition of carbon fibers, it had a negative effect on the transverse and torsional damping
ratio. The damping ratio in the longitudinal mode was 2% for 6% CFRPC systems,
compared to 1.3% for 6% GFRPC systems. The transverse mode damping ratio was 1.8%
for 6% GFRPC systems and in torsional mode 4.96% for 6% GFRPC systems.

Table 3.4 Non-Destructive Tests Results of Cylindrical Specimens


Impact Resonance
Pulse Velocity
Long.
Shear
P-wave Shear
Wave
Wave
velocity wave
EiL
EiTr
Gi Velocity
EP
Gp
Velocity
(Vp) velocity
No. (GPa) (m/sec) (GPa) (GPa) (m/sec) i (m/sec) (Vs)
p (GPa) (GPa)
14%
1 21.09
3198 18.7 8.16
1989 0.29
3250
2031 0.18 8.69 20.49
Polymer
2 21.37
3224 18.7 8.19
1995 0.31
3262
2052 0.17 8.84 20.73
PC

GFRPC

CFRPC

18%
Polymer

1 21.75

3234

17.9

8.06

1968 0.35

3339

2074

0.19

9.13 21.65

22.5

3235

18.9

8.49

1986 0.33

3397

2096

0.19

9.64 22.99

20%
Polymer

1 20.82

3186

17.1

7.87

1959 0.32

3275

2026

0.19

8.59 20.44

2 20.98

3206

17.3

8.01

1980 0.31

3257

2020

0.19

8.5 20.19

0%
Fiber

1 21.75

3234

17.9

8.06

1968 0.35

3339

2074

0.19

9.13 21.65

3235

18.9

8.49

1986 0.33

3397

2096

0.19

9.64 22.99

4%
Fiber

1 21.54

3182

18.8

8.55

2003 0.26

3210

2031

0.17

8.95 20.89

2 21.64

3176

18.9

8 .5

1990 0.27

3236

2031

0.18

9.03 21.22

6%
Fiber
0%
Fiber

1
2
1
2

24.05
23.51
20.82
20.98

3328
3302
3186
3206

20
20.4
17.1
17.3

8.57
9.42
7.87
8.01

1985 0.40
2090 0.25
1959 0.32
1980 0.31

3363
3361
3275
3257

2111
2105
2026
2020

0.17
0.18
0.19
0.19

9.88
9.76
8.59
8.5

4%
Fiber

1 21.84
2 20.29

3217
3193

18.4
17.6

8.46
7.88

2001 0.29
1990 0.29

3273
3257

2010
2031

0.2
0.18

8.7 20.84
8.37 19.79

6%
Fiber

1 21.05

3197

18.2

7.98

1968 0.32

3230

2031

0.17

8.67 20.34

2 21.26

3196

17.9

8.07

1969 0.32

3248

2036

0.18

8.81 20.72

22.5

23.22
22.97
20.44
20.19

Pulse velocity tests were also performed on both cylindrical and prismatic
specimens. Wavelengths of the test P-wave pulses were in the range of 20-23 mm, ten
times the largest aggregate size used in test specimens, having a minimum dimension of 50
mm. Wavelengths of the test shear wave pulses were in the range of 13-14 mm. Measuring

45

the wave velocities, wavelengths were calculated from known test pulse frequency using
Eq.3.14. From the P-wave and shear wave velocity, dynamic Poissons ratio was first
calculated from Eq.3.4 and then dynamic Youngs modulus and dynamic shear modulus
were calculated from Eq.3.5 and Eq.3.1 respectively. This was another way to calculate the
Youngs modulus from pulse velocity method other than ASTM C 597, where calculation
of Youngs modulus requires assumption of Poissons ratio. Shear modulus can also be
calculated from this method. From the pulse velocity test, P-wave velocities (Vp) and shear
wave velocities (Vs) for PC, GFRPC and CFRPC systems were measured and shear wave
velocity to P-wave velocity ratios (Vs/Vp), dynamic Youngs moduli, dynamic shear
moduli and dynamic Poissons ratios were calculated and are summarized in Table 3.3 for
prism specimens and in Table 3.4 for cylindrical specimens. The test result showed that the
dynamic Youngs modulus, dynamic shear modulus, as well as the dynamic Poissons ratio
was independent of specimen shape. The velocity ratio (Vs/Vp) for 14% PC system was
0.63 and this ratio was almost independent to polymer and fiber addition in the PC system
as well as to shape of the specimen.

The longitudinal and shear wave velocities obtained from impact resonance tests for
both cylindrical and prism specimens for all the PC, GFRPC and CFRPC systems studied
were on an average 2-3% less than those calculated from pulse velocity tests; this was
partly because the test wavelengths in impact resonance tests were higher than those of
pulse velocity test. While the test wavelengths in the pulse velocity tests were between 1323 mm, the test wavelengths in impact resonance tests were in the range of 400-600 mm.
Since longitudinal wave velocity is directly proportional to dynamic Youngs modulus

46

[Eq.3.5], and shear wave velocity is directly proportional to dynamic shear modulus
[Eq.3.1], dynamic Youngs moduli and dynamic shear moduli, calculated from impact
resonance test should be less than those calculated from pulse velocity tests. But the
dynamic Youngs moduli and dynamic shear moduli, calculated from impact resonance
tests using ASTM C 215 correction factors for the PC, GFRPC and CFRPC systems
studied were on an average 10% higher than respective moduli calculated from pulse
velocity tests, except for shear moduli for cylindrical specimens, where impact resonance
method predicted a lower value.

Table 3.5- Damping Ratios of PC, GFRPC and CFRPC


Longitudinal
Flexural Damping Ratio
Damping Ratio (%)
(%)

PC

Specimen Prism

Cylinder

Prism

Cylinder

Prism

Cylinder

14 %
Polymer

0.80

0.92

1.15

1.32

2.74

2.84

0.88

0.92

1.03

1.32

2.82

2.79

18 %
Polymer

0.92

0.95

1.3

1.41

2.9

2.95

0.86

0.95

1.2

1.38

2.83

2.89

20 %
Polymer

1.18

1.2

1.78

1.65

3.32

3.51

1.1

1.1

1.82

1.79

3.41

3.39

0.92

0.95

1.3

1.41

2.9

2.95

0.86

0.95

1.2

1.38

2.83

2.89

1.01

1.13

1.6

1.65

3.81

3.99

1.05

1.19

1.52

1.49

3.72

4.01

1.17

1.26

1.75

1.8

4.96

4.9

1.23

1.3

1.7

1.76

4.91

4.85

1.18

1.2

1.78

1.65

3.32

3.51

0 % Fiber
GFRPC

4 % Fiber
6 % Fiber
0 % Fiber

CFRPC

Torsional Damping
Ratio (%)

4 % Fiber
6 % Fiber

1.1

1.1

1.82

1.79

3.41

3.39

1.71

1.82

0.3

0.23

2.2

2.41

1.69

1.77

0.27

0.38

2.29

2.53

1.98

2.01

0.35

0.3

2.42

2.53

1.89

1.92

0.32

0.33

2.3

2.44

47

This is contradicting but partly explainable since the correction factors used in
ASTM C 215 are approximate12. The equation relating dynamic Youngs modulus and
longitudinal wave velocity for impact resonance test [combining Eqn.3.11 and Eqn.3.14]
and the equation relating dynamic Youngs modulus and longitudinal wave velocity for
pulse velocity tests [Eqn.3.5] are compared in Fig.3.17 and Fig.3.19 for prism and
cylindrical specimens respectively along with experimental results. Similarly, the equation
relating dynamic shear modulus and shear wave velocity for impact resonance tests
[combining Eqn.3.12 and Eqn.3.14] and the equation relating dynamic shear modulus and
shear wave velocity for pulse velocity tests [Eqn.3.1] are compared in Fig.3.18 and
Fig.3.20 for prism and cylindrical specimens respectively, along with experimental results.
These plots showed that for the same wave velocity, impact resonance tests predict higher
moduli (except for shear moduli for cylindrical specimen) and pulse velocity tests predict
lower moduli. Though the discrepancies in Youngs moduli and shear moduli were within
acceptable limit, the Poissons ratios were much more sensitive. The dynamic Poissons
ratios calculated from the impact resonance method for prism specimens were far less than
those calculated from pulse velocity method and for cylindrical specimens they were about
30% higher. The dynamic Poissons ratios calculated from pulse velocity tests matched
reasonably well with static Poissons ratio reported in the literature13, 17 for PC systems.

While comparing the Youngs moduli obtained from static tests with those obtained
from dynamic tests, the dynamic moduli for all the PC systems were higher than the
respective static moduli; this was due to the fact that moduli were determined at low strain
(order of 10-6) in the dynamic tests. The moduli obtained from the pulse velocity method

48

predicted the static moduli more closely than the moduli from the impact resonance
method.

Young's Modulus, E (GPa)

25

Impact Resonance (Eqn.3.11 & Eqn.3.14 coupled)

20

15

Pulse Velocity (Eqn.3.5 with = 0.2)


10

0
2600

Test data - Pulse Velocity


Test Data - Impact Resonance
2700

2800

2900

3000

3100

3200

3300

3400

3500

Longitudinal Wave Velocity (m/sec)

Figure 3.17-Comparison of Youngs moduli obtained from pulse velocity


and impact resonance tests for prism specimens
16

Shear Modulus, G (GPa)

14

Impact Resonance (Eqn.3.12 & Eqn.3.14 coupled)

12
10
8

Pulse Velocity (Eqn.3.1)

6
4

Test data - Pulse Velocity

Test Data - Impact

0
1500

1700

1900

2100

2300

2500

Shear Wave velocity (m/sec)

Figure 3.18-Comparison of shear moduli obtained from pulse velocity


and impact resonance tests for prism specimens

49

30

Impact Resonance (Eqn.3.11 & Eqn.3.14coupled)


Young's Modulus, E (GPa)

25

20

Pulse Velocity (Eqn.3.5 with = 0.2)


15

10

Test Data - Impact Resonance


5

0
3000

Test Data - Pulse Velocity


3100

3200

3300

3400

3500

3600

Longitudinal Wave Velocity (m/sec)

Figure 3.19-Comparison of Youngs moduli obtained from pulse velocity


and impact resonance tests for cylindrical specimens
12

Shear Modulus, G (GPa)

Pulse Velocity (Eqn.3. 1)

10

Impact Resonance (Eq.3.12


& Eq.3.14 coupled)
8

Test Data - Pulse Velocity


Test data - Impact Resonance
6
1700

1900

2100

2300

2500

Shear Wave Velocity (m/sec)

Figure 3.20-Comparison of shear moduli obtained from pulse velocity


and impact resonance tests for cylindrical specimens

50

3.5 Strength Model


Inspection of failure surfaces of the fiber reinforced polymer concrete in tension
indicated adhesive and cohesive failure of polymer with mainly fiber pullout; hence,
strength model, as suggested by Dharmarajan (1987) was used to predict the tensile
strength of the CFRPC systems. The tensile strength of FRPC systems can be represented
as

Incremental Strength Ratio

0.080

K1 (GFRPC) = 1.42

0.060

0.040

K1 (CFRPC) = 0.86

0.020

CFRPC Tension Test Data


GFRPC Tension Test Data

0.000
0

Fiber Volume Fraction, %

Figure 3.21-Experimental and predicted tensile strength of CFRPC and GFRPC

tFRPC = (AT)(/4) + ( 1-/4)(1-CV)tP + AS Xf(3.16)

and considering the change in strength (Model 3)

(tFRPC tPC)/tfP = K1 Xf

51

K1=S / tfP(3.17)

where tFRPC = tensile strength of fiber reinforced polymer concrete; tPC = tensile
strength of polymer concrete without fibers; tP = tensile strength of polymer;
AT = adhesive tensile strength between aggregate and polymer; S = adhesive shear
strength between fiber and polymer; CV = porosity of the polymer matrix; = fiber
efficiency factor ; Xf = fiber content by volume. Hence incremental tensile strength ratio of
the CFRPC system was directly proportional to the volume fraction of fibers. The
comparison of model predictions with experimental data for both GFRPC and CFRPC
systems are shown in Fig.3.21. The factor K1 for GFRPC systems was 1.42, whereas for
CFRPC systems it was 0.86.

3.6 Stress-Strain Model


A model (Vipulanandan and Paul, 1990; Vipulanandan and Paul, 1993; Mebarkia,
1993; Mantrala, 1996) that has been used in the past to describe PC behavior is proposed to
predict the stress-strain behavior of PC, GFRPC and CFRPC systems in both tension and
compression. The proposed model is (Mantrala, 1996)

c
=
( p + q ) c ..(3.18)

q + (1 p q ) + p

c
c

where p and q = material parameters; c and c = peak stress and strain at peak stress; and
= stress and corresponding strain.

52

By satisfying the condition of initial slope of the stress-strain curve ( = 0), the
following relationship can be obtained
q=

Es
.(3.19)
Ei

where Es is the secant modulus at peak and Ei is the initial tangent modulus. The parameter
p can be calculated using regression analysis (least square method). A typical way of
calculating the material constant p is shown in Fig.3.22. From the known values of c, c, q
and assumed value of p, stress was predicted at a particular strain level and the square of
error of predicted stress with the experimental stress was calculated. This procedure was
repeated assuming different values of p and a plot as shown in Fig.3.22 was obtained. The
value of p corresponding to the least square of error was assumed to be the material
constant p for that material. The model parameter q signifies the ratio of secant modulus at
peak to initial tangent modulus and q =1 signifies linear material up to peak stress; hence,
the lower value of q represents more non-linear material. On the other hand, the parameter
p controls mainly the post peak behavior, though it influences the pre-peak behavior also.
Parameter p can be related to the toughness of the material and in Fig.3.24 the variation of

d
c c
0

pre-peak normalized toughness,



d
c c
1

1.5

and post-peak normalized toughness,

with the material parameter p is compared. Both pre-peak and post-peak

toughness increased with an increase in the material parameters p and q. But q has the
greatest influence on the pre-peak toughness, whereas p controls the post-peak toughness
and also the influence of p on the pre-peak toughness of the material is negligible in the
range of p obtained for the PC, GFRPC, and CFRPC systems studied; hence for PC

53

systems material constant p was used to compare the post-peak toughness. The
experimental and predicted stress-strain behavior of 4% and 6% CFRPC is compared in
Fig.3.22a and Fig.3.23b respectively. The modulus of elasticity, strength, failure strain and
model values c, c, p and q for different PC, GFRPC, and CFRPC system are summarized
in Table 3.2.
0.025

Square of error

0.02

0.015

0.01

0.005

0
0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

value of p

Figure 3.22-Material constant p for 2.5% CFRPC in tension

For PC systems, the increase in polymer content above 14% w/w did not increase
the peak strength, but material parameter q decreased and p increased. Reduction in q
indicates the fact that material is becoming more non-linear with an increase in polymer
content. An increase in parameter p with polymer content indicates that post peak
toughness increased with polymer content. Similarly, for the GFRPC systems, the initial
stiffness remained more or less the same with an increase in fiber content and hence q
remain unchanged but post-peak toughness increased substantially (increase in parameter

54

p) with the addition of 6% glass fibers. For CFRPC systems, there was not much change in
initial stiffness as well as toughness with the addition of carbon fibers, since both the
parameters p and q remained more or less constant. Similarly, comparing the tensile model
values (p and q) of optimum combinations of PC, GFRPC, and CFRPC systems we can
conclude that GFRPC was the stiffest (comparing q) and also the toughest (comparing p).

3.7 Summary
For PC systems, the optimum combination (lowest polymer content at which strength
and/or modulus are maximum) was obtained at 14% polymer content. For the GFRPC
system, the mechanical properties improved with the addition of glass fibers but based on
workability, the optimum combination was 6% glass fibers with 18% polymer. For the
CFRPC system, though the compressive properties were almost independent to fiber
addition, tensile properties improved significantly. Based on the tensile properties and
workability, for the CFRPC systems the optimum system was with 6% carbon fibers and
20% polymer. Based on the destructive and non-destructive tests performed on over fortyfive PC, GFRPC and CFRPC specimens the following can be summarized.
1. Polymer concrete is a bi-modulus material with a tensile-to-compressive modular
ratio of 0.75. A combination of series and parallel iso-stress models, predicted the
modulus of PC systems reasonably well.
2. The tensile strength of the PC system improved by 85% and 60% with the addition
of 6% glass fibers and 6% carbon fibers respectively. The tensile strength model
predicted the strength increase in PC systems with fibers. Glass fiber addition
improved the compressive strength of polymer concrete, but carbon fibers did not.

55

70

PC with 4% Carbon Fibers

60

Stress (MPa)

50
40
30
20

Predicted (Eq.3.18)
10

Experimental

0
0

0.2

0.4

0.6

0.8

Strain (%)

80

PC with 6% Carbon Fibers

70

Stress (MPa)

60
50
40
30

Predicted (Eq.3.18)

20

Experimental

10
0
0

0.2

0.4

0.6

0.8

1.2

Strain (%)

Figure 3.23-Predicted and measured compressive stress-strain relationships for


CFRPC systems with (a) 4% fibers (b) 6% fibers

56

0.8

Normalized Toughness

0.7

Pre-peak toughness

0.6
0.5
0.4

Post-peak toughness

0.3
0.2

Range of material parameter p


in different PC systems

0.1

q = 0.8
q = 0.4

0
0

0.2

0.4

0.6
Value of p

0.8

1.2

Figure 3.24 -Variation of toughness with material constant p

3. Polymer concrete with glass and carbon fibers also behaved as a bi-modulus
material with a tensile to compressive modular ratio of 0.85 and 0.78 respectively.
4. The p-q model for stress-strain relationships predicted both the pre-peak and postpeak behavior with a single function for both tension and compression. While the
parameter q influenced the stress-strain relationships before peak, the parameter p
influenced the toughness, mainly the post-peak toughness.
5. The impact resonance and the pulse velocity methods can be used to characterize
the behavior of PC with and without fibers. The ratio of shear wave velocity to Pwave velocity for 14% PC was 0.63. The addition of polymer or fibers did not
influence the velocity ratio. While the pulse velocity method was independent of
specimen shape, the impact resonance method was dependent on specimen shape

57

with dynamic shear modulus and dynamic Poissons ratio having the greatest effect.
Dynamic Youngs moduli and dynamic shear moduli obtained from the pulse
velocity method were within 10% of the respective static moduli; whereas those
from the impact resonance test showed larger variations. Dynamic Poissons ratio
obtained from the pulse velocity method was closer to static Poissons ratio. Wave
velocities measured from the pulse velocity test were higher than those determined
from the impact resonance method.
6. The damping ratio increased with increase in polymer and fiber contents in the PC
systems. Glass fibers increased the damping ratio in the longitudinal, flexural, and
torsional modes. Carbon fibers increased the longitudinal damping ratio but it also
reduced the flexural and torsional damping ratios.

58

CHAPTER 4
SELF-MONITORING BEHAVIOR OF CARBON FIBER
REINFORCED POLYMER CONCRETE
4.1 Introduction
As a structural material, carbon fiber reinforced polymer concrete has received
much attention in terms of mechanical properties, but relatively little attention in terms of
electrical conduction properties. Nevertheless, the electrical properties are relevant to the
use of the structural materials for non-structural purposes such as sensing stress (selfmonitoring property), electromagnetic interference shielding and electrostatic discharge.
In this chapter, the electrical properties of interest for civil engineering applications were
investigated. The properties are compressive and tensile piezoresistive coefficients and
gage factors.

4.2 Specimen Design and Instrumentation


The same types of specimens (cylindrical and dumb-bell shaped) that were used to
measure the structural properties of CFRPC had been used to measure the electrical
properties with resistance measurements. Initially, 2 mm diameter brass electrodes were
used to measure the resistance of CFRPC material between two points. Electrodes were
embedded into the specimens during preparation in the fresh mix (Fig.4.1(a)). But stiff
electrodes prevented proper compaction of the PC material during preparation around the
electrodes, which left air voids at the contact between electrodes and the specimen.
Because of the presence of the air voids resistance of two identical specimens under same
condition was not reproducible, especially when resistance was measured under

59

compressive or tensile stress. To overcome this problem of poor reproducibility, brass


electrodes were replaced by much less stiffer 3-conductor cables. These 3-conductor
cables were embedded into the specimen during preparation in the same way as the brass
electrodes (Refer Fig. 4.1 (b)). But due to lesser stiffness of these cables, formation of air
voids at the contact between the cable and specimen were minimized through proper
compaction during preparation. This ensures better reproducibility of resistance
measurements. All the resistance data that have been presented in this chapter were
measured with 3-conductor cables.

Resin content of CFRPC specimens chosen for this study was 20% by weight.
Carbon fiber content was varied up to 6% by weight in this study. The variation of
volume resistivity of CFRPC with fiber weight fraction was as shown in Fig. 4.2. Hence
insulator material became a conductor around 3% fiber content. Also the resistivity of
CFRPC was almost indifferent to fiber weight fraction above 6% fiber loading. Hence in
this investigation the carbon fiber loading of CFRPC system had been limited to 3% and
6% by weight.

4.3 Testing Program


Cylindrical specimens were tested under uniaxial compression and dumb-bell
shaped specimens were tested under uniaxial tension to obtain the compressive and
tensile piezoresistive properties respectively. The typical compression and tension test set
up are shown in Fig. 4.3 and Fig. 4.4 respectively. The strains were measured along

60

(a)

Brass electrodes for


longitudinal resistivity
measurement

Wires for longitudinal


resistivity measurement

(b)

Wires for transverse


resistivity measurement

Strain gage for transverse


strain measurement
Strain gage for longitudinal
strain measurement

Figure 4.1 Instrumentation of CFRPC specimen for compressive electrical


property measurement (a) with brass electrodes and (b) with 3-conductor cables
61

Volume Resistivity (Ohm-cm)

7000

Coefficient of
correlation = 0.95

6000
5000
4000
3000
2000
1000
0
0

Weight Fraction of Fiber (%)

Figure 4.2 Variation of volume resistivity of CFRPC system with fiber weight
fraction

the stress axis by using commercially available 12 mm strain gages having least count of
10-6 strain/strain. Strain gages were attached to the specimens directly. During testing,
DC electrical resistance measurements were made both along the stress axis between two
known points, as shown in Fig. 4.2, using two wires, embedded in to the specimen during
preparation with the help of a Hewlett Packard 34420A Nano Volt / Micro Ohm Meter,
having a least count of 1 . Although the spacing between contacts increased on tensile
deformation and decreased on compressive deformation, the change was so small that the
measured resistance remained essentially proportional to the volume resistivity. Testing
was performed either in one cycle up to the failure stress or in multiple cycles on loading
up to a fraction (~1/12th under compression and ~1/5th under tension) of the failure stress.
At least two specimens were loaded up to failure in one cycle and at least

62

(a)

Compression testing
machine

Strain gage read-out

Specimen, Refer Fig. 4.3 (b)


Resistance meter

(b)

Insulator

Wires for longitudinal


resistance measurement

Wire to strain
gage read-out

Figure 4.3 Uniaxial compression test (a) the complete test set up and
(b) the specimen

63

(a)
Tension testing
machine
Specimen, Refer Fig. 4.4 (b)

Resistance meter
Strain gage read-out

(b)

Wires for transverse


resistance measurement

Wire to strain
gage read-out

Wires for longitudinal


resistance measurement

Figure 4.4 Uniaxial tension test (a) the complete test set up and (b) the specimen

64

one was tested in minimum of three cycles on loading up to fraction of failure strength
before loading to failure. Plain Papers were used to insulate the specimens from the
testing machine as shown in Fig.4.3 (b).

4.4 Test Results and Discussions


The typical experimental observation of the relative change in resistance of carbon
fiber reinforced polymer concrete dumb-bell shaped specimen as a function of externally
applied stress is shown in Fig.4.5 for fiber weight fraction of 3%. Three successive
tensile cycles are shown in Fig.4.5. It was observed that the first cycle did not coincide
with the others, giving evidence of hysterisis and resistance increased at zero-load at the
end of each cycle. The resistance at unstressed state increased by 0.1% after the first
loading-unloading cycle. This increase in resistance at unstressed state (R0) reduced
with increase in loading-unloading cycles as shown in Fig.4.6. Increasing the number of
successive cycles on a given specimen decreased the hysterisis and resistance change at
zero stress condition. Hence the reproducibility of the measurements was improved with
higher cycles.

But when the specimen was initially subjected to a much higher stress in the first
cycle than the ones used during following cycles, the reproducibility was obtained much
faster. This is demonstrated by the results shown in Fig.4.7 obtained on a tensile
specimen, with the same composition with the specimen as in Fig.4.5, but had been

65

2.5

1.5
Stress (MPa)

1
Cycle 1
Cycle 2
Cycle 3

0.5

0
0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

Change in Longitudinal Resistance (%)

Figure 4.5 Change in resistance of 3% CFRPC system under cyclic tensile stress
with all the cycles subjected to same stress level
3.98
Unstressed Resistance (k-Ohm)

Coefficient of correlation = 0.99


3.978

3.976

3.974

3.972

3.97
0

Cycle (No.)

Figure 4.6 Increase in unstressed resistance of 3% CFRPC system with cycles


under cyclic tensile stress with all the cycles subjected to 25% of failure strength

66

Tensile Stress (11) (MPa)

4
3
2

2
1

Cycle 1
Cycle 2
Cycle 3

0
0

0.1

0.2

0.3

0.4

0.5

0.6

Change in Longitudinal Resistance (R/R0)1 (%)

Figure 4.7 Change in resistance of 3% CFRPC system under cyclic tensile stress
with first cycle subjected to much higher stress than the following cycles
subjected to a tensile stress of 5 MPa on the first cycle and between 0 to 2.5 MPa during
next two cycles. It was observed that resistance returned to its unstressed value from 2nd
cycle onwards and there was good reproducibility between subsequent cycles. Hence all
samples were subjected to a higher stress at the first run. All the results had been obtained
following such a procedure.

Table 4.1 and Table 4.2 summarize the electrical resistivity measurements from
compressive and tensile tests. The mechanical testing results include failure strength,
strain at failure and modulus. The electrical measurement results include change in
resistivity at failure (Chung, 1996), piezoresistivity coefficient (Amin, 2001) and gage
factor (Amin, 2001). The electrical measurement gave, as raw result, the resistance (R)

67

between two known points on the specimen were measured during the testing. The
resistance was then converted to resistivity, by the equation

=R

A
.(4.1)
l

where A is the cross-sectional area and l is the distance between the electrodes. Since
resistance in the material is directional (anisotropic), based on the random distribution of
the fibers, the resistivity will be represented as a vector. Hence the piezoresistivity coefficient (ijk) was defined as follows:


= ijk jk ..(4.2)

0 i

where it relates the specific change in electrical resistivity (/0)i to the change in stress
tensor (jk). The piezoresistivity tensor (ijk) had the dimension of reciprocal stress
(m2/N). Eqn.4.2 was rewritten in terms of conjugate strain (mn) as follows [assuming the
material is incrementally elastic [ ( jk = C jkmn mn )]:


= ijk C jkmn mn = M ijk jk ..(4.3)

0 i

68

where Cjkmn was the elasticity matrix and the tensor Mijk was the elasto-resistance tensor
known as the gage factor. Gage factor signifies the sensitivity of change in resistivity
measurement to strain measurement. Both piezoresistivity coefficient and gage factor are
used in this study to quantify the sensitivity of CFRPC material to stress and strain
respectively.

11

Under uniaxial stress condition 0


0

0 0

0 0 , Eqn.4.2 simplified to:


0 0

= 111 11 (4.2a)
0 1

11
Now for uniaxial stress condition the strain tensor reduced to 0
0

0
m 22
0

0
0 and
m 33

assuming 22 = 33 = c 11 (c is the Poissons ratio of the composite) and hence


Eqn.4.3 reduced to:


= M 111 2 c M 222 11 (4.3a)

0 1

69

4.4.1 Compressive Behavior (Uniaxial)


Under uniaxial compression, the percentage change in resistivity was initially
negative (resistivity decreased as deformation took place), but after a certain threshold
value the percentage change in resistivity became positive and remained positive up to
failure (resistivity increased with deformation). The change in resistivity behavior of
CFRPC with uniaxial compressive stress is characterized with threshold change in
resistivity, initial instantaneous piezoresistivity coefficient, initial instantaneous gage
factor,

post-threshold

instantaneous

piezoresistivity

coefficient,

post-threshold

instantaneous gage factor and change in resisitivity at failure as shown below in Fig.4.8.

Stress/Strain

Failure stress/strain

Threshold change in
resistivity

Post-threshold instantaneous
piezoresistivity coefficient/gage factor
Change in resistivity at
[111th/(M111-2cM222)th]
failure
Initial instantaneous piezoresistivity
coefficient/gage factor
[1110/(M111-2cM222)0]

Change in Longitudinal Resistivity (%)

Figure 4.8 Typical stress/strain versus change in resistivity relationship of CFRPC


in uniaxial compression

70

During uniaxial compressive testing the electrical resistance along the stress axis
was conducted on 3% and 6% CFRPC systems. The results for 3% CFRPC system are
plotted in Figs. 4.9 through 4.11 and for 6% CFRPC system the results are shown in Figs.
4.12 through 4.14. Under cyclic compressive loading both the strain and change in
resistivity essentially returned to zero for both 3% and 6% systems, indicating elastic
deformation. The slopes of stress (11)-change in resistivity (/0)1 relationship were
found to be different in 3% and 6% CFRPC systems. Though the threshold change in
resistivity and initial piezoresistivity coefficient (1110) was higher for 6% CFRPC
system, the post-threshold piezoresistivity coefficient was higher for 3% CFRPC system.
The initial instantaneous piezoresistivity coefficients (1110) and initial instantaneous
post-threshold piezoresistivity coefficients (111th) obtained for 3% CFRPC system were
-700 X 10-12 m2/N and 650 X 10-12 m2/N respectively and for 6% CFRPC system were
-1000 X 10-12 m2/N and 300 X 10-12 m2/N respectively. Though for both 3% and 6%
CFRPC systems the pre-threshold piezoresistivity coefficient decreased non-linearly with
increase in stress up to the threshold change in resistivity and the post-threshold
piezoresistivity co-efficient increased non-linearly with increase in stress up to failure,
the rate of change of piezoresistive coefficient with stress was found to be higher for the
3% CFRPC system. The variations of instantaneous piezoresistivity coefficient with
stress are shown in Fig.4.21. Also it had been observed that the change in piezoresistivity
coefficient (111th) with stress for both 3% and 6% CFRPC system was very rapid
beyond 70% of the failure stress of the respective system. Gage factors (M111-2cM222)
were also calculated for both the system from the strain (11)-change in resistivity
(/0)1 relationships (Figs.4.11 and 4.13). The initial instantaneous gage factor [(M111-

71

2cM222)0] and instantaneous post-threshold gage factor [(M111-2cM222)th] obtained for


3% CFRPC system were -8 and 7.2 respectively and for 6% CFRPC system were 10.2
and 4.4 respectively. Increasing the fiber content from 3% to 6% reduced the initial
instantaneous gage factor [(M111-2cM222)0] and instantaneous post-threshold gage factor
[(M111-2cM222)th]. The stress dependences of gage factors were similar to those of
piezoresistivity coefficients but much more linear as shown in Fig.4.22. Gage factors also
showed the higher pre-threshold sensitivity of 6% CFRPC system over 3% CFRPC
system. As the stress level is increased beyond 5 MPa, 3% CFRPC system becomes more
sensitive to the stress (greater gage factor, Fig.4.22 (b)).

Table 4.1 Results of Simultaneous Mechanical and Electrical Measurements of CFRPC


under Uniaxial Compression
Mechanical

Electrical
Instantaneous
piezoresistive coefficient

Change
Strain
in
Threshold
Initial Unstressed resistivity change in
at
Strength failure modulus resistivity at failure resistivity
3%
CFRPC 60 MPa 0.95% 17.8 GPa

6300
Ohm-cm

+ 225%

350
6%
CFRPC 65 MPa 1.10% 17.2 GPa Ohm-cm + 150%
6%
6% more 3% more
6%
more 3% stiffer
Remarks
conductive sensitive
stronger
ductile

Initial
(1110)

Instantaneous
gage factor
PostPostInitial
threshold
[(M111threshold [(M111(111th) 2cM222)0] 2cM222)th]

-700 X 650 X 10-12


m2/N
- 0.06 % 10-12 m2/N

- 0.22%
6% more
sensitive

-1000 X 300 X 10-12


m2/N
10-12 m2/N
6% more
sensitive

3% more
sensitive

-8

7.2

-10.2

4.4

6% more
sensitive

3% more
sensitive

Comparing the simultaneous mechanical and electrical measurements of CFRPC


systems with varied fiber content under uniaxial compression, it can be concluded that
though the increase in fiber content did not change the mechanical properties
significantly; it had significant effect on the self-monitoring properties. Both the post-

72

70

(a) Overall Behavior


Compressive Stress, (- 11) (MPa)

60

0.7 c
3% CFRPC System

50
40
30
20
10

specimen - 1
specimen - 2

Refer Fig. 4.9 (b)

0
0

0.2

0.4
0.6
0.8
Compressive strain, (- 11) (%)

1.2

5.0

Compressive Stress, ( 11) (MPa)

4.5

(b) Initial Behavior

4.0
3.5
3.0
2.5
2.0
1.5
1.0

3% CFRPC System

0.5

Specimen 2
Cycle - 1
Cycle - 2
Cycle - 3

0.0
0

0.005

0.01

Compressive Strain, (- 11) (%)

0.015

0.02

Figure 4.9 Compressive stress-strain relationship for 3% CFRPC system in


uniaxial compression (a) overall behavior and (b) initial behavior

73

70

(a) Overall Behavior


Compressive Stress, (- 11) (MPa)

60

3% CFRPC System

50
40
30

specimen 1
specimen 2

Refer Fig. 4.10 (b)


-50

50

100

150

200

20
10
0
250

Change in Longitudinal Resistivity, (/0)1 (%)

4.5
Compressive Stress, (- 11) (MPa)

(b) Initial Behavior

4.0

3% CFRPC System

3.5
3.0
2.5
2.0

Specimen 2
cycle 1
cycle 2
cycle 3

-0.1

-0.05

0.05

0.1

1.5
1.0
0.5
0.0
0.15

Change in Longitudinal Resistivity, /0)1 (%)

Figure 4.10 Compressive stress-change in resistivity relationship for 3% CFRPC


system in uniaxial compression (a) overall behavior and (b) initial behavior

74

1.2

Compressive Strain, (- 11) (%)

(a) Overall Behavior


1

0.8

3% CFRPC System

0.6

0.4

0.2

specimen 1
specimen 2

Refer Fig. 4.11 (b)


0
-50

50

100

150

200

250

Change in Longitudinal Resistivity, (/0)1 (%)

0.02

Compressive Strain, (- 11) (%)

(b) Initial Behavior

0.018

3% CFRPC System

0.016
0.014
0.012
0.01

0.008

Specimen 2
cycle 1
cycle 2
cycle 3

0.006
0.004
0.002
0

-0.1

-0.05
0
0.05
0.1
Change in Longitudinal Resistivity, (/0)1(%)

0.15

Figure 4.11 Compressive strain-change in resistivity relationship for 3% CFRPC


system in uniaxial compression (a) overall behavior and (b) initial behavior

75

Compressive Stress, ( 11) (MPa)

80

(a) Overall Behavior

70

0.7 c

60

6% CFRPC System

50
40
30
20
10

specimen - 1
specimen - 2

Refer Fig. 4.12 (b)

0
0

0.2

0.4

0.6

0.8

1.2

1.4

Compressive Strain, (- 11) (%)

5.0

(b) Initial Behavior

Compressive Stress, (- 11) (MPa)

4.5
4.0
3.5
3.0
2.5
2.0
1.5

Specimen 2

1.0

Cycle - 1
Cycle - 2
Cycle - 3

6% CFRPC System

0.5
0.0
0

0.005

0.01

0.015

0.02

0.025

0.03

0.035

0.04

Compressive Strain, (- 11) (%)

Figure 4.12 Compressive stress-strain Relationship for 6% CFRPC system in


uniaxial compression (a) overall behavior and (b) initial behavior

76

80

(a) Overall Behavior


Compressive Stress, (- 11) (MPa)

70
60

6% CFRPC System

50
40
30
20

specimen - 1
specimen - 2

Refer Fig. 4.13 (b)


-50

0
50
100
150
Change in Longitudinal Resistivity, (/0)1 (%)

10
0
200

5.0

Compressive Stress, (- 11) (MPa)

(b) Initial Behavior

4.5
4.0

6% CFRPC System

3.5
3.0
2.5
2.0

Specimen 2
cycle 1
cycle 2
cycle 3

1.5
1.0
0.5
0.0

-0.25

-0.2

-0.15

-0.1

-0.05

Change in Longitudinal Resistivity, (/0)1 (%)

Figure 4.13 Compressive stress-change in resistivity relationship for 6% CFRPC


system in uniaxial compression (a) overall behavior and (b) initial behavior

77

1.4

(a) Overall Behavior


Compressive Strain, (- 11) (%)

1.2
1

6% CFRPC System
0.8
0.6
0.4

specimen - 1
specimen - 2

Refer Fig. 4.14 (b)


-50

50

100

150

0.2

0
200

Change in Longitudinal Resistivity, (/0)1 (%)

0.04
Compressive Strain, (- 11) (%)

(b) Initial Behavior

0.035

6% CFRPC System

0.03
0.025
0.02
0.015

Specimen 2

0.01

cycle 1
cycle 2
cycle 3

0.005
0

-0.25

-0.2

-0.15

-0.1

-0.05

Change in Longitudinal Resistivity, (/0)1 (%)

Figure 4.14 Compressive strain-change in resistivity relationship for 6% CFRPC


system in uniaxial compression (a) overall behavior and (b) initial behavior

78

-threshold stress and strain sensing properties as well as the change in resistivity at failure
improved with decreased fiber content, though the initial (up to 5 MPa stress) sensitivity
(piezoresistivity coefficient and gage factor) of 6% CFRPC system was 30% more than
that of 3% CFRPC system (Figs. 4.21 (b) and 4.22 (b)).

4.4.2 Tensile Behavior (Uniaxial)


Under uniaxial tension, the percentage change in resistivity was always positive
(resistivity increased with deformation). Tensile strain and electrical resistance were
measured along the stress axis on 3% and 6% CFRPC systems. The variations of
electrical resistance with stress and strain for 3% CFRPC system are shown in Figs.4.15
through 4.17 and those for 6% CFRPC system are shown in Figs.4.18 through 4.20.
Under cyclic uniaxial tensile loading (up to 20% of failure stress) both the strain () and
change in resistivity (/0)1 essentially returned to zero for both 3% and 6% CFRPC
systems, indicating elastic deformation. Under uniaxial tension also the slopes of the
stress (11)-change in resistivity (/0)1 relationships were different for 3% CFRPC and
6% CFRPC systems, with the piezoresistivity coefficient (111) of 3% CFRPC system
was higher than that for 6% CFRPC system. The initial instantaneous piezoresistivity
coefficients (1110) obtained for 3% CFRPC system and 6% CFRPC system were 800 X
10-12 m2/N and 600 X 10-12 m2/N respectively as compared to 66 X 10-12 m2/N for p-Si
semiconductor (Amin, 2001) and 45 X 10-12 m2/N for ESL D3414 (Ruthenium-based
MIM thick-film resistors; Amin, 2001). For both 3% and 6% CFRPC systems the
piezoresistivity coefficient (111) increased non-linearly with increase in stress up to
failure, with the rate of change of piezoresistivity coefficient (d111) with stress (d11)
79

was higher for 3% CFRPC system as compared to 6% CFRPC system as shown in


Fig.4.21. The rates of change in piezoresistivity coefficients (d11/d11) were very rapid
beyond 70% of failure stress of the respective CFRPC system. The initial instantaneous
gage factors [(M111-2cM222)0] obtained for 3% CFRPC and 6% CFRPC systems were
9.2 and 7.2 respectively. The secant gage factors [(M111-2cM222)f] obtained were 75.5
and 29.1 for 3% and 6% CFRPC systems respectively, as compared to 89 for carbon
fiber-Silica fume-Cement paste (Chung, 2000). Gage factors for both the system
increased with stress with the rates of change of gage factors with stress were higher for
3% CFRPC system as shown in Fig.4.22.

Comparing the simultaneous mechanical and electrical measurements of CFRPC


systems with varied fiber content under uniaxial tension, it can be concluded that though
the increase in fiber content improved the mechanical properties it deteriorated the selfmonitoring properties. Both the piezoresistivity coefficient and gage factor as well as the
change in resistivity at failure reduced with increased fiber content.

Figs.4.21 and 4.22 show the variations of piezoresistivity coefficients and gage
factors respectively of CFRPC system with varied fiber content under uniaxial
compression and uniaxial tension. Test results showed that the instantaneous values of
piezoresistivity coefficients and gage factors as well as their rate of change with stress of
respective CFRPC system were higher under uniaxial tension than under uniaxial
compression.

80

Table 4.2 Results of Simultaneous Mechanical and Electrical Measurements of CFRPC


under Uniaxial Tension
Mechanical

Strength
3% CFRPC
6% CFRPC

9 MPa
12 MPa

Electrical

Change
in
Strain at Initial Unstressed resistivity
failure modulus resistivity at failure
14.1
6300
0.09%
GPa
Ohm-cm + 6.80%
0.12%

p-Sia

13.9
GPa

330
Ohm-cm + 3.50%
7.8
Ohm-cm

ESL D3414a
Carbon fiber25300
Silica fume1.88 MPa
+ 5.10%
Ohm-cm
Cement Pasteb
6%
6% more
6% more 3% more
Remarks
stronger
ductile
conductive sensitive
a
Amin, 2001; b Wen and Chung, 2000; Chen and Chung, 1996

Secant
Initial
Initial
gage
instantaneous instantaneous factor at
piezoresistive gage factor failure
[(M111co-efficient
[(M111(111)0
2cM222)0] 2cM222)f]
800 X 10-12
m2/N
9.2
75.5
600 X 10-12
m2/N
66 X 10-12
m2/N
45 X 10-12
m2/N

7.2

29.1

89
3% more
sensitive

3% more
sensitive

3% more
sensitive

4.5 Summary
Based on the mechanical and electrical measurements of CFRPC under uniaxial
loading the following can be summarized:
1. Caron fiber reinforced polymer concrete (CFRPC) was found to be a selfmonitoring material that can sense stress and strain both in compression and
tension. The electrical resistivity changed with application stress and strain.
2. Self-monitoring behavior of CFRPC system was observed at fiber loading of
about 3% by weight or higher. Self-monitoring properties such as piezoresistivity
coefficient and gage factor changed with fiber weight fraction. Increasing the
fiber content from 3% to 6% reduced the piezoresistivity coefficient and gage
factor parameters.

81

Tensile Stress, 11 (MPa)

10
9

(a) Overall Behavior

3% CFRPC System

0.7 t

7
6
5
4
3
2

specimen 1
specimen 2

Refer Fig. 4.15 (b)

1
0
0

0.02

0.04

0.06

0.08

0.1

0.12

Tensile Strain, 11 (%)


2.5

(b) Initial Behavior


3% CFRPC System

Tensile Stress, 11 (MPa)

1.5

Specimen 2
Cycle 1
Cycle 2
cycle 3

0.5

0
0

0.002

0.004

0.006

0.008

0.01

0.012

0.014

Tensile Strain, 11 (%)

Figure 4.15 Tensile stress-strain relationship for 3% CFRPC system in uniaxial


tension (a) overall behavior and (b) initial behavior

82

Tensile Stress, 11 (MPa)

10
9

(a) Overall Behavior

3% CFRPC System

7
6
5
4
3
2
1

specimen 1
specimen 2

Refer Fig. 4.16 (b)

0
0

Change in Longitudinal Resistivity, (/0)1 (%)

Tensile Stress, 11 (MPa)

2.5

(b) Initial Behavior


3% CFRPC System

1.5

Specimen 2
cycle 1
cycle 2
cycle 3

0.5

0
0

0.05

0.1

0.15

0.2

0.25

0.3

Change in Longitudinal Resistivity, (/0)1 (%)

Figure 4.16 Tensile stress-change in resistivity relationship for3% CFRPC system


in uniaxial tension (a) overall behavior and (b) initial behavior

83

0.12

(a) Overall Behavior


0.1

Tensile Strain, 11 (%)

3% CFRPC System

0.08

0.06

0.04

specimen 1
specimen 2

Refer Fig. 4.17 (b)

0.02

0
0

Change in Longitudinal Resistivity, (/0)1 (%)

0.014

(b) Initial Behavior

0.012

3% CFRPC System
Tensile Strain, 11 (%)

0.01
0.008
0.006

Specimen 2
cycle 1
cycle 2
cycle 3

0.004
0.002
0

0.05

0.1

0.15

0.2

0.25

0.3

Change in Longitudinal Resistivity, (/0)1 (%)

Figure 4.17 Tensile strain-change in resistivity relationship for 3% CFRPC


system in uniaxial tension (a) overall behavior and (b) initial behavior

84

14

(a) Overall Behavior

Tensile Stress, 11 (MPa)

12

0.7 t

6% CFRPC System

10
8
6
4
2

specimen 1
specimen 2

Refer Fig. 4.18 (b)

0
0

0.02

0.04

0.06

0.08

0.1

0.12

0.14

Tensile Strain, 11 (%)

2.5

(b) Initial Behavior


6% CFRPC System

Tensile Stress, 11 (MPa)

1.5

Specimen 2
cycle 1
cycle 2
cycle 3

0.5

0
0

0.002

0.004

0.006

0.008

0.01

0.012

0.014

0.016

Tensile Strain, 11 (%)

Figure 4.18 Tensile stress-strain relationship for 6% CFRPC system in uniaxial


tension (a) overall behavior and (b) initial behavior

85

14

(a) Overall Behavior


6% CFRPC System

Tensile Stress, 11 (MPa)

12
10
8
6
4
2

specimen 1
specimen 2

Refer Fig. 4.19 (b)

0
0

0.5

1.5

2.5

3.5

Change in Longitudinal Resistivity, (/0)1 (%)

2.5

(b) Initial Behavior


6% CFRPC System

Tensile Stress, 11 (MPa)

1.5

Specimen 2
cycle 1
cycle 2
cycle 3

0.5

0
0

0.02

0.04

0.06

0.08

0.1

0.12

0.14

0.16

0.18

Change in Longitudinal Resistivity, (/0)1 (%)

Figure 4.19 Tensile stress-change in resistivity relationship for 6% CFRPC


system in uniaxial tension (a) overall behavior and (b) initial behavior

86

0.14

(a) Overall Behavior


6% CFRPC System
Tensile Strain, 11 (%)

0.12
0.1
0.08
0.06
0.04

specimen 1
specimen 2

Refer Fig. 4.20 (b)

0.02
0

0.5

1.5

2.5

3.5

Change in Longitudinal Resistivity, (/0)1(%)

0.016

(b) Initial Behavior

0.014

Tensile Strain, 11 (%)

6% CFRPC System

0.012
0.01
0.008
0.006

Specimen 2
cycle 1
cycle 2
cycle 3

0.004
0.002
0

0.02

0.04

0.06

0.08

0.1

0.12

0.14

0.16

0.18

Change in Longitudinal Resistivity, (/0)1 (%)

Figure 4.20 Tensile strain-change in resistivity relationship for 6% CFRPC system


in uniaxial tension (a) overall behavior and (b) initial behavior

87

Absolute Piezoresistivity Coefficient (m2/MN)

0.035

3% CFRPC under
uniaxial compression

0.03

3% CFRPC under
uniaxial tension

0.025
0.02

6% CFRPC under
uniaxial tension

0.015
0.01
0.005

6% CFRPC under
uniaxial compression

0
0

10

20

30

40

50

60

70

80

Axial Stress (MPa)

Refer Figure 4.21(b)

Absolute Piezoresistivity Coefficient (m2/MN)

0.008
0.007

3% CFRPC under
uniaxial compression

3% CFRPC under
uniaxial tension

0.006
0.005

6% CFRPC under
uniaxial compression

6% CFRPC under
uniaxial tension

0.004
0.003
0.002
0.001
0
0

10

12

14

Axial Stress (MPa)

Figure 4.21 Variation of secant piezoresistivity coefficient of different CFRPC


systems with stress (a) overall behavior and (b) initial behavior

88

250

3% CFRPC under
uniaxial tension

200
Absolute Gage Factor

3% CFRPC under
uniaxial compression

150

6% CFRPC under
uniaxial tension

100

50

6% CFRPC under
uniaxial compression

0
0

10

20

30

Refer Figure 4.22 (b)

40

50

60

70

80

Axial Stress (MPa)

70

3% CFRPC under
uniaxial tension

Absolute Gage Factor

60

6% CFRPC under
uniaxial compression

50

6% CFRPC under
uniaxial tension

40

3% CFRPC under
uniaxial compression

30
20
10
0
0

10

12

14

Axial Stress (MPa)

Figure 4.22 Variation of secant gage factor of different CFRPC systems with stress
(a) overall behavior and (b) initial behavior

89

3. The change in resistivity response with stress was found to be different in uniaxial
compression and tension. Under compression, resistivity initially decreased with
increasing stress but after a certain threshold value, it increased with stress up to
failure. On the other hand, under tension resistivity continuously increased with
increasing tensile stress up to failure.
4. The stress and strain sensing ability of CFRPC systems under uniaxial tension
was higher than under uniaxial compression and quantified in terms of
piezoresistivity coefficients and gage factors.
5. Both pizoresistivity coefficient and gage factor as well as percentage change in
resistivity increased rapidly when stress level was higher than 70% of failure
strength of the CFRPC system which suggested fracture in the material.

90

CHAPTER 5
MODELING OF SELF-MONITORING BEHAVIOR OF
FIBER REINFORCED POLYMER CONCRETE
5.1 Introduction
Modeling of piezoresistive behavior of carbon fiber reinforced polymer concrete
has been presented in this chapter. Percolation theory has been used to predict the
unstressed resistivity of carbon fiber reinforced polymer concrete system. An electromechanical constitutive model based on the principle of continuum mechanics and
percolation theory has been proposed to model the piezoresistivity of carbon fiber
reinforced polymer concrete system under loading. The mechanical properties of CFRPC
systems that were used in this chapter were obtained in Chapter 3. The mechanical
properties of carbon fibers were from manufacturers specification. The piezoresistive
model developed in this chapter is used in Chapter 6 to predict the piezoresistive behavior
of CFRPC under bending.

5.2 Analytical Formulation


At microstructural level carbon fiber reinforced polymer concrete can be idealized
as random arrangement of polymer-surrounded-carbon-fiber lattices in granular (sand)
medium. At low concentration of carbon fibers (below percolation threshold) the
conducting carbon fibers in neighboring lattices are well apart and are insulated by
surrounding polymer matrix (connectivity 0-3). At 0-3 connectivity, the composite
resistivity is limited by polymer matrix. Polymer matrix can be idealized as a perfect
insulator and hence at very low concentration of carbon fiber the composite will act as an

99

insulator. If we increase the volume fraction of carbon fiber or if we apply external stress
or due to internal stress (shrinkage or curing), the connectivity will change from 0-3 to x3, where x > 0 and x < 3. At x = 3 i.e. at 3-3 connectivity, the composite will act as a
perfect conductor and the fiber resistivity will govern the composite resistivity. For
connectivity 1-3 or 2-3 the composite resistivity of carbon fiber reinforced polymer
concrete can be best described with the help of percolation theory.

Sand
particles

Polymer-surrounded-carbon
fiber

Figure 5.1 - Idealized microstructure of carbon fiber reinforced polymer concrete

From percolation theory, the resistivity of a composite can be represented as (Carmona


et. al. 1987)

i = A fiber ( crit )t

...(5.1)

where i is the composite resistivity in the i-direction (i = 1,2,3); A is a prefactor; fiber is


the resistivity of carbon fibers; is the volume fraction of the carbon fibers; crit is the
critical volume fraction of the carbon fibers; ti is the critical conductivity exponent in the
i-direction (i = 1,2,3).

100

Taking logarithm on both sides of Eqn.5.1 and differentiating with respect to stress
tensor jk

1
1 fiber
1 ( crit )
= fiber
.(5.2)
t i ( crit )
jk
jk
jk 0 i 0

Neglecting the intrinsic piezoresistivity of the fibers, Eqn.5.2 can be written as

1
1 crit
= ti ( crit )1
+ ti ( crit )
(5.3)
jk 0 i
jk
jk

Let us concentrate on the first part of Eqn.5.3. The volume fraction of the conducting
carbon fibers () can be written as

Vf
...(5.4)
V f +V p +V s

where Vf, Vp, Vs are the volume of conducting fibers, polymer matrix and sand
respectively. Taking logarithm on both sides of Eqn.5.4 and differentiating with respect
to the stress tensor, jk

(V f + V p + V s )
1
1 V f
1
(5.5)
= f
f
jk V jk V + V p + V s
jk

101

Now, lets consider, the following term in Eqn.5.5. Rearranging, this term results in the
following relationship (when dVf 0):

1 V f
1
.(5.6)
=
f
jk
V jk
V f
f
V

V f
f
V

is the infinitesimal volumetric strain in fibers due to infinitesimal stress increase

jk and hence can be represented (by tensor notation) as ll . Hence Eqn.5.6 can be
represented as follows:

1 V f
1
1
1
1
=
=
=
=
f

( m jk + S jk ) ( m jk ) S jk
m S jk
V jk jk
+
m jk + jk
+
ll
ll
ll
ll
ll
ll
ll
..(5.7)

where m is the mean stress, jk is the kronekar delta and Sjk is the deviatoric stress
tensor. Hence Eqn.5.7 reduces to the following:

1 V f
V f jk

=
S
K f jk + jk
ll

102

..(5.8)

fiber

where Kf is the co-efficient of volume compressibility of fiber. So the value of K is +ve if


the mean tensile stress causes volume expansion and is ve if the mean compressive
stress causes volume reduction. Similarly, the second part of Eqn.5.5 reduces to the
following relationship [d(Vf+Vp+Vs)0]

(V + V + V )
1
1
=
f
p
s
S
jk
(V + V + V )
K c jk + jk
ll

(5.9)

composite

where Kc is the volume compressibility of the composite. By Combining Eqn.5.8 and

1
Eqn.5.9, the first term t i ( crit )
of Eqn.5.3 can be represented as:
jk

1
1
1
t i ( crit )
= t i ( crit )

S jk
jk
K jk +
ll

K jk + jk
ll
fiber

composite

(5.10)

For a composite the critical volume fraction (crit) in Eqn.5.3 can change due to
change in microstructure of the composite. Change in the microstructure of the composite
can be achieved by externally applied strain (or stress) as shown in Fig.5.2.

103

(a)

(b)

Figure 5.2 - Effect of strain on microstructure (a) percolating before straining


(b) not percolating after straining

From Fig.5.2, an initially percolating system can become an insulator (specific


resistivity approaching infinity) with shearing strain (or stress) because of the change in
the microstructure. Note that in Fig.5.2(a), the conducting fibers formed a connective
path across the boundary of the system [path 1 to 2] whereas in Figure 5.2(b) the
conductive passage was broken because of shearing strain. Hence the system in Fig.
5.2(b) will require more fibers to become conductive and hence the percolation threshold
of the system in Fig.5.2(b) will require more fibers (higher crit) compared to Fig.5.2(a).
The change of criit with stress tensor can be written as follows:

crit
1
1
1
=
=
=
(5.11)
jk
( m jk + S jk )
jk
jk
m S jk
m
+ jk
+
crit
crit
crit
crit crit

104

If criit in Eqn.5.3 is influenced by the deviatoric stress tensor, then Eqn.5.11 can be
simplified as follows:

crit crit
=
..(5.12)
jk
S jk

If the change in crit with respect to deviatoric stress tensor is proportional to the
deviatoric stress tensor, Eqn.5.12 reduces to

crit crit
=
= ZS jk .(5.13)
jk
S jk

where, Z is the constant of proportionality. Combining Eqn.5.13 and Eqn.5.10, Eqn.5.3


can be written as

1

jk 0 i

1
1
= t i ( crit )
S
K f jk + jk
ll

S
K c jk + jk

ll
fiber

+ t i ( crit ) ZS jk

composite

..(5.14)

Therefore, the change in resistivity due to the applied stress can be represented in terms
of stress as follows:

105



0 i

1
= t i ( crit )
S

K f jk + jk
ll

K c jk + jk

ll
fiber

+ t i ( crit ) ZS jk d jk

composite

(5.15)

1
1
= t i ( crit ) c

K jk + B cjk

f
f

composite K jk + B jk

+ t i ( crit )1 ZS jk d jk

fiber

(5.16)

where, B jk is the shear parameter, which quantifies volume change of the composite/fiber
under deviatoric stress. For an isotropic homogeneous and linear elastic material, the coefficient of volume compressibility (K) can be written in terms of Youngs modulus (E)
and Poissons ratio () as

K=

E
(5.17)
3(1 2 )

Therefore, the change in resistivity of linear elastic, isotropic and homogeneous


composite under applied stress can be written in most general form as follows:

0 i

1
1

= t i ( crit )

+ ZS jk
c
f

E
E

jk + B cjk
jk + B jkf
c
f

3(1 2 )

3(1 2 )

jk

.(5.18)

106

The above relationship is identical to the assumed relationship Eqn.4.2 of chapter 4,


which relates change in resistivity to the stress tensor. Hence the piezoresistivity
coefficient (ijk) used in chapter 4 can be represented as follows:

ijk

1
1

= t i ( crit )

+ ZS jk
c
f

E
E


jk + B cjk
jk + B jkf
c
f

3(1 2 )
3(1 2 )

(5.18a)

Eqn.5.18 relates the piezoresistivity of carbon fiber reinforced polymer concrete to


the fiber properties (Ef, f and Bf) and composite properties (Ec, c, Bc and Z), fiber
volume fraction () and critical volume fraction (crit).

Notice that if we assume crit is independent of stress (i.e. Z = 0) and represent

1
1

= a jk ,

f
c

E
E

+ B cjk
jk + B jkf
f
3(1 2 c ) jk


3(1 2 )

Eqn.5.18

reduces

to

the

following relationship:

a jk t i

(
)

crit
i

107

jk

.(5.19)

which is in agreement with the relationship developed by Carcia et al. (1983, Eqn.2(b))
and Caroma et al. (1987, Eqn.14).

Eqn.5.18 (the most general equation of piezoresistivity) can be solved


incrementally, updating the values of and crit for each small stress increment. The
relation between volume fraction of fiber, and the stress tensor, ij (from Eqn.5.10) can
be written as

1
1

=
S
jk
K f jk + jk
ll

K c jk + jk
ll
fiber

(5.20)

composite

or


S jk
jk
K f jk +
ll

K c jk + jk
ll
fiber

=0
.(5.21)

composite

which is of the form

y ' + ay = 0 ..(5.22)

The solution of Eqn.35 can be written as

108

= Ce

S
f
K jk + jk

ll

S
c

K jk + jk

ll
fiber

jk

composite

.(5.23)

Applying the initial condition: At jk = 0, = 0 therefore C = 0


Therefore Eqn.5.23 can be written as

= 0 e

S
K f jk + jk

ll

K c jk + jk

ll
fiber

jk

composite

(5.24)

Now, the relation between the critical volume fraction of fiber, crit and the stress tensor,
jk can be written as (from Eqn.5.13)

crit crit
=
= ZS jk (5.25)
jk
S jk

Therefore,

crit

= Z S jk dS jk ....(5.26)

Therefore,

109

crit = Z

S jk S jk
2

+ C ..(5.27)

Applying the initial condition: When S jk = 0, crit = ( crit ) 0


and hence Eqn.5.27 can be written as

crit = ( crit ) 0 + Z

S jk S jk
2

..(5.28)

5.3 Resistivity at Unstressed State


The resistances of carbon fiber reinforced polymer concrete specimens (cylinder,
dumb-bell, beam and sandwich pipe) were measured between two points of known
distance at room temperature and at no-load condition. From the measured resistance,
resistivities were calculated and plotted as a function of carbon fiber volume
concentration and shown in Fig.5.3. The expected insulator-to-conductor transition
(percolation phenomena) was clearly evidenced in the experimental observation.
Percolation equation (Eqn.5.1) was curve-fitted with the experimental data by leastsquare method and the values for the critical volume fraction of fiber (crit), the
conductivity exponent (t1) and the pre-factor (A) were calculated for the composite
material. The values of critical volume fraction of the carbon fibers (crit), conductivity
exponent (t1) and pre-factor (A) for the CFRPC systems were determined to be 0.028,
1.7315 and 427.39 respectively. The value of the conductivity exponent was found to be
in good agreement with the universal value predicted by the percolation theory for 3-D

110

conductor-insulator multiphase solids (Strumpler et. al., 1999). The relationship with
these values as parameters (crit, t1 and A) is compared to the experimental results in
Fig.5.3 and a good agreement (Coefficient of correlation = 0.95) with the experimental
data was observed.

Longitudinal Resistivity (Ohm-cm)

7000
6000
5000
4000

-t1
1 = Ac(-crit)

Model

A = 427.39
c= 1.73X10-3 Ohm-cm
crit= 0.028
t1 = 1.7315

Coefficient of
correlation = 0.95

3000
2000

Percolation Threshold

1000
0
0.000

0.010

0.020

0.030

0.040

0.050

0.060

0.070

Volume Fraction of Fiber

Figure 5.3 Prediction of longitudinal resistivity of CFRPC systems with varying


fiber content

5.4 Piezoresistivity under Uni-axial Compression


The change in resistance between two points separated by a known distance along
longitudinal axis of carbon fiber reinforced polymer concrete cylindrical specimens of
varying fiber content were measured during compression loading. It was observed that
the change in resistivity depends on fiber content as well as on the magnitude of stress.
Irrespective of fiber content, the stress-change in resistivity behavior showed a similar
111

trend resistivity decreased initially with compressive stress up to a threshold value, a


function of fiber content, and then increased with compressive stress up to failure.
Eq.5.18 has been used to predict the piezoresitivity under uni-axial compression. The
stress tensor in uni-axial compression can be written in terms of mean stress and
deviatoric stress as

11
0

11

0 0 3
0 0 = 0

0 0
0

11
3
0

2
0 11
3

0 + 0

11
0
3

11
3
0

0 (5.29)

11

Therefore, considering the mechanical properties of the material as homogeneous


and isotropic, from Eqn.5.18, the fractional change in resistivity along the direction of
stress due to uni-axial compression (Using the sign of K as ve) can be written as


1
1
2 Z 11
1

= t1 ( crit )
11

+
3

E cFRPC
E cf
0 1
FRPC
f


+B
+B

f

3(1 2 cFRPC )
3(1 2 c )

..(5.30)

From Eqn.5.30 it can inferred that the change in resistivity of carbon fiber
reinforced polymer concrete subjected to uniaxial compression, in the direction of stress
depends on the volume fraction of the carbon fibers (), the difference between the

112

volume fraction of carbon fibers and the crtical volume fraction of the carbon fibers (crit), the material parameters of CFRPC system in compression (EcFRPC, cFRPC, BFRPC),
the material parameters of the carbon fibers (Ecf, cf, Bf) and the rate of change in
microstructure (Z) of the CFRPC system.

Using the value of Youngs modulus and Poissons ratio of CFRPC system
(obtained in Chapter 3), Youngs modulus and Poissons ratio of the carbon fibers
(obtained from manufacturer specifications) and assuming the shear parameters of the
CFRPC system and the carbon fibers (BFRPC, Bf and Z), the differential equation of
piezoresistivity under uniaxial compression (Eqn.5.30) was solved incrementally,
updating the value of (Eqn.5.24) and crit (Eqn.5.28) for each small stress increment
(11). Total change in resistivity under applied uniaxial compressive stress was
calculated as the cumulative change in resistivity [(/0)1] of each small stress
increment (11) and was fitted with the experimental data by multiple regression
analysis. Under uniaxial compression, Eqn.5.24 reduced to

= 0 e

K cFRPC + B FRPC

1

FRPPC

FRPC

1
B

FRPC

1
K cFRPC + B FRPC
1
B FRPC


1

f
f
B
Kc + B

1
1

B FRPC
Bf

1
1

FRPC
FRPC
Kc
+B
Bf

1

FRPC

1
f

B
1

K cf + B f
1
Bf

B
11 0 0

1

0 0 0
Bf
0 0 0

f
f
Kc + B
1

(5.31)

113

= 0 e

1

FRPC
+ B FRPC
Kc

K f +B f
c

11

......(5.32)

and Eqn.5.28 reduced to

crit = ( crit )0

2 11
3
1
+ Z 0
2

= ( crit )0 +

11
3
0

2
0 11
3

0 0

11

0
3

11
3
0

0 (5.33)

11

1
Z 112 .(5.34)
3

The values of BFRPC and Z, which correspond to the best-fit curve, were assumed to
be the shear parameters for the CFRPC system and the value of Bf as the shear parameter
for the carbon fibers. The model predictions of the change in resistivity have been plotted
with experimental observations in Fig.5.4 and Fig.5.5 for 3% and 6% CFRPC systems
respectively. The values of shear parameters (BFRPC and Z) of 3% CFRPC system were
found to be 90 GPa and 3.2 X 10-6 m4/N2 respectively and for 6% CFRPC system, the
respective values obtained were 2.5 GPa and 6.4 X 10-6 m4/N2. The value of the shear
parameter (Bf) of the fiber was found to be 1030 GPa.

114

70

(a) Overall Behavior


Compressive Stress, (- 11) (MPa)

60
50
40

3% CFRPC System

30
20

specimen 1 (no cycle)


specimen 2 (cycle 3)
Model

Refer Fig. 5.4 (b)


-50

50

100

150

200

10
0
250

Change in Longitudinal Resistivity, (/0)1 (%)

Compressive Stress, (- 11) (MPa)

5.0

-0.12

(b) Initial Behavior

4.5
4.0
3.5

3% CFRPC System

3.0
2.5
2.0

-0.1

-0.08

-0.06

-0.04

-0.02

Specimen 2

1.5

cycle 2
cycle 3
Model

1.0

0.02

0.04

0.5
0.0
0.06

Change in Longitudinal Resistivity, (/0)1 (%)

Figure 5.4 Model prediction of piezoresistivity under uni-axial compression of 3%


CFRPC system (a) overall behavior and (b) initial behavior

115

Compressive Stress, (- 11) (MPa)

80
70

(a) Overall Behavior

60
50

6% CFRPC System

40
30
20
10

specimen 1 (no cycle)


specimen 2 (cycle 3)
Model

Refer Fig. 5.5 (b)

0
-50

50

100

150

200

Change in Longitudinal Resistivity, (/0)1 (%)

Compressive Stress, (- 11) (MPa)

(b) Initial Behavior

7
6
5

6% CFRPC System

4
3

Specimen 2

cycle 2
cycle 3
Model
-0.25

-0.20

-0.15

-0.10

-0.05

0
0.00

Change in Longitudinal Resistivity, (/0)1 (%)

Figure 5.5 Model prediction of piezoresistivity under uniaxial compression of 6%


CFRPC system (a) overall behavior and (b) initial behavior

116

5.5 Piezoresistivity under Uni-axial Tension


The change in resistance between two points separated by a known distance along
longitudinal axis of carbon fiber reinforced polymer concrete specimens of varying fiber
content were measured during tensile loading. It was observed from the experimental
observation, as presented in Chapter 4, that change in resistivity depended on the fiber
content as well as on the magnitude of stress. For both 3% CFRPC system and 6%
CFRPC system, the change in resistivity increased non-linearly with tensile stress up to
failure.

Eqn.5.18 has been used to predict the piezoresitivity under uni-axial tension. The
stress tensor in uniaxial tension can be written in terms of mean stress and deviatoric
stress as

11
0

11

0 0 3
0 0 = 0

0 0
0

11
3
0

2
0 11
3

0 + 0

11
0
3

11
3
0

0 (5.35)

11

Therefore, considering the mechanical properties of the material as homogeneous


and isotropic, from Eqn.5.18, the fractional change in resistivity along the direction of
stress due to uniaxial tension (Using the sign of K as +ve) can be written as

117


1
1
2 Z 11
1

= t1 ( crit )

+
11
FRPC
f
3

E
E
0 1
FRPC
f
t
t


+B
+ B

f

3(1 2 tFRPC )

3
(
1
2
)

..(5.36)

From Eqn.5.36 it can inferred that the piezoresistivity of carbon fiber reinforced
polymer concrete under uniaxial tension depends on the volume fraction of the carbon
fibers (), the difference between the volume fraction of the carbon fibers and the crtical
volume fraction of the carbon fibers (-crit), the material parameters of the CFRPC
system in tension (EtFRPC, tFRPC, BFRPC), the material parameters of the carbon fibers (Etf,
tf, Bf) and the rate of change in microstructure (Z) of CFRPC system.

Using the values of Youngs modulus and Poissons ratio of CFRPC system in
tension (EtFRPC and tFRPC - obtained from Chapter 3), Youngs modulus and Poissons
ratio of the carbon fibers (Etf and tf - obtained from manufacturers specification) and the
shear parameters of the CFRPC system and the carbon fibers (BFRPC, Bf and Z - obtained
from curve-fitting of compressive piezoresistive data), the differential equation of
piezoresistivity under uniaxial tension (Eqn.5.36) was solved incrementally, updating the
value of (Eqn.5.24) and crit (Eqn.5.28) for each small stress increment. Total change
in resistivity was calculated as the cumulative change in resistivity [(/0)1] of each
small increment (11). The model predictions of change in resistivity are compared with
the experimental observations in Fig.5.6 and Fig.5.7 for 3% and 6% CFRPC systems
respectively. The model predicted the change in resistivity reasonably up to 70% of
118

failure stress but under predicted the change in resistivity near failure. This might be
because of formation of cracks at higher stress levels.

Under uniaxial tension, Eqn.5.24 reduced to

= 0 e

FRPC
Kt
+ B PC

1

B FRPC

FRPC

1
B

FRPC

1
KtFRPC + B PC
1
B FRPC


1
f
f
B
Kt + B

1
1

f
B
B FRPC

1
1

FRPC
FRPC
f
Kt
+B
B
1

FRPC

B
1

Ktf + B f
1
Bf

B
11 0 0

1

0 0 0
f

B
0 0 0

Kt f + B f
1

.(5.37)

= 0 e

1
FRPC FRPC
+B
Kt

K f +B f
t

11

........(5.38)

and Eqn.5.28 reduced to

crit = ( crit )0

2 11
3
1
+ Z 0
2

11
3
0

2
0 11
3

0 0

11

0
3

11
3
0

0 (5.39)

11

1
= ( crit )0 + Z 112 ..(5.40)
3

119

12

(a) Overall Behavior


Tensile Stress, (11) (MPa)

10

3% CFRPC System

specimen 1 (no cycle)


specimen 2 (cycle 3)

Refer Fig. 5.6 (b)

Model

0
0

2
3
4
5
6
Change in Longitudinal Resistivity, (/0)1 (%)

(b) Initial Behavior


Tensile Stress, (11) (MPa)

2.5
2
1.5
1

Specimen 2
cycle 2
cycle 3
Model

3% CFRPC System
0.5
0
0

0.05

0.1

0.15

0.2

0.25

0.3

Change in Longitudinal Resistivity, (/0)1 (%)

Figure 5.6 Model prediction of piezoresistivity of 3% CFRPC system under uniaxial tension (a) overall behavior (b) initial behavior

120

14

(a) Overall Behavior

Tensile Stress, (11) (MPa)

12
10
8

6% CFRPC System

6
4

specimen 1 (no cycle)

specimen 2 (cycle 3)

Refer Fig. 5.7 (b)

Model

0
0

0.5

1.5

2.5

3.5

Change in Longitudinal Resistivity, (/0)1 (%)

(b) Initial Behavior


Tensile Stress, (11) (MPa)

2.5
2
1.5

Specimen 2

cycle 2
cycle 3
Model

6% CFRPC System
0.5
0
0

0.05

0.1

0.15

0.2

Change in Longitudinal Resistivity, (/0)1 (%)

Figure 5.7 Model prediction of piezoresistivity of 6% CFRPC system under


uni-axial tension (a) overall behavior and (b) initial behavior

121

5.6 Parametric Study


The primary concern to this study is to investigate the effect of individual properties
of the composite material on the tensile and compressive piezoresistive behavior of
CFRPC system to improve the sensitivity of self-monitoring capacity of the composite
system. For this purpose the effect of fiber Youngs modulus, fiber Poissons ratio and
fiber weight fraction on the percentage change in longitudinal resistivity was studied
using the model equations (Eqns.5.30 and 5.36) for both uniaxial compression and
uniaxial tension.

5.6.1 Compressive Behavior


5.6.1.1 Fiber Properties
Fig.5.8 shows the effect of Youngs modulus of carbon fibers on the compressive
piezoresistivity of CFRPC system. Both the overall behavior as well as the initial
behavior has been shown in Fig.5.8. Though the Youngs modulus of fiber (Ef) has no
significant effect on the post-threshold piezoresistivity co-efficient (111th), it has
significant effect on the initial piezoresistivity coefficient (1110) as well as on the
threshold negative change in resistivity. The initial piezoresistivity coefficient increased
by 100% and the negative threshold change in reistivity increased by 600%, when the
fiber Youngs modulus (Ef) was increased from 150 GPa to 240 GPa.

Fig.5.9 shows the effect of Poissons ratio of carbon fibers (f) on compressive
piezoresistivity of CFRPC system. Clearly, both the negative threshold change in
resistivity and the initial piezoresistivity coefficient (1110) increased with decrease in
122

Compressive Stress, (- 11) (MPa)

80
70

Ecf = 230 GPa

(a) Overall Behavior


Ecf = 240 GPa

60

Ecf= 150 GPa

50

t1 = 1.7315
= 0.064
crit = 0.028
EcFRPC = 17 GPa
cFRPC = 0.15
cf = 0.1
BFRPC = 2.5 GPa
Bf = 1030 GPa
Z = 6.4 X 10-6

40
30

6% CFRPC System

20
10

Refer Fig. 5.8 (b)

0
-10

10

20

30

40

50

60

Change in Longitudinal Resistivity ( / 0 )1(%)

7
Compressive Stress, (- 11) (MPa)

(b) Initial Behavior


6
f

Ec = 240 GPa

t1 = 1.7315
= 0.064
crit = 0.028
Ec

FRPC

Ec = 230 GPa

4
f

Ec = 150 GPa

= 17 GPa

cFRPC = 0.15
cf = 0.1

FRPC

B
= 2.5 GPa
f
B = 1030 GPa
-6
Z = 6.4 X 10

6% CFRPC System

0
-0.8

-0.7

-0.6

-0.5

-0.4

-0.3

-0.2

-0.1

Change in Longitudinal Resistivity, (/0)1 (%)

Figure 5.8 Predicted effect of Youngs modulus of fiber on compressive


piezoresistive behavior of CFRPC system (a) overall behavior and (b) initial
behavior

115

Compressive Stress, (- 11) (MPa)

80

(a) Overall Behavior

70

c = 0.1

c = 0.2

60
50
40

6% CFRPC System

30
20
10

Refer Fig. 5.9 (b)

0
-5

15

25

35

t1 = 1.7315
= 0.064
crit = 0.028
EcFRPC = 17 GPa
cFRPC = 0.15
cf = 228 GPa
BFRPC = 2.5 GPa
Bf = 1030 GPa
Z = 6.4 X 10-6
45

55

Change in Longitudinal Resistivity, (/0)1 (%)

7
Compressive Stress, (- 11) (MPa)

(b) Initial Behavior


f
c

= 0.1
f

6% CFRPC System
-0.35

-0.25

-0.15

-0.05

c = 0.2

t1 = 1.7315
= 0.064
crit = 0.028
EcFRPC = 17 GPa
cFRPC = 0.15
cf = 228 GPa
BFRPC = 2.5 GPa
Bf = 1030 GPa
Z = 6.4 X 10-6

0.05

3
2
1

0
0.15

Change in Longitudinal Resistivity, (/0)1 (%)

Figure 5.9 Predicted effect of Poissons ratio of fiber on compressive piezoresistive


behavior of CFRPC system (a) overall behavior and (b) initial behavior

116

carbon fiber Poissons ratio (f), though the effect on post-threshold piezoresistivity
(111th) coefficient was minimal.

5.6.1.2 Fiber Reinforced Polymer Concrete Properties


Fig.5.10 shows the effect of carbon fibers weight fraction () on the compressive
piezoresistivity of CFRPC system. In studying the effect of carbon fibers weight fraction
(), the mechanical properties of CFRPC system were adjusted accordingly. The
variations of mechanical properties of CFRPC system (Youngs modulus, EcFRPC and
EtFRPC; shear parameters, BFRPC and Z) were assumed as shown in Fig.5.11. The Poissons
ratio of CFRPC system (FRPC) was assumed to be independent of fiber weight fraction.
Both the initial piezoresistivity coefficient (1110) and post-threshold piezoresistivity
coefficient (111th) increased with decrease in fiber weight fraction.

Fig.5.12 shows the effect of the Poissons ratio of CFRPC system (FRPC) on its
compressive piezoresitivity behavior. Clearly, the Poissons ratio of CFRPC system
(FRPC) has significant effect on its compressive piezorestive behavior. Though the postthreshold stress-change in resistivity behavior did not change much, the initial
piezoresistivity coefficient (1110) increased by 80% and threshold negative change in
resistivity increased by 500% when the Poisons ratio of CFRPC was reduced to 0 from
0.40.

123

35
6%

5% 4%

3%

Compressive Stress, (- 11) (MPa)

30
25

(a) Overall Behavior

20
15

t1 = 1.7315
crit = 0.028
cf = 228 GPa
cf = 0.1
Bf = 1030 GPa

10
5

Refer Fig. 5.10(b)


0
-5

10

15

20

25

30

35

40

Change in Longitudinal Resistivity, ( / 0 ) 1 (%)

Compressive Stress, (- 11) (MPa)

7
3%

6
5

6%

(b) Initial Behavior

4
3

t1 = 1.7315
crit = 0.028
cf = 228 GPa
cf = 0.1
Bf = 1030 GPa

2
1
0
-0.3

-0.2

-0.1
0
0.1
0.2
0.3
Change in Longitudinal Resistivity, (/0)1 (%)

0.4

0.5

Figure 5.10 Predicted effect of fiber weight fraction on the compressive


piezoresistive behavior of CFRPC system (a) overall behavior and (b) initial
behavior
118

Shear Parameter, Z (m4/N2)

0.000007

0.0000035

0
0

2
4
6
Fiber weight Fraction, (%)

Shear Parameter, BFRPC (GPa)

Modulus, E FRPC (GPa)

20

15

10

Compressive modulus

Tensile modulus
0
0

Fiber weight Fraction, (%)

100
80
60
40
20
0
0

Fiber weight Fraction, (%)

Figure 5.11 Variation of mechanical properties of CFRPC with fiber content

5.6.1 Tensile Behavior


5.6.2.1 Fiber Properties
Fig.5.13 shows the effect of Youngs modulus of carbon fibers (Ef) on the tensile
piezoresistivity of CFRPC system. Though increase in Youngs modulus of carbon fibers
(Ef) increased the piezoresistivity coefficient (111), the effect was very minimal. The

124

80
70
Compressive Stress, (- 11) (MPa)

FRPC

(a) Overall Behavior

FRPC
c

=0

= 0.15

60

50

30

6% CFRPC System

10

Refer Fig. 5.12 (b)

0
-10

10

20

30

= 0.40

t1 = 1.7315
= 0.064
crit = 0.028
EcFRPC = 17 GPa
cf = 228 GPa
cf = 0.1
BFRPC = 2.5 GPa
Bf = 1030 GPa
Z = 6.4 X 10-6

40

20

FRPC

40

50

60

Change in Longitudinal Resistivity, (/0)1 (%)

Compressive Stress, (- 11) (MPa)

7
6

FRPC

=0

FRPC

= 0 .1 5

FRPC

(b) Initial Behavior

t1 = 1.7315
= 0.064
crit = 0.028
EcFRPC = 17 GPa
cf = 228 Gpa
cf = 0.1
BFRPC = 2.5 GPa
Bf = 1030 GPa
Z = 6.4 X 10-6

4
3
2

6% CFRPC System

1
0
-0.7

= 0 .4 0

-0.5

-0.3

-0.1

0.1

0.3

0.5

Change in Longitudinal Resistivity, ( / 0 ) 1 (%)

Figure 5.12 Predicted effect of Poissons ratio of composite on compressive


piezoresistive behavior of CFRPC system (a) overall behavior and (b) initial
behavior

120

16

(a) Overall Behavior

Tensile Stress, (11) (MPa)

14

Etf = 100 GPa


Etf = 1000 GPa

12

6% CFRPC System

10

t1 = 1.7315
= 0.064
crit = 0.028
EtFRPC = 14 GPa
tFRPC = 0.15
tf = 0.1
BFRPC = 2.5 GPa
Bf = 1030 GPa
Z = 6.4 X 10-6

8
6
4
2

Refer Fig. 5.13 (b)

0
0

Tensile Stress, (11) (MPa)

2.5

0.5
1
1.5
2
Change in Longitudinal Resistivity, (/0)1 (%)

2.5

(b) Initial Behavior


Etf = 100 GPa

t = 1.7315
2 1
= 0.064
crit = 0.028
FRPC
= 14 GPa
1.5 Et
tFRPC = 0.15
tf = 0.1
FRPC
= 2.5 GPa
1 B
f
B = 1030 GPa
Z = 6.4 X 10-6

Etf = 1000 GPa

6% CFRPC System

0.5

0
0

0.02

0.04

0.06

0.08

0.1

0.12

Change in Longitudinal Resistivity, (/0)1 (%)

Figure 5.13 Predicted effect of Youngs modulus of fiber on tensile piezoresistive


behavior of CFRPC system (a) overall behavior and (b) initial behavior

121

16

Tensile Stress, (11) (MPa)

14

(a) Overall Behavior


f
t = 0.1

12

f
t = 0.5

6% CFRPC System

10

t1 = 1.7315
= 0.064
crit = 0.028
EtFRPC = 14 GPa
tFRPC = 0.15
tf = 228 GPa
BFRPC = 2.5 GPa
Bf = 1030 GPa
Z = 6.4 X 10-6

8
6
4
2

Refer Fig. 5.14 (b)

0
0

0.5

1.5

2.5

Change in Longitudinal Resistivity, (/0)1 (%)

2.5

(b) Initial Behavior

Tensile Stress, (11) (MPa)

t1 = 1.7315
= 0.064
crit = 0.028
EtFRPC = 14 GPa
1.5
tFRPC = 0.15
tf = 228 GPa
FRPC
= 2.5 GPa
1 B
f
B = 1030 GPa
Z = 6.4 X 10-6

tf = 0.1

tf = 0.5

6% CFRPC System

0.5

0
0.00

0.02

0.04

0.06

0.08

0.10

0.12

Change in Longitudinal Resistivity, (/0)1 (%)

Figure 5.14 Predicted effect of Poissons ratio of fiber on tensile piezoresistive


behavior of CFRPC system (a) overall behavior and (b) initial behavior

122

piezoresistivity coefficient (111) increased by 5% when the Youngs modulus of fibers


was varied from 100 GPa to 1000 GPa.

The effect of Poissons ratio of carbon fibers (f) on the tensile piezoresitivity of
CFRPC system is shown in Fig.5.14. Though decrease in Poissons ratio of carbon fibers
(f) increased the piezoresistivity coefficient (111), the effect was very minimal.

5.6.2.2 Fiber Reinforced Polymer Concrete Properties


Fig.5.15 shows the effect of carbon fibers weight fraction () on the tensile
piezoresistivity of CFRPC system. The mechanical properties of CFRPC system were
assumed to vary according to Fig.5.11. Clearly, the piezoresistivity coefficient (111)
decreased with increase in carbon fibers weight fraction ().

Fig.5.16 shows the effect of the Poissons ratio of CFRPC system (FRPC) on its
tensile piezoresistive behavior. Increase in Poissons ratio from 0 to 0.5 decreased the
piezoresistivity coefficient (111) by 100%. Since the Poissons ratio of plain polymeric
composite (c) is in the range 0.4-0.5 (Vipulanandan and Mantrala, 1995) and that of
CFRPC is in the range of 0.15-0.2, it can be said that presence of sand improved the selfmonitoring sensitivity of CFRPC when compared to the plane polymeric composite.

5.7 Summary
Percolation theory was used to model the insulator-conductor transition occurring in
CFRPC system when conducting carbon fibers a reached threshold concentration. A
125

16

Tensile Stress, ( 11) (MPa)

4%

6%

14

3%

(a) Overall Behavior

12
10
8

t1 = 1.7315
crit = 0.028
tf = 228 GPa
tf = 0.1
Bf = 1030 GPa
Z = 6.4 X 10-6

6
4
2

Refer Fig. 5.15 (b)

0
0.0

1.0

2.0

3.0

4.0

5.0

6.0

7.0

8.0

Change in Longitudinal Resistivity, ( / 0 ) 1 (%)

Tensile Stress, (11) (MPa)

2.5

(b) Initial Behavior


6%

3%

1.5

t1 = 1.7315
crit = 0.028
tf = 228 GPa
tf = 0.1
Bf = 1030 GPa
Z = 6.4 X 10-6

0.5

0
0.00

0.02

0.04

0.06

0.08

0.10

0.12

0.14

0.16

0.18

Change in Longitudinal Resistivity, (/0)1 (%)

Figure 5.15 Predicted effect of fiber weight fraction on composite tensile


piezoresistive behavior (a) overall behavior and (b) initial behavior

124

16

(a) Overall Behavior

Tensile Stress, ( 11) (MPa)

14

FRPC

= 0.5

FRPC

= 0.3
FRPC

=0

12
10

t1 = 1.7315
= 0.064
crit = 0.028
EtFRPC = 14 GPa
tf = 228 Gpa
tf = 0.1
BFRPC = 2.5 GPa
Bf = 1030 GPa
Z = 6.4 X 10-6

8
6

6% CFRPC System

4
2

Refer Fig. 5.16 (b)

0
0.00

0.50

1.00

1.50

2.00

2.50

3.00

Change in Longitudinal Resistivity, ( / 0 ) 1 (%)

Tensile Stress, (11) (MPa)

2.5

(b) Initial Behavior

FRPC

FRPC
t

= 0.5

tFRPC = 0

= 0.3

t1 = 1.7315
= 0.064
crit = 0.028
EtFRPC = 14 GPa
tf = 228 Gpa
tf = 0.1
BFRPC = 2.5 GPa
Bf = 1030 GPa
Z = 6.4 X 10-6

1.5

6% CFRPC System

0.5

0
0.00

0.02

0.04

0.06

0.08

0.10

0.12

0.14

Change in Longitudinal Resistivity, (/0)1 (%)

Figure 5.16 Predicted effect of Poissons ratio of composite on composite tensile


piezoresistive behavior (a) overall behavior and (b) initial behavior

125

simple analytical model was developed based on the principle of continuum mechanics
and percolation theory to explain the piezoresistive behavior of CFRPC system under
uniaxial compression and uniaxial tension. The sequence of modeling and important
findings from the modeling are listed below:
1. Percolation theory predicted the resistivity of CFRPC system reasonably well.
The value of the conductivity exponent (t1) was 1.732 and was found to be in
good agreement with the universally accepted value for 3-D insulator-conductor
multiphase solids.
2. A general 3-D analytical model was developed to explain the piezoresistive
behavior of CFRPC system and later simplified for uniaxial compression and
uniaxial tension. It has been demonstrated that the piezoresistivity of CFRPC
arises due to two factors: the change in volume concentration of conducting
carbon fibers with stress and the change in micro-structure of the heterogeneous
CFRPC with stress. The former is a function of the both the hydrostatic and the
deviatoric component of the stress tensor and the later was assumed to be purely a
function of deviatoric stress. The model predicted the change in resistivity with
stress reasonably well up to 70% the failure stress of CFRPC systems for both
uniaxial compression and uniaxial tension. The model under predicted the change
in resistivity at higher stress level. This might because of formation of cracks in
the material at higher stress level.
3. Parametric study was performed to understand the effect of individual parameter
on the piezoresistive behavior of the CFRPC system. From parametric study it can
be summarized that the carbon fiber Youngs modulus (Ef) and the carbon fiber

126

Poissons ratio (f) had significant effect on the compressive piezoresistive


behavior, though they do not have much influence on the tensile piezoresistive
behavior. High Youngs modulus of the carbon fibers (Ef) increased both the
threshold change in resistivity and the piezoresistivity coefficient (111).
Similarly, low Poissons ratio of the carbon fibers (f) increased both the
threshold change in resistivity and the initial piezoresistivity coefficient (1110).
Sensitivity of CFRPC system in both compression and tension decreased with
increase in fiber weight fraction () as it reduced the piezoresistivity coefficient
(111). It was also shown that higher value of Poissons ratio of the composite (c)
reduced the self-monitoring sensitivity of the composite.

127

CHAPTER 6
APPLICATIONS OF SELF-MONITORING BEHAVIOR OF
CFRPC IN STRUCTURAL ELEMENTS
6.1 Introduction
Bending is the most commonly encountered loading in civil engineering structures.
In this chapter the piezoresistive behavior of CFRPC system under beam and pipe
configurations were investigated. CFRPC system was subjected to two different types of
bending behavior, CFRPC beam under four-point loading and PVC-CFRPC-PVC
sandwich ring under diametrically opposite compressive loading. In both the cases, the
change in resistance behavior with load was studied. The beam response was modeled
analytically using the uniaxial compressive and uniaxial tensile response, presented in
Chapter 4 and modeled in Chapter 5, and was compared with experimental observations.

6.2 CFRPC Beam Under Four-Point Bending Test


6.2.1 Specimen Design and Instrumentation
Both 3% and 6% CFRPC systems were studied under four-point loading using
300 mm X 50 mm X 50 mm prismatic specimens, prepared as per CIGMAT PC-1
(Appendix). Specimens were instrumented to measure the mechanical and electrical
properties simultaneously. Strain gages were glued to the specimens both at the top
(compression) and bottom (tension) faces of the beam to measure the variation of strain
along the cross-sections of the specimens during loading [Fig.6.1 (b)]. 3-conductor cables
were inserted in to the specimens during preparation both in the compression side and the
tension side to measure the changes in resistances during loading [Fig.6.1 (b)].
131

6.2.2 Testing Program


Two specimens of each CFRPC system were tested in minimum of four cycles on
loading up to fraction of failure load. The typical four-point bending test set up is shown
in Fig.6.1(a). The tests were performed as per CIGMAT PC-4 (Appendix), using the
Instron Universal Testing Machine (Model 1322) under displacement-controlled mode
with a crosshead speed of 0.0635 mm/min. Loads were measured from the machine readout with an accuracy of 0.5 lbs. During loading strains () and resistances (R) were
measured simultaneously. Strains were measured using 10 mm strain gages, having a
least count of 10-6 mm/mm, on both extreme compression fiber and extreme tension fiber.
The DC electrical measurements were made along the length of the beam both on the
compression side and on the tension side using the Hewlett Packard 34420A Nano Volt /
Micro Ohm Meter, which had a least count of 1. In order to insulate the specimen
from the testing machine paper was used as shown in Fig.6.1(b).

6.2.3 Test Results and Discussions


Similar to the uniaxial compression and tension measurements, in bending also,
the specimens were subjected to a higher stress at the first cycle, compared to the
subsequent cycles, to get better reproducibility in electrical measurements. All the results
reported were done using the same procedure.

132

(a)

Instraon bending
testing machine

Specimen, Refer
Fig. 6.1(b)

Machine control and


read-out

Strain gage
read-out

Resistance mater

(b)

Wires for compression side


longitudinal resistance
measurement

Wires to strain
gage read-out

Insulator

Wires for tension side


longitudinal resistance
measurement

Figure 6.1 Four-point bending test (a) the complete test set up and (b) the
specimen

133

The self-monitoring sensitivity of the beam specimens will be characterized by


piezoresistance coefficient and gage factor. In line with the piezoresistivity coefficient, as
defined in Chapter 4 for uniaxial compressive and uniaxial tensile behavior, the
instantaneous piezoresistance coefficient (Cbeam) for beam bending can be defined as
follows:

R
= C beam P ..(6.1)
R

where R / R is the change in resistance of the beam specimen for a load increment of

P and Cbeam is the instantaneous piezoresistance coefficient in beam bending. The


piezoreistance coefficient in beam bending has the dimension of reciprocal load (1/N).
Similarly, the instantaneous gage factor (Mbeam) for beam bending can be defined as
follows:

R
= M beam ..(6.2)
R

where R / R is the change in resistance of the beam specimen for an increase in midspan deflection and M is the instantaneous gage factor in beam bending. The gage factor
in beam bending has the dimension of reciprocal deflection (1/mm). It should be noted
that though the piezoresistivity coefficient (ijk) and gage factor (Mijk), as defined in
Chapter 4, are material constants, the piezoresistance coefficient (Cbeam) and gage factor
(Mbeam) for beam bending depend upon the dimensions of the specimen and hence not

134

comparable with those defined in Chapter 4. The instantaneous stiffness (kbeam) of the
beam under four-point bending can be calculated by coupling Eqn.6.1 and Eqn.6.2 as
follows:

P =

M beam
= k beam (6.3)
beam
C

Figs.6.2(a) and 6.2(b) show the results of simultaneous mechanical and


electrical measurements of 3% CFRPC beam specimen under four-point bending in terms
of load with change in resistance (Fig.6.2(a) and mid-span deflection with change in
resistance (Fig.6.2(b) respectively. The resistance initially decreased with load but after a
certain threshold value it increased with increase in load, resembling the behavior under
uniaxial compression. The initial instantaneous piezoresistance coefficient (C0beam) and
initial instantaneous gage factor (M0beam) obtained were 2.8 X 108 1/N and -5 X 104
1/mm respectively. The variations of piezoresistance coefficient and gage factor with
load have been plotted in Figs.6.3 and 6.4, where the instantaneous and the secant values
are compared. Though both the piezoresistance coefficient (Cbeam) and gage factor
(Mbeam) increased non-linearly with the increase in load, the ratio of gage factor to
piezoresistance coefficient (Mbeam/Cbeam), which is the stiffness (kbeam) of the beam,
decreased linearly with load as shown in Fig.6.5.

The mechanical and electrical responses of 6% CFRPC beam specimen under


four-point loading were shown in Figs 6.6(a) and 6.6(b). The difference in behavior of
6% beam specimen with 3% beam specimen must to be noted. The resistance of 6%

135

beam specimen always increased with increasing load, resembling the behavior under
uniaxial tension. The initial piezoresistance coefficient (C0beam) and initial gage factor
(M0beam) obtained were 1.4 X 108 1/N and 2 X 104 1/mm respectively. The variations of
piezoresistance coefficient, gage factor and stiffness with load are shown in Figs.6.7
through 6.9. Comparing the mechanical and electrical measurement results of identical
CFRPC beam specimens with varied fiber content, it can be concluded that though
increase in fiber content increased the stiffness of the beam, it deteriorated the selfmonitoring sensitivities (Cbeam and Mbeam) of the beam.

6.2.4 Modeling
From the electrical resistance point-of-view, the beam specimen can be idealized
as a combination of resistors connected in series and parallel to each other as shown in
Fig.6.10. Now, if it is assumed the material is isotropic and homogeneous, at unstressed
state the resistances of all the resistors are equal in the same direction. But due to
application of load, because of different stress level in different cross-sections of the
specimen, the resistors will have different resistances depending upon the magnitude and
type of stress. Hence, the equivalent resistance, R, between the points A & B or C & D
(Fig.6.10) can be written as

1
1
1
=
+
..(6.4)
R R1 + R 2 + R3 R 4 + R5 + R6

Now, lets calculate R1 at location 1 due to increase in stress. The resistance is given by

136

2.5

Load, (P) (kN)

3% CFRPC
1.5

cycle 2 - compression side


cycle 3 - compression side
cycle 4 - tension side
cycle 5 - tension side

0.5

0
-0.02

0.02

0.04

0.06

0.08

0.1

Change in Resistance, (R/R0) (%)

0.16

Midspan Deflection, ( ) (mm)

0.14
0.12

3% CFRPC

0.1
0.08
0.06

cycle 2 - compression side


cycle 3 - compression side

0.04

cycle 4 - tension side

0.02

cycle 5 - tension side


0
-0.02

0.02
0.04
0.06
Change in Resistance, (R/R0)(%)

0.08

0.1

Figure 6.2 Mechanical and electrical Response of 3% CFRPC beam under fourpoint bending test (a) load versus change in resistance relationship and (b) mid-span
deflection versus change in resistance relationship

137

Piezoresistance Coefficient, (Cbeam) (1/kN)

0.0018

3% CFRPC

0.0016

Instantaneous

0.0014
0.0012
0.001
0.0008
0.0006
0.0004
0.0002
0
-0.0002

Secant

-0.0004
0

0.5

1
1.5
Load, (P) (kN)

2.5

Figure 6.3 Variation of piezoresistance coefficient in bending of 3% CFRPC beam


specimen under four-point bending test with load
0.03

3% CFRPC

Instantaneous

) (1/mm)

0.025

beam

0.02

Gage Factor, (M

0.015
0.01
0.005
0

Secant
0

0.5

1
1.5
Load, (P) (kN)

-0.005
2.5

Figure 6.4 Variation of gage factor in bending of 3% CFRPC beam specimen


under four-point bending test with load

138

Secant Stiffness, (kbeam) (kN/mm)

25

20

15

3% CFRPC
10

0
0

0.5

1
1.5
Load, (P) (kN)

2.5

Figure 6.5 Variation of secant stiffness in bending of 3% CFRPC beam specimen


under four-point bending test with load

R=

l
(6.5)
A

where, = resistivity of the material; l = length and A = area of cross-section of the


specimen. Now taking logarithm on both sides and differentiating Eqn.6.5 with respect to
the stress tensor we have,

1 R
1
1 l
1 A
(6.6)
+

=
R jk jk l jk A jk

Neglecting the small change in l and A Eqn.6.6 can be written as

139

2.5

Load, (P) (kN)

6% CFRPC
1.5

cycle 2 - compression side


cycle 3 - compression side

0.5

cycle 4 - tension side


cycle 5 - tension side
0
0

0.05

0.1

Change in Resistance, (R/R0) (%)

0.14

Midspan Deflection, () (mm)

0.12
0.1

6% CFRPC

0.08
0.06

cycle 2 - compression side

0.04

cycle 3 - compression side


cycle 4 - tension side

0.02

cycle 5 - tension side


0
0

0.02

0.04

0.06

0.08

0.1

Change in Resistance, (R/R0) (%)

Figure 6.6 Mechanical and electrical response of 6% CFRPC beam under fourpoint bending test (a) load versus change in resistance relationship and (b) mid-span
deflection versus change in resistance relationship

140

0.0007
Piezoresistance Coefficient, (Cbeam) (1/kN)

6% CFRPC

Instantaneous

0.0006
0.0005
0.0004
0.0003
0.0002
0.0001

Secant

0
-0.0001

0.5

1
1.5
Load, (P) (kN)

2.5

Figure 6.7 Variation of piezoresistance coefficient in bending of 6% CFRPC beam


specimen under four-point loading with load

0.014

Instantaneous

Gage Factor, (M beam) (1/mm)

6% CFRPC

0.012
0.01
0.008
0.006
0.004
0.002

Secant

0
-0.002
0

0.5

1
1.5
Load, (P) (kN)

2.5

Figure 6.8 Variation of gage factor in bending of 6% CFRPC beam specimen


under four-point loading with load

141

Secant Stiffness, (kbeam) (kN/mm)

30
25
20
15

6% CFRPC
10
5
0
0

0.5

1.5

2.5

Load, (P) (kN)

Figure 6.9 Variation of secant stiffness in bending of 6% CFRPC beam specimen


under four-point bending test with load

1 R
1
(6.7)
=
R jk jk

Hence, combining Eqn.6.7, Eqn.5.14 can be written as

1 R

jk R0 i

1
1
= t i ( crit )
S
K f jk + jk
ll

K c jk + jk

ll
fiber

+ t i ( crit ) ZS jk

composite

.(6.8)
Hence, at location 1, the fractional change in resistance in the i-direction due to the stress
tensor ij can be written as: (from Eqn.5.16)

142


R1
1

= t i ( crit )1 c

K jk + B c jk
R10 i

f
K +Bf

jk
composite

jk


+ t i ( crit )1 ZS jk d jk

fiber

.(6.9)

R1

R2

R3

R6

R5

R4

Figure 6.10 Idealized beam specimen

Now, in case of specimen subjected to four-point loading, the stress tensor


(normal stress (11) and shear stress (12) across the cross-section) can be written in terms
of mean stress and deviatoric stress as

11 12

21 0
0
0

where 11 =

11
0 3
0 = 0

0
0

11
3
0

2
0 11
3

0 + 21

11
0
3

12

11
3
0

0 .(6.10)

11

M
VQ
y and 12 = 21 =
where M is the bending moment at any crossI
Ib

section; I is the moment of inertia of the cross-section; y is the distance of any point from
the neutral axis; V is the shear force at the cross-section; Q is the first area moment

143

Q = ydA ; b is the width of the cross-section. Hence normal stress (11) is a function

of bending moment (M) and distance from the neutral axis (y) and deviatoric stress (12)
is a function of shear force (V) and square of distance from the neutral axis (y2). The
variations of shear force and bending moment along the length of the beam under fourpoint loading have been shown in Figs.6.11(b) and 6.11(c) respectively. Hence, the
respective normal stresses (11) and deviatoric stresses (12) at the idealized resistor
locations (Fig.6.10) can be calculated as a function of applied load (P) and physical
dimensions (length, L; breadth, b and depth, d) as follows:

Location 1/3/4/6 (X = L/6)

11

PL
M
d
PL
=
.(6.11)
y = 63 =
I
bd 4 2bd 2
12

12 = 21 =

VQ
=
Ib

Pb

d d d
+
4 4 8 9P
=
..(6.12)
8bd
bd 3
b
12

Location 2/5 (X = L/2):

11

PL
M
d
PL
=
y = 3 3 = 2 ...(6.13)
I
bd 4 bd
12

144

2
P

d/4

d/4

d/4
d/4

L/3

L/3

(a)

L/3

P
P

(b)

PL/3

PL/3

(c)
Figure 6.11 Beam under four-point bending test (a) the stresses at different
locations (b) shear force diagram and (c) bending moment diagram

145

12 = 21 =

VQ
= 0 .(6.14)
Ib

Hence, considering both CFRPC and fiber as isotropic and homogeneous


(i.e. K =

E
and B11 = B12 = B21 = B), at location 1, fractional change in resistance
3(1 2 )

in the 1-direction (along the longitudinal axis of the beam) can be written with respect to
normal stress and deviatoric stress as

R1

= t1 ( crit
R1 0 1

+ t1 ( crit

)1

1
1
2 Z 11
1
)

11
+
3

E cFRPC
E cf
FRPC
f

+B
+B

FRPC
f

)
3
(
1
2

)
3(1 2 c


1
FRPC

1
Bf

+ Z 12 21

..(6.15)

and expressing the change in resistance at location 1 with respect to the load (P) [from
Eqns.6.11, 6.12 and 6.15], we have

L

2Z
P
2
R1
1
1

1
L P
bd
2

= t1 ( crit )

+
FRPC
f
2bd 2

Ec
Ec
R10 1
FRPC
f


+B
+B

FRPC
f

3(1 2 c )

3(1 2 c )
1
1
9 P 9
1
P
+ t1 ( crit ) FRPC f + Z

8 bd 8bd
B
B
..(6.16)

146

Using the value of compressive Youngs modulus (EFRPC) and Poissons ratio
(FRPC) of fiber reinforced polymer concrete (obtained in Chapter 3), Youngs modulus
(Ef) and Poissons ratio (f) of the carbon fibers (obtained from manufacturers
specification) and the shear parameters of the CFRPC and carbon fibers (BFRPC, Z and Bf
obtained from curve-fitting of uniaxial compression test results), the differential
equation of change in resistance (R1/R10) at location 1 of beam specimen under fourpoint loading (Eqn.6.16) was solved incrementally, updating the value of (Eqn.5.24)
and crit (Eqn.5.28) for each small load increment (P). The total change in resistance at
location 1 for a given stress was calculated as the cumulative change in resistance
[(R1/R10)] for each small load increment.

Under four-point loading at location 1, Eqn.5.24 reduced to


1
1
1
1
FRPC FRPC
f f
FRPC
FRPC
K
K
B
B
B

+
c
c +B

1
1
1
1

f
FRPC
FRPC
FRPC
FRPC
B
B
Kc +B

1
1
1
1

f
FRPC
FRPC
FRPC FRPC
B
B
B
K
B

+
c

= 0e

1
f

B
1

Kcf +B f
1
Bf

B
11 12 0

1

21 0 0
f
B

0 0 0

1

f
f
Kc +B
1

..(6.17)


1
1
FRPC FRPC f f

Kc +B
Kc +B

= 0e

11+2 1 1 12

BFRPC B f

147

..(6.18)

PL 1
1 9P
+2

bd2 BFRPC B f 8bd


1
1
FRPC FRPC f f
Kc +B
Kc +B

= 0e

...(6.19)

and Eqn.5.28 reduced to

crit = ( crit )0

2 11
3
1
+ Z 21
2

12

11
3
0

2
0 11
3

0 21

11

0
3

12

11
3
0

0 .(6.20)

11

= (crit )0 + Z 11 + 12 ..(6.21)
3

9
L
= (crit )0 + Z
+
P ..(6.22)
2
6bd 8bd

Similarly, the changes in resistances at different locations (R2, R3, R4, R5 and
R6) of the idealized beam were calculated and their individual variations with load were
shown in Figs6.12 through 6.15 for both 3% and 6% CFRPC systems. The total
resistance between the points A & B or C & D was predicted from Eqn.6.4. The model
predictions of the change in resistance with applied load were compared with
experimental observations in Figs.6.16 and 6.17 for 3% CFRPC system and 6% CFRPC
system respectively.

148

4.5

6% CFRPC system

4.0
3.5

3% CFRPC system
Load, (P) (kN)

3.0
2.5

R1

R3

2.0
1.5
1.0
0.5

-0.08

-0.06

-0.04

-0.02

0.02

0.0
0.04

Change in Resistance, (R/R0) (%)

Figure 6.12 Predicted variations of change in resistance at locations R1 and R3


(compression and shear) with load for 3% and 6% CFRPC beam specimen
4.5
4.0
3.5

6% CFRPC system

3.0

Load, (P) (kN)

3% CFRPC system

2.5
2.0

R2

1.5
1.0
0.5
0.0

-0.3

-0.2

-0.1

0.1

0.2

0.3

0.4

0.5

Change in Resistance, (R/R0) (%)

Figure 6.13 Predicted variations of change in resistance at location R2 (pure


compression-no shear) with load for 3% and 6% CFRPC beam specimen

146

4.5
4.0
3.5

3% CFRPC system
Load, (P) (kN)

3.0
2.5
2.0

R4

1.5

R6

1.0

6% CFRPC system

0.5
0.0

0.1

0.2

0.3

0.4

0.5

Change in Resistance, (R/R0) (%)

Figure 6.14 Predicted variations of change in resistance at locations R4 and R6


(tension and shear) with load for 3% and 6% CFRPC beam specimen
4.5
4.0

6% CFRPC system

3.5
Load, (P) (kN)

3% CFRPC system

3.0
2.5
2.0
1.5

R5

1.0
0.5
0.0

0.2

0.4
0.6
0.8
1
Change in Resistance, (R/R0) (%)

1.2

Figure 6.15 Predicted variations of change in resistance at location R5 (pure


tension no shear) with load for 3% and 6% CFRPC beam specimen
147

Load, (P) (kN)

cycle 2 - compression side


cycle 3 - compression side
cycle 4 - tension side
cycle 5 - tension side

2.5

1.5

3% CFRPC
1

0.5

Model
0
-0.1

-0.05

0.05

0.1

Change in Resistance, (R/R0) (%)

Figure 6.16 Experimental and model prediction of piezoresistance of 3% CFRPC


system under four-point bending test (a) overall behavior and (b) initial behavior

2.5

Load, (P) (kN)

cycle 2 - compression side


cycle 3 - compression side
cycle 4 - tension side
cycle 5 - tension side

1.5

6% CFRPC

0.5

Model
0
-0.1

-0.05

0.05

0.1

Change in Resistance, (R/R0) (%)

Figure 6.17 Experimental and model prediction of piezoresistance of 6% CFRPC


system under four-point bending test (a) overall behavior and (b) initial behavior

148

6.3 PVC-CFRPC-PVC Sandwich Ring


6.3.1 Specimen Design and Instrumentation
6% CFRPC system was tested under parallel plate loading. CFRPC system was
used as the core material in the sandwich ring developed at CIGMAT laboratory
(Mantrala, 1996). Figure 6.18 (b) shows the design configurations of the sandwich ring
with 6% CFRPC system as the core material. PVC rings used in this research were
conforming to ASTM D 3034-89. The nominal inside diameters of the outer and inner
PVC rings were 304.8 mm and 203.2 mm respectively. This resulted in a pipe wall
thickness of about 50.8 mm and a pipe radius to wall thickness ratio of about 3. The
sandwich ring was cast by placing and compacting the CFRPC in the annular space
between two PVC rings. The CFRPC core material was prepared as per CIGMAT PC-1
(Appendix). A base plate and a top plate made of aluminum were used to keep the PVC
rings concentrically in place. The core material was compacted in three layers, at 25
blows per layer all around the circumference. The curing procedure is described in
CIGMAT PC-1 for polymer concrete specimens with and without fibers. This sandwich
ring specimen was instrumented to measure the mechanical and electrical properties
simultaneously. Strain gages were glued to the outside and inside surfaces of both the
outer and inner PVC rings at the crown, invert and spring line locations before casting. 3conductor cables were inserted in to the core material during casting both at the top and
bottom fiber locations at both invert and spring line of the sandwich ring to measure
change in resistances during loading.

149

6.3.2 Testing Program


One PVC-CFRPC-PVC sandwich ring specimen was tested in eight cycles by
loading up to a fraction of failure load. The typical test set up is shown in Fig.6.18(a).
The test were performed using the Tinius Olsen Universal Testing Machine under
displacementcontrolled mode. During loading the change in diameter of the ring and
resistances were measured simultaneously. Change in diameter was measured using
strain gages glued to the outer surface of the inner PVC ring and recorded using a
standard data acquisition system. A dial gage was also used to measure the change in ring
diameter. The relation between circumferential strain on the outer surface of the inner
PVC ring at crown location and the change in sandwich ring diameter is plotted in
Fig.6.19. DC electrical measurements were made between crown and invert and between
the spring lines during loading with the help of 3-conductor cables embedded in to the
specimen during casting and recorded using Hewlett Packard 34420A Nano Volt/ Micro
Ohm Meter.

6.3.3 Test Results and Discussions


Similar to the compression, tension and beam bending measurements, the
sandwich ring specimen was also subjected to a higher stress at the first cycle, compared
to the subsequent cycles, to get better reproducibility in electrical measurements.

The self-monitoring sensitivity of PVC-CFRPC-PVC sandwich ring was


characterized by piezoresistance coefficient and gage factor in ring bending. The
piezoresistance coefficient in ring bending (Cring) was defined as

150

Data acquisition system


for strain gage readings

Compression
testing machine

Ring specimen,
Refer Figure 6.14 (b)

Resistance meter
(a)

PVC ring

6% CFRPC system
(core material)

Wires to strain gage data


acquisition system
3-Conductor cables for
resistance measurement

(b)

Figure 6.18 PVC-CFRPC (6% fiber)-PVC sandwich ring under parallel plate
loading (a) test set up for simultaneous mechanical and electrical measurements and
(b) the specimen

151

Change in Diameter, mm

3.5

Strain Gage

3
2.5

Dial Gage

2
1.5
1
0.5

y = 1226.9x + 39.832x

0
0

0.01

0.02

0.03

0.04

0.05

Strain in Outer Surface of Inner PVC Ring, %

.
Figure 6.19 Calibration of strain gage readings with dial gage readings

R
P
= C ring (6.23)
R
L

where R / R is the change in resistance between two diametrically opposite points of the
sandwich ring specimen for an increment ( P / L) of load per unit length and Cring is the

instantaneous piezoresistance coefficient in ring bending. The piezoreistance coefficient


in ring bending had the dimension (meter/Newton). Similarly the instantaneous gage
factor in ring bending (Mring) can be defined as follows:

R
= M ring ..(6.24)
R

152

where R / R is the change in resistance between two diametrically opposite points of the
sandwich ring specimen for a change in diameter of and Mring is the instantaneous
gage factor in ring bending. The gage factor in ring bending (Mring) had the dimension of
(1/meter). Similar to beam bending the piezoreistance coefficient and gage factor in ring
bending were not pure material property; it also depended upon the dimensions of the
ring. The instantaneous stiffness of the ring (kring) under parallel plate loading was
calculated from the simultaneous mechanical and electrical measurements by coupling
Eqn.6.23 and Eqn.6.24 as follows:

ring
P M
= ring = k ring .(6.25)
L C

Figs.6.20 and 6.21 show the results of simultaneous mechanical and electrical
measurements of PVC-CFRPC (6%)-PVC sandwich ring specimen under parallel plate
loading in terms of load versus change in resistance and change in diameter versus
change in resistance respectively. The resistance continuously increased with increase in
load, resembling the uniaxial tension behavior. The initial instantaneous piezoresistance
coefficient (C0ring) and initial instantaneous gage factor (M0ring) in ring bending obtained
were 20 X 10-3 mm/kN and 30 X 10-4 1/mm respectively. Similar to beam bending in
sandwich ring also the self-monitoring sensitivity, the piezoresistance coefficient and the
gage factor, increased with increasing load. The variations of piezoresistance coefficient
and gage factor have been plotted in Figs.6.22 and 6.23. In both the plots both the
instantaneous and secant values have been plotted. Piezoresistance coefficient (Cring)
increased non-linearly and gage factor (Mring) increased linearly with load but the ratio of
153

12

Coefficient of correlation = 0.96


10

Load, (P) (kN)

8
6

Cycle 2 - crown-invert
Cycle 3 - crown-invert
Cycle 4 - crown-invert
Cycle 5 - springline
Cycle 6 - springline
Cycle 7 - springline

4
2
0
0

0.2

0.4

0.6

0.8

1.2

1.4

1.6

Change in Resistance, (R/R0) (%)

Figure 6.20 Load versus change in resistance behavior of


PVC-CFRPC (6%)-PVC sandwich ring under parallel plate loading
4

Coefficient of correlation = 0.99

Change in Diameter, () (mm)

3.5
3
2.5
2

Cycle 2 - crown-invert
Cycle 3 - crown-invert
Cycle 4 - crown-invert
cycle 5 - springline
Cycle 6 - springline
Cycle 7 - springline

1.5
1
0.5
0
0

0.2

0.4

0.6

0.8

1.2

1.4

1.6

Change in Resistance, (R/R0) (%)

Figure 6.21 Change in diameter versus change in resistance behavior of


PVC-CFRPC (6%)-PVC sandwich ring under parallel plate loading

154

Piezoresistance Coefficient, (C ring) (mm/kN)

0.25

Seacnt piezoresistance coefficient


Initial piezoresistance coefficient

0.2

0.15

Instantaneous
0.1

Secant
0.05

0
0

10

12

Load, (P) (kN)

Figure 6.22 Variation of piezoresistance coefficient with load for


PVC-CFRPC (6%)-PVC sandwich ring under parallel plate loading
0.012

Initial piezoresistance coefficient


Seacnt piezoresistance coefficient

Gage Factor, (M ring)(1/mm)

0.01
0.008

Instantaneous

0.006
0.004
0.002

Secant

0
0

6
Load, (P) (kN)

10

12

Figure 6.23 Variation of gage factor with load for PVC-CFRPC (6%)-PVC
sandwich ring under parallel plate loading

155

gage factor to the piezoresistance coefficient (Mring/Cring), which is the stiffness of the
sandwich ring (kring), decreased linearly with load as shown in Fig.6.24.

Ring Stiffness, (k ring) (MPa)

250

200

150

100

50

10

12

Load, (P) (kN)

Figure 6.24 Variation of ring stiffness with load for PVC-CFRPC (6%)-PVC
sandwich ring under parallel plate loading

6.4 Summary
The applicability of piezoresistive CFRPC system as a bulk sensor in structural
application has been studied. CFRPC beam under four-point loading and PVC-CFRPCPVC sandwich ring under parallel plate loading showed that the applied load or the
deflection due to that load can be predicted by monitoring the change in resistance
behavior between two points inside those structural elements. Based on the experimental
and modeling study the following can be summarized:

156

1. Piezoresistance coefficient (C) and gage factor (M) can be used to characterize the
bulk sensing property of the structural elements, CFRPC beam and PVC-CFRPCPVC sandwich ring. (It should be noted that in beam or ring bending the
piezoresistanace coefficients and gage factors are not purely material property.
These also depend upon the dimensions of the structural element).
2. In beam bending the self-monitoring sensitivity reduced with increased fiber
content, similar to the response in uniaxial compression and uniaxial tension. Also
it was observed that the trend of change in resistance behavior with load was
different for 3% CFRPC beam as compared to 6% CFRPC beam. The resistance
of 3% CFRPC beam initially decreased and then after a threshold change in
resistance it increased with increasing load, resembling the uniaxial compressive
behavior. The resistance of 6% CFRPC beam always increased with increasing
load, resembling the uniaxial tensile behavior. Analytical modeling showed that
this difference in trend was because of difference in the shear parameter, Z for 3%
6% CFRPC system.
3. In beam bending similar to uniaxial compression and uniaxial tension, the selfmonitoring sensitivity was higher at higher load. Both the piezoresistance
coefficient (Cbeam) and gage factor (Mbeam) increased non-linearly with increase in
load, with the rate of change of these coefficients were higher for the 3% CFRPC
beam as compared to the 6% CFRPC beam.
4. A series-parallel resistors idealization of beam predicted the change in resistance
behavior with load for both 3% CFRPC beam and 6% CFRPC beam and the
predictions agreed well with the experimental results.

157

5. In sandwich ring response under parallel plate loading, the self-monitoring


sensitivity was observed to be higher at higher load. Piezoresistance coefficient
(Cring) increased non-linearly and gage factor (Mring) increased linearly with load.

158

CHAPTER 7
CONCLUSIONS AND RECOMMENDATIONS
7.1 Conclusions
This study was focused on characterizing the structural and self-monitoring
behavior of fiber reinforced polymer concrete. Glass and carbon fibers were used as
matrix reinforcement. The structural properties of interest in civil engineering
applications such as compressive strength, tensile strength, compressive modulus, tensile
modulus, stress-strain relationship and non-destructive properties such as low strain
moduli and damping ratio were studied experimentally for polymer concrete with and
without fibers. Strength, stiffness and stress-strain models were used to predict the
experimental observations. Self-monitoring behavior of fiber reinforced polymer concrete
was quantified in terms of compressive piezoresistive coefficient, tensile piezoresistive
coefficient, compressive gage factor and tensile gage factor. An analytical model was
developed by combining the principle of percolation theory and continuum mechanics to
predict the stress-change in resistivity relationship. Based on the experimental and
analytical investigation the following conclusions can be advanced:

1. The tensile strength of the PC system improved by 85% and 60% with the
addition of 6% glass fibers and 6% carbon fibers respectively. The tensile strength
model predicted the strength increase in PC systems with fibers. Addition of glass
fibers improved the compressive strength of polymer concrete, but carbon fibers
did not.

159

2. Polymer concrete with glass and carbon fibers behaved as a bi-modulus material
with tensile-to-compressive modular ratio of 0.85 and 0.78 respectively as
compared to 0.75 for polymer concrete without fibers.

3. The p-q model for stress-strain relationship predicted both the pre-peak and postpeak behavior with a single function for both tension and compression. While the
parameter q influenced the stress-strain relationships before peak, the parameter p
influenced the toughness, mainly the post-peak toughness.

4. The impact resonance and the pulse velocity methods can be used to characterize
the behavior of PC with and without fibers. The ratio of shear wave velocity to Pwave velocity for 14% PC was 0.63. The addition of polymer or fibers did not
influence the velocity ratio. While the pulse velocity method was independent of
specimen shape, the impact resonance method was dependent on specimen shape
with dynamic shear modulus and dynamic Poissons ratio having the greatest
effect. Dynamic Youngs moduli and dynamic shear moduli obtained from the
pulse velocity method were within 10% of the respective static moduli; whereas
those from the impact resonance test showed larger variations. Dynamic Poissons
ratio obtained from the pulse velocity method was closer to static Poissons ratio.
Wave velocities measured from the pulse velocity test were higher than those
determined from the impact resonance method.

160

5. The damping ratio increased with increase in polymer and fiber contents in the PC
systems. Glass fibers increased the damping ratio in the longitudinal, flexural, and
torsional modes. Carbon fibers increased the longitudinal damping ratio but it also
reduced the flexural and torsional damping ratios.

6. Polymer concrete system with carbon fibers was found to exhibit two unique
functional properties, electrical conduction and piezoresistance, which polymer
concrete system with glass fibers and polymer concrete system without fibers did
not. Piezoresistive properties of polymer concrete system with carbon fibers can
be utilized to self-monitor the material. In monitoring itself against stress,
strain or fracture, the material emits signal in terms of electrical resistivity,
which changes upon application of stress or strain and monitoring this change in
resistivity the stresses in the material can be predicted.

7. Both the electrical conductivity and self-monitoring properties of CFRPC system


was observed at fiber loading of ~3% by weight or higher. Increase in fiber
loading to 6% by weight improved the electrical conductivity but deteriorated the
self-monitoring properties.

8. The change in resistivity response with stress was found to be different under
uniaxial compression and under uniaxial tension. Under uniaxial compression
resistivity initially decreased with increasing stress but after a certain threshold
value, it increased with stress up to failure. On the other hand, under uniaxial

161

tension resistivity continuously increased with increasing tensile stress up to


failure. Also, the stress and strain sensing ability of CFRPC system under uniaxial
tension was found to be much higher than those under uniaxial compression,
which were evident from piezoresistivity coefficients and gage factors
respectively.

9. Both pizoresistivity coefficient and gage factor as well as percentage change in


resistivity increased rapidly at stress level above 70% of failure strength of the
CFRPC system which suggested fracture in the material.

10. Percolation theory predicted the resistivity of CFRPC system at unstressed


condition reasonably well. The value of the conductivity exponent was found to
be in good agreement with the universally accepted value for 3-D insulatorconductor multiphase solids.

11. A general 3-D analytical model was developed to explain the change in resistivity
of CFRPC system with stress tensor and later simplified for uniaxial compression
and uniaxial tension. It has been demonstrated that the piezoresistivity of CFRPC
system arises due to two factors: the change in volume concentration of
conducting carbon fibers with stress and the change in micro-structure of the
heterogeneous CFRPC with stress. The former is a function of the both the
hydrostatic and the deviatoric component of the stress tensor and the later was
assumed to be purely a function of deviatoric stress. The model predicted the

162

change in resistivity with stress reasonably well up to 70% the failure stress of
CFRPC systems for both uniaxial compression and uniaxial tension. The model
under predicted the change in resistivity at higher stress level. This might because
of formation of cracks in the material at higher stress level.

12. Parametric study was performed to understand the effect of individual parameter
of the model on the piezoresistive behavior of the CFRPC system. From
parametric study it was concluded that fiber Youngs modulus and fiber Poissons
ratio have significant effect on the compressive piezoresistive behavior, though
they do not have much influence on the tensile piezoresistive behavior. High
Youngs modulus of fiber increased both the threshold change in resistivity and
the piezoresistivity coefficient. Similarly, low Poissons ratio of fiber increased
both the threshold change in resistivity and the initial piezoresistivity coefficient.
Sensitivity of CFRPC system in both compression and tension decreased with
increase in fiber weight fraction as it reduced the piezoresistivity coefficient. Also
higher value of Poissons ratio of the composite reduced the self-monitoring
sensitivity of the composite.

13. CFRPC beam under four-point loading and PVC-CFRPC-PVC sandwich pipe
configuration under parallel plate loading showed that the applied load or the
deflection due to that load can be predicted by monitoring the change in resistance
behavior between any two points inside those structural elements. In other words
the structural element made out of piezoresistive CFRPC system is itself a bulk

163

sensor and there is no need to embed or glue external load cells or strain gages to
monitor its behavior.

14. Piezoresistance coefficient and gage factor was to characterize the bulk sensing
property of the structural element, CFRPC beam and PVC-CFRPC-PVC
sandwich ring. It should be noted that in bending the piezoresistanace coefficients
and gage factors were not pure material property. These also depended upon the
dimensions of the structural elements.

15. As observed in uniaxial compression and uniaxial tension, in beam bending also
the self-monitoring sensitivity reduced with increase in fiber content. Also it was
observed that the trend of change in resistance behavior with load was different
for 3% CFRPC beam as compared to 6% CFRPC beam. The resistance of 3%
CFRPC beam initially decreased and then after a threshold change in resistance it
increased with increasing load, resembling the uniaxial compressive behavior but
the resistance of 6% CFRPC beam always increased with increasing load,
resembling the uniaxial tensile behavior. Analytical modeling showed that this
difference in trend was because of different shear parameter, Z of 3% CFRPC and
6% CFRPC system.

16. In beam bending, similar to uniaxial compression and uniaxial tension, the selfmonitoring sensitivity was higher at higher load. Both the piezoresistance
coefficient and gage factor increased non-linearly with increase in load, with the

164

rate of change of these coefficients were higher for the 3% CFRPC beam as
compared to the 6% CFRPC beam.

17. In sandwich pipe configuration also the self-monitoring sensitivity was observed
to be higher at higher load. Piezoresistance coefficient increased non-linearly and
gage factor increased linearly with load.

18. A series-parallel resistors idealization of beam predicted the change in resistance


behavior with load for both 3% CFRPC beam and 6% CFRPC beam reasonably
well.

7.2 Recommendations
Based on the results of experimental and analytical investigation, the following
suggestions are recommended for future work:

1. From the present study, it was known that the glass fibers improved both the
compressive and tensile strength of PC systems but carbon fibers did not improve
the compressive strength, though it improved the tensile strength of PC system. It
is suggested that the microstructures of both the glass and carbon fibers be studied
to investigate the difference in behavior of both the fibers.

2. The result of change in resistance behavior of piezoresistive CFRPC with load


was very promising but cyclic load test showed that if the material is loaded to a

165

higher stress in the first cycle compared to the subsequent cycles, only then the
subsequent cycles are repeatable. Repeatability of results is very important before
using the material as a bulk sensor and this might be achieved by passing a
relatively higher voltage of current through the material before loading and thus
causing internal dielectric breakdown.

3. In the present study the self-monitoring behavior of piezoresistive CFRPC was


characterized with the change in resistivity along the stress axis only. It is
suggested that the material should also be characterized with change in resistivity
perpendicular to the stress axis and thus verifying the 3-D analytical model
proposed.

4. In modeling the change in resistance behavior analytically, it was assumed that


the critical volume fraction of conducting fibers (c) is a function of square of
deviatoric stress. It is suggested to simulate the critical volume fraction
numerically using Monte Carlo technique to verify the assumption.

5. The values of shear parameter of CFRPC and carbon fiber were obtained by
curve-fitting the change in resistivity behavior under uniaxial behavior and later
used in predicting the change in resistivity behavior under uniaxial tension and
beam bending. It is suggested to verify the parameters experimentally.

166

6. All the experiments in this investigation were carried out at room temperature and
hence temperature effect was neglected. But since temperature also induces
internal stresses and thus affects the change in resistivity behavior, it is suggested
that the effect of temperature should also be investigated before using the material
as a sensor.

167

REFERENCES
1. ACI Committee 548, Guide for the use of Polymers in Concrete
2. Adolf, D. and Martin, J.E., Calculation of Stresses in Crosslinking Polymers,
Journal of Composite Materials, V.30, No.1, 1996, pp.13-34.
3. Amin, A., Piezoresistivity in Ruthenium-Based Metal-Insulator-Metal Structures,
Journal of Materials Research, V.16, No.8, August 2001, pp.2239-2243.
4. ASTM C 39M-01, Test Method for Compressive strength of Cylindrical Concrete
Specimens, Vol. 04.02.
5. ASTM C 215-97e1, Standard Test Method for Fundamental Transverse,
Longitudinal and Torsional Frequencies of Concrete Specimens, Vol. 04.02.
6. ASTM C 597, Standard Test Method for Pulse Velocity through Concrete, Vol.
04.02.
7. ASTM C 469-94e1, Standard Test Method for Static Modulus of Elasticity and
Poissons Ratio of Concrete in Compression, Vol. 04.02.
8. ASTM C 617-98, Standard Practice for Capping Cylindrical Concrete
Specimens, Vol. 04.02.
9. Bay, J.A. and Stokoe, K.H. II, Field and Laboratory Determination of Elastic
Properties

of

Portland

Cement

Concrete

Using

Seismic

Techniques,

Transportation Research Record 1355.


10. Becerra, R.T., Polymer Concrete for Electrical Applications, SP-69, American
Concrete Institute, Detroit, 1981, pp.145-153.
11. Beek, L.K.H. et al., Journal of Applied Polymer Science, V.6, No.24, 1962, pp.651.

168

12. Bloomfield T.D., Sewers and Manholes with Polymer Concrete, Proceedings of
the 1997 ASCE Conference on Trenchless Pipeline Project, June 1997, pp.466-472.
13. Bontea, D. M.; Chung, D.D.L. and Lee, G.C., Damage in Carbon FiberReinforced Concrete, Monitored by Electrical Resistance Measurement, Cement
and Concrete Research, V.30, No.1, 2000, pp.651-659.
14. Bush, A.W., Contact Mechanics, in Rough Surfaces Edited By Thomas, T.R.,
Longman, London, 1982.
15. Carcia, P.F.; Suna, A. and Childers, W.D., Electrical Conduction and Strain
Sensitivity in RuO2 Thick Film Resistors, Journal of Applied Physics, V.54,
No.10, October 1983, pp.6002-6008.
16. Caroma, F.; Canet, R. and Delhaes, P., Piezoresistivity of Heterogeneous Solids,
Journal of Applied Physics, V.61, No.7, April 1987, pp.2550-2557.
17. Chen, P.W. and Chung, D.D.L., Carbon Fiber Reinforced Concrete as an
Intrinsically Smart Concrete for Damage Assesment during Static and Dynamic
Loading, ACI Materials Journal, V.93, No.4, July-August 1996, pp.341-350.
18. Chen, P.W. and Chung, D.D.L., Carbon Fiber Reinforced Concrete for Smart
Structures capable of Non-Destructive Flaw Detection, Smart Material and
Structure, V.2, No.1, 1993, pp.22-30.
19. Cho, J.W. and Choi J.S., Relationship Between Electrical Resistance and Strain of
Carbon Fibers upon Loading, Journal of Applied Polymer Science, V.77, 2000,
pp.2082-2087.
20. Chung, D.D.L., Structural Health Monitoring by Electrical Resistance
Measurement, Smart Material and Structure, V.10, No.1, 2001, pp.624-636.

169

21. Chung, D.D.L., Strain Sensors based on the Electrical Resistance Change
accompanying the Reversible Pull-Out of Conducting Short Fibers in a less
Conducting Matrix, V.4, No.1, 1995, pp.59-61.
22. Chung, D.D.L., Cement Reinforced with Short Carbon Fibers: A Multifunctional
Material, Composite Part B: Engineering, V.31, No.1, 2000, pp.511-526.
23. CIGMAT PC 1-00, Standard Practice for Making and Curing Polymer Concrete
Test Specimens in Laboratory.
24. CIGMAT PC 2-00, Standard Test Method for Tensile Properties of Polymer
Concrete.
25. CIGMAT PC 5-00, Standard Test Method for Compressive Properties of Polymer
Concrete.
26. Fan, Z., A new approach to the Electrical Resistivity of Two-Phase Composites,
Acta Metal. Mater., V.43, No.1, 1995, pp.43-49.
27. Fontana, J.J. and Bartholomew, J., Use of Concrete Polymer Materials in the
Transportation Industry, Application of Polymer Concrete, SP-69, American
Concrete Institute, Detroit, 1981, pp.21-30.
28. Fowler, D.W., Future Trends in Polymer Concrete, Polymers in Concrete:
Advances and Applications, SP-116, American Concrete Institute, Detroit, 1989,
pp.129-143.
29. Frenkel, J., Physical Review, V.36, 1930, pp.1640.
30. Fu, X. and Chung, D.D.L., Effect of Curing Age on the Self-Monitoring Behavior
of Carbon Fiber Reinforced Mortar, Cement and Concrete Research, V.27, No.9,
1997, pp.1313-1318.

170

31. Fu, X. and Chung, D.D.L., Self-Monitoring of Fatigue Damage in Carbon Fiber
Reinforced Cement, Cement and Concrete Research, V.26, No.1, 1996, pp.15-20.
32. Holm, R., Electrical Contacts Theory and Application, Springer, New York,
1967.
33. Karasek, L.; Meissner, B; Asai, S. and Sumita, M., Percolation Concept: PolymerFiller Gel Formation, Electrical Conductivity and Dynamic Electrical Properties of
Carbon-Black-Filled Rubbers, Polymer Journal, V.28, No.2, 1996, pp.121-126.
34. Kuchaka, L.E., Polymer Concrete Materials for use in Geothermal Energy
Processes, Proceedings of 2nd International Congress on Polymers in Concrete,
University of Texas, Austin, October 1978, pp.157-172.
35. Kuchaka, L.E.; Fontana, J. and Steinberg, M., Polymer Concrete for Repairing
Deteriorated

Bridge

Decks,

Transportation

Research

Record,

No.542,

Transportation Research Board, Washington D.C., 1975, pp.20-28.


36. Lott, J.; Naus, D. and Howdyshell, P., Polymer Concrete-Reinforced Concrete
Composite Beams, SP-40, American Concrete Institute, Detroit, 1973, pp.295318.
37. Leslie, J.R. and Chessman, W.J., An Ultrasonic Method of Studying Deterioration
and Cracking in Concrete Structures, Journal of the American Concrete Institute,
V21, No.1, September, 1949, pp. 17-35.
38. Mantrala, S.K and Vipulanandan, C., Nondestructive Evaluation of Polyester
Polymer Concrete, ACI Materials Journal, V.92, No.6, November-December
1995, pp. 660-668.

171

39. Mantrala, S.K.,

Role of Core Material on Structural Behavior of Sandwich

Pipes, PhD Thesis, University of Houston, 1996.


40. McLachlan, D.S., Blaszkiewicz, M., Newnham, R.E., Electrical Resistivity of
Composites, Journal of American Ceramic Society, V.73, No.8, 1990, pp.21872203.
41. Ueda, N. and Taya, M., Prediction of Electrical Conductivity of Twodimensionally Misoriented Short Fiber Composites by a Percolation Model,
Journal of Applied Physics, V.60, No.1, July 1986, pp.459-461.
42. Mebarkia, S and Vipulanandan, C.,

Compressive Behavior of Glass Fiber

Reinforced Polymer Concrete, Journal of Materials in Civil Engineering, V.4,


No.1, Feb. 1992, pp. 91-105.
43. Mebarkia, S., Mechanical and Fracture Properties of High Strength Polymer
Concrete Under Various Loading Conditions and Corrosive Environments, PhD
Thesis, University of Houston, 1993.
44. Obert, L. and Duvall, W.I, Discussion of Dynamic Methods of Testing Concrete
with Suggestions for Standardization, Proceedings of ASTM, V.41, 1941, pp.
1053-1070.
45. Ohama, Y and Nishimura, T., Properties of Steel Fiber Reinforced Polyester
Resin Concrete, In Proceedings 22nd. Congress on Material Research (Society of
Material Sciences, Kyoto, Japan, 1979), pp. 364-367.
46. Picket, G.,

Equations for Computing Elastic Constants from Flexural and

Torsional Resonant Frequencies of Vibration of Prisms and Cylinders,

172

Proceedings, ASTEA, American Society for Testing and Materials, V.45, 1945, pp.
846-865.
47. Prusinski, R.C., Study of Commercial Development in Precast Polymer concrete,
Polymers in Concrete: International Symposium, SP-58, American Concrete
Institute, Detroit, 1978, pp.75-101.
48. Rejon, L.; Rosas-Zavala, A; Porcayo-Calderon, J. and Cstano, V.M., Percolation
Phenomena in Carbon Black-Filled Polymeric Concrete, Polymer Engineering
and Science, V.40, No.9, September 2000, pp.2101-2104.
49. Reza, F.; Baston, G.B.; Yamamuro, J.A. and Lee, J.S., Volume Electrical
Resistivity of Carbon Fiber Cement Composites, ACI Materials Journal, V.98,
No.1, January-February 2001, pp.25-35.
50. Shui, X. and Chung, D.D.L., Improved Composite Piezoresistive Strain Sensors,
Proceedings of SPIE, V.2716, No.1, 1996, pp.251-258.
51. Shui, X. and Chung, D.D.L., A Piezoresistive Carbon Filament Polymer-Matrix
Composite Strain sensor, Smart Material and Structure, V.5, No.1, 1996, pp.243246.
52. Strumpler, R., Glatz-Reicnenbach, J., Conducting Polymer Composites, Journal
of Electroceramics, V.3, No.4, 1999, pp.329-346.
53. Taya, M., Kim, W.J., Piezoresistivity of Short Fiber Elastomer Composite,
Proceedings of US-Japan Workshop on Smart Materials and Structures, 1997,
pp.243-250.
54. Taya, M., Kim, W.J. and Ono, K., Piezoresistivity of a Short Fiber/Elastomer
Matrix Composite, Mechanics of Materials, V.28, 1998, pp.53-59.

173

55. Taya, M. and Ueda N., Prediction of In-Plane Electrical Conductivity of a


Misoriented Short Fiber Composite: Fiber Percolation Model Versus Effective
Medium Theory, Journal of Engineering Materials and Technology, V.109, July
1987, pp.252-256.
56. Vipulanandan, C. and Dharmarajan N., Flexural Behavior of Polyester Polymer
Concrete, Cement and Concrete Research, V.17, 1987, pp. 219-230.
57. Vipulanandan, C. and Paul, E, Characterization of Polyester Polymer and Polymer
Concrete, Journal of Materials in Civil Engineering, V.5, No.1, Feb. 1993, pp. 6282.
58. Vipulanandan, C and Paul, E., Performance of Epoxy and Polyester Polymer
Concrete, ACI Materials Journal, V.87, No.3, May-June 1990, pp. 241-251.
59. Wang, S., Chung, D.D.L., Apparent Negative Electrical Resistance in Carbon
Fiber Composites, Composites: Part B, V.30, 1990, pp. 579-590.
60. Wang, X., Wang, S. and Chung, D.D.L., Sensing Damage in Carbon Fiber and its
Polymer-Matrix and Carbon-Matrix Composites by Electrical Resistance
Measurement, Journal of Material Science, V.34, 1999, pp.2703-2713.
61. Weber, M. and Kamal M.R., Estimation of the volume Resistivity of electrically
Conductive Composites, Polymer Composites, V.18, No.6, December 1997,
pp.711-725.
62. Wen, S. and Chung, D.D.L., Uniaxial Tension in Carbon Fiber Reinforced
Cement, sensed by Electrical Resistivity Measurement in Longitudinal and
Transverse Directions, Cement and Concrete Research, V.30, no.1, 2000,
pp.1289-1294.

174

APPENDIX

1. CIGMAT PC 1-02: Standard Practice for Making and Curing polymer Concrete
Test Specimens With and Without Fibers in the Laboratory.
2. CIGMAT PC 2-02: Standard Test Method for Tensile Properties of Polymer
Concrete.
3. CIGMAT PC 4-02: Standard Test Method for Flexural Properties of Polymer
Concrete.
4. CIGMAT PC 5-02: Standard Test Method for Compressive Strength Stress-Strain
Relationship of Cylindrical Polymer Concrete Specimens.
5. Operating Instructions for Dynamic Analyzer.

175

CIGMAT TESTING STANDARD


CIGMAT PC 1-02

Standard Practice for Making and


Curing Polymer Concrete Test Specimens
With and Without Fibers in the
Laboratory
1994

C I G MAT

Prepared by
C.Vipulanandan and Kallol Sett

Center for Innovative Grout Materials and Technology (CIGMAT)


Department of Civil and Environmental Engineering
University of Houston
Houston, Texas 77204-4003

May 2003

Modified 5/20/2003

Designation: CIGMAT PC 1-02

Standard Practice for Making and Curing Polymer Concrete Test


Specimens With and Without Fibers in the Laboratory
1. Scope
1.1 This practice covers procedures for making and curing test specimens of polymer
concrete in the laboratory. The polymer concrete can be consolidated by tamping as
described herein.
1.2 The values stated in inch-pound units are to be regarded as the standard. The values
given in parentheses are for information purposes only.
1.3 This standard does not purport to address the safety problems associated with its use.
It is the responsibility of the user of this standard to establish appropriate safety and
health practices and determine the applicability of regulatory limitations prior to use.
2. Referenced Documents
ASTM C 192-Standard Practice for Making and Curing Concrete Test Specimens in the
Laboratory.
ASTM C 125 -Standard Terminology Relating to Concrete and Concrete Aggregates.
ASTM C127-Standard Test Method for Specific Gravity and Absorption of Coarse
Aggregate.
ASTM C 128-Standard Test Method for Specific Gravity and Absorption of Fine
Aggregate.
ASTM C 138-Standard Test Method for Unit Weight, Yield, and Air Content
(Gravimetric) of Concrete.
ASTM C 330-Standard Specification for Lightweight Aggregates for Structural
Concrete.
ASTM C 470-Standard Specification for Molds for Forming Concrete Test Cylinders
Vertically.
ASTM C 511- Standard Specification for Moist Cabinets, Moist Rooms, and Water
Storage Tanks Used in the Testing of Hydraulic Cements and Concretes.
ASTM C 617-Standard Practice for Capping Cylindrical Concrete Specimens.
ASTM C 1064-Standard Test Method for Temperature of Freshly Mixed PortlandCement Concrete.
ASTM C 1312-Standard Practice for Making and Conditioning Chemical-Resistant
Sulfur Polymer Cement Concrete Test Specimens in the Laboratory.
ASTM D448-Standard Classification for Sizes of Aggregate for Road and Bridge
Construction.
3. Apparatus
3.1 Molds, GeneralMolds for specimens or fastenings thereto in contact with the
polymer concrete shall be made of Teflon. Molds shall conform to the dimensions and
tolerances specified in the method for which the specimens are required. Inner surfaces of

Modified 5/20/2003

the molds shall be lightly coated with mineral oil or a suitable no reactive release
material before use.
3.2 Cylinder Molds:
3.2.1 Molds for Casting Specimens shall conform to the above requirement 3.1 and
ASTM C 470.
3.3 Beam and Prism Molds:
3.3.1Beam and Prism Molds shall be rectangular in shape (unless otherwise specified)
and of the dimensions required to produce the desired specimen size. The inside surfaces
of the molds shall be smooth and free from indentations. The sides, bottom, and ends
shall be at right angles to each other and shall be straight and true and free of warpage.
Maximum variation from the nominal cross section shall not exceed 1/8 in. (3.2 mm) for
molds with depth or breadth of 6 in. (152 mm) or more, or 1/16 in. (1.6 mm) for molds of
smaller depth or breadth. Except for flexure specimens, molds shall not vary from the
nominal length by more than l/16 in. Flexure molds shall not be shorter than 1/16 in. of
the required length, but may exceed it by more than that amount.
3.4 Tamping RodsTwo sizes are specified in ASTM standards. Each shall be a round,
straight steel rod with at least the tamping end rounded to a hemispherical tip of the same
diameter as the rod. Both ends may be rounded, if preferred.
3.4.1 Larger Rod shall be 5/8 in. (16 mm) in diameter and approximately 24 in. (610 mm)
long.
3.4.2 Smaller Rod shall be 3/8 in. (10 mm) in diameter and approximately 12 in. (305
mm) long.
3.5 Small ToolsTools and items such as shovels trowels, wood float, blunted trowels,
rulers, rubber gloves, and mixing bowls shall be provided.
3.6 Sampling and Mixing ContainerThe container shall be flat bottom and of heavygage paper, watertight, of convenient depth and of sufficient capacity to allow easy
mixing by shovel or trowel of the entire batch.
3.7 ScalesScales for weighing batches of materials and polymer concrete shall be
accurate within 0.3 % of the test load at any point within the range of use. They shall
meet the requirements for sensitivity and tolerances prescribed by the National Institute
of Standards and Technology. Where the scales are graduated in decimal fractions of a
pound instead of ounces, or where the metric system is used, the equivalent percentage
sensitivity requirement and tolerances shall apply.
4. Specimens
4.1 Cylindrical SpecimensCylinders for such tests as compressive strength, Young's
modulus of elasticity, creep, and splitting tensile strength may be of various sizes with a
minimum of 1.5-in. diameter by 3-in. length or 2.5 by 5 in.
4.2 Prismatic SpecimensBeams for flexural strength, cubes for compressive strength,
prisms for freezing and thawing, length change, and volume change shall be formed with
their long axes horizontal, unless otherwise required by the method of test in question,
and shall conform in dimension to the requirements of the specific test method.
4.3 Other SpecimensOther shapes and sizes of specimens for particular tests may be
molded as desired following the general procedures set forth in this practice.

Modified 5/20/2003

4.4 Number of SpecimensThe number of specimens and the number of test batches are
dependent on established practice and the nature of the test program. Guidance is usually
given in the test method or specification for which the specimens are made. Generally
three or more specimens shall be made for each test age and test condition unless
otherwise specified. Specimens involving a given variable shall be made from three
separate batches mixed on different days. An equal number of specimens for each
variable shall be made on any given day. When it is impossible to make at least one
specimen for each variable on a given day, the mixing of the entire series of specimens
shall be completed in least possible time and one of the mixtures shall be repeated each
day as a standard of comparison.
5. Preparation of Materials
5.1 TemperatureBefore mixing the polymer concrete the materials shall be brought to
room temperature in the range of 68 to 86F (20 to 30C).
5.2 Resin, cobalt napthenate, methyl ethyl ketone peroxide The resin shall be stored in
a dry place, in moisture-proof containers, preferably made of metal.
5.3 AggregatesThe sand shall be constituted by mixing five grades of commercially
available blasting sand of equal weight. It shall be well graded and shall have a
coefficient of concavity (Cc) of 0.9 and a coefficient of uniformity (Cu) of 5.8.
5.4 The compositions of the polymer concrete (PC), glass fiber reinforced polymer
concrete (GFRPC) and carbon fiber reinforced polymer concrete (CFRPC) are
recommended in Table 1, Table 2 and Table 3 respectively. Earlier studies have shown
that these are the mix proportions of PC that has the highest strength and modulus.
Table 1 Compositions of PC
Constituent Materials
Polyester resin
MEKPO, Initiator
Cobalt napthenate, Promoter
Blasting sand
a
by weight of resin

Relative Proportions by weight (%)


14
1.5a
0.3a
86

Table 2 Compositions of GFRPC


Constituent Materials
Polyester resin
MEKPO, Initiator
Cobalt napthenate, Promoter
Blasting sand
Glass Fibers
a
by weight of resin
b
by weight of resin+polymer

Relative Proportions by weight (%)


18
1.5a
0.3a
82
6b

Modified 5/20/2003

Table 3 Compositions of CFRPC


Constituent Materials
Polyester resin
MEKPO, Initiator
Cobalt napthenate, Promoter
Blasting sand
Carbon Fibers
a
by weight of resin
b
by weight of resin+polymer

Relative Proportions by weight (%)


20
1.5a
0.3a
80
6b

6. Procedure
6.1 In preparing PC specimens cobalt napthenate shall be added to the polyester resin and
the solution shall be mixed for at least 2 minutes before adding methyl ethyl ketone
peroxide. After mixing, the sand or the sand-fiber mixture shall be slowly added to the
polyester resin and shall be mixed long enough to obtain a uniform mixture.
6. 2 Number of Layers The mixture shall then be poured slowly into the Teflon molds
to cast the specimens. Specimens shall be made in three layers with proper tamping.
6.3 TampingEach layer shall be tamped with the rounded end of a rod. Bottom layer
shall be tamped throughout its depth. Strokes shall be distributed uniformly over the
cross section of the mold and for each upper layer the rod shall be allowed to penetrate
about 1/2 in. into the underlying layer when the depth of the layer is less than 4 in. and
about 1 in. when the depth is 4 in. or more. After tamping, the outside face of the mold
shall be tapped lightly for 10 to 15 times with the mallet to close any holes left by
tamping and to release any large air bubbles that may have been trapped. Hand can be
used to tap light-gage single-use molds, which are susceptible to damage if tapped with a
mallet. After tapping, the polymer concrete shall be spade along the sides and ends of
beam and prism molds with a trowel or other suitable tool to smoothen the outside
surface.
7. Curing
7.1 Removal from MoldsSpecimens shall be removed from the molds after 24 hours
after casting.
7.2 Curing Environment The specimens shall be allowed to cure for one day at room
temperature followed by at 800 C for 24 hours.
7.3 Before testing, the ends of the cylindrical specimens shall be trimmed using a
diamond saw to ensure smooth and parallel surface.

This standard is subject to revision at any time by the responsible technical committee and subject to review every five years and if not
revised, either re-approved or withdrawn. Your comments are invited either for revision of this standard or for additional standards and
should be addressed to CIGMAT Headquarters. Your comments will receive careful consideration at a meeting of the responsible
technical committee, which you may attend. If you feel that your comments have not received a fair hearing you should make your
views known to the CIGMAT Committee on Standards, 4800 Calhoun, Houston, TX77204-4791.

Modified 5/20/2003

CIGMAT TESTING STANDARD


CIGMAT PC 2-02

Standard Test Method for Tensile


Properties of Polymer Concrete

1994

C I G MAT

Prepared by
C.Vipulanandan and Kallol Sett

Center for Innovative Grout Materials and Technology (CIGMAT)


Department of Civil and Environmental Engineering
University of Houston
Houston, Texas 77204-4003

May 2003

Modified 5/20/2003

Designation: CIGMAT PC 2-02

Standard Test Method for Tensile Properties of Polymer Concrete


1. Scope
1.1 This method covers the determination of the tensile strength of polymer concrete
specimens by the use of simple dog bone specimens.
2. Referenced Documents
ASTM C 78-Standard Test Method for Flexural Strength of Concrete (Using Simple
Beam with Third-Point Loading).
ASTM C 192-Standard Practice for Making and Curing Concrete Test Specimens in the
Laboratory.
ASTM D 638-Standard Test Method for Tensile Properties of Plastics.
ASTM D 3574-Standard Test Method for Flexible Cellular Materials.
3. Apparatus
3.1 Testing Machine The machine shall be capable of applying loads at a uniform rate
without shock or interruption.
3.2 Loading Apparatus It shall ensure that all forces are applied perpendicular to the
face of the specimen without eccentricity. Reactions shall be parallel to the direction of
applied load at all times during the test.
4. Test Specimen
Dog bone shaped specimens as shown in Fig.1 shall be used. The sides of the specimen
shall be at right angles with the top and bottom. All surfaces shall be smooth and free of
scars, indentations, holes, or inscribed identifications.

A
1.250

A'
0.960
1.925

0.660

3.000

section
A-A'

7.720

Figure1-Geometry of tensile test specimens (inches)


1

Modified 5/20/2003

5. Procedure
5.1.Tensile testing shall be performed using a screw type mechanical testing system.
Loading shall be continuous and without shock. The crosshead speed shall be 1.0
mm/min. During tensile loading up to fracture, the strain shall be measured by a strain
gage attached at middle height of the specimens.
6. Calculations
Tensile strength shall be calculated by dividing the maximum load carried by the
specimen during the test by the average cross-sectional area and the result shall be
expressed to the nearest psi (6.9 kPa).

This standard is subject to revision at any time by the responsible technical committee and will be reviewed regularly and if not
revised, either re-approved or withdrawn. Your comments are invited either for revision of this standard or for additional standards
and should be addressed to CIGMAT Headquarters. Your comments will receive careful consideration at a meeting of the responsible
technical committee. If you feel that your comments have not received a fair hearing you should make your views known to the
CIGMAT Committee on Standards, 4800 Calhoun., Houston, TX77204-4003.

Modified 5/20/2003

CIGMAT TESTING STANDARD


CIGMAT PC 4-02

Standard Test Method for Flexural


Properties of Polymer concrete
1994

C I G MAT

Prepared by
C.Vipulanandan and Kallol Sett

Center for Innovative Grout Materials and Technology (CIGMAT)


Department of Civil and Environmental Engineering
University of Houston
Houston, Texas 77204-4003

May 2003

Modified 5/20/2003

Designation: CIGMAT PC 4-02

Standard Test Method for Flexural Properties of Polymer concrete


1. Scope
1.1.1
1.1.2
1.1.3
1.1.4
1.1.5

This test method covers the determination of flexural properties of polymer


concrete.
The values stated in meter-newton units shall be regarded as the standard.
A four-point loading system-utilizing center loading on a simply supported
beam shall be adopted.
This standard may involve hazardous materials, operations, and equipment.
This standard does not purport to address all of the safety problems associated
with its use. It is the responsibility of whoever uses this standard to consult
and establish appropriate safety and health practices and determine the
applicability of regulatory limitations prior to use.

2. Reference Documents
ASTM C 31M-Standard Practice for Making and Curing Concrete Test Specimens in
the Field
ASTM C 192M-Standard Practice for Making and Curing Concrete Test Specimens
in the laboratory
ASTM C 293-Standard Testing Method for Flexural Strength of Concrete
ASTM D 790-Standard Test Method for flexural properties of un-reinforced and
reinforced plastics and electrical insulating materials.
3. Summary of Test Method
3.1.1
3.1.2
3.1.3

A beam specimen shall be tested in flexure.


The beam shall be rested on two supports and shall be loaded by means of two
loading noses between the supports.
The specimen shall be loaded until rupture occurs.

4. Significance and Use


4.1
4.2

Flexural properties are useful for quality control and material specifications.
The flexural properties may vary with specimen depth, temperature,
atmospheric conditions and the difference in the rate of straining.

Modified 5/20/2003

5. Apparatus
5.1.1

5.1.2

Testing machine---A properly calibrated testing machine that can be operated


at constant rates of cross head motion over the range indicated, and in which
the error in the load measuring system shall not exceed 1% either way of the
maximum load expected to be measured. It shall be equipped with a
deflection-measuring device.
CIGMAT Testing Machine: INSTRON MODEL 1322

6. Test specimen
The recommended specimen for molding materials is 50 mm by 50 mm by 225 mm
tested on a support span of 170 mm.
7. Number of test specimens
At least two specimens shall be tested for each sample.
8. Conditioning
ConditioningTest specimen shall be conditioned at 23+2 o C and 50+5%
relative humidity for not less than 40 hr prior to test.
8.1.2 Test conditionsTest shall be conducted in standard laboratory atmosphere
of 23+2 o C and 50+5% relative humidity.
8.1.1

9. Procedure
9.1.1 The support span range shall be 170 mm in four point bending test.
9.1.2 Rate of crosshead motion shall be determined and the machine shall be set for
the calculated rate.
CIGMAT Recommendations:
Use a crosshead speed of 0.0635 mm/min
The machine settings to reach the desired speed are:
INPUT
1.0
9.1.3

9.1.4

RANGE
0.01

MULTIPLIER
0.02

Loading noses and the supports shall be aligned in such a way that the axes of
the cylindrical surfaces are parallel and the loading noses are symmetrical.
Specimen shall be centered on the supports, with the long axis of the specimen
perpendicular to the loading noses and the supports.
Specimen shall be loaded at the specified crosshead speed. Load and
deflection data shall be taken simultaneously. Deflection shall be measured
either by LVDT or by machine movements or by strain gages.

Modified 5/20/2003

9.1.5
9.1.6
9.1.7
9.1.8

Load-deflection curves may be plotted to determine the flexural strength,


secant or tangent modulus of elasticity and the total work measured by the
area under the load-deflection curve.
Test shall be terminated when specimen ruptures.
The deflection at which this strain occurs may be calculated as follows:
D=

0.21rL2
d

(1)

where:
D= deflection at midspan, mm
r= strain, mm/mm
L= support span, mm
d= height of beam , mm
10. Retest:
Material properties at rupture shall not be calculated for any specimen that
breaks at some obvious, fortuitous flaw, unless such flaw constitutes a
variable being studied. Retests shall be made for any specimens on which
values are not calculated.
11

Calculations:

11.1

Maximum fiber stress


When a beam of homogenous, elastic material is tested in four-point flexure
as a simple supported at two points, the maximum stress in the outer fibers
occurs at the midspan. This stress shall be calculated for any point on the
load-deflection curve by the following equation:

= PL bd 2

(2)

where,
= stress in the outer fibers at mid-span, MPa
P = load at a given point on the load-deflection curve, N,
L = support span, mm
B =width of the beam, tested, mm and
d = height of the beam tested, mm.
Note: The above equation is applicable for which the stress is linearly
proportional to the strain upto the point of rupture and for which the
strains are small.

Modified 5/20/2003

11.2 Flexural Strength (Modulus of rupture)


The flexural strength shall be obtained as the maximum stress in the outer
fiber at rupture. It shall be calculated in accordance to Eq (2) by equating P to
the load at rupture.
11.3 Stress at any given point
The stress at any given strain may be calculated in accordance with Eq (2)
by equating P to the load read from the load-deflection curve at the
deflection corresponding to the desired strain.
11.4 Maximum strain
The maximum strain in the outer fibers occurs at the midspan, and may be
calculated as follows:

r=

dD
0.21L2

(3)

where:
r = maximum strain in the outer fibers, mm/mm
D = maximum deflection of the center of the beam, mm,
L = support span,mm
d =depth, mm
11.5 Tangent Modulus of Elasticity

The tangent modulus of elasticity, often called, as the modulus of


elasticity is the ratio, within the elastic limits of the stress to the
corresponding strain and shall be expressed in MPa. It shall be calculated
by drawing a tangent to the steepest initial straight-line portion of the loaddeflection curve.
CIGMAT Recommendations:
The value of E was determined
graphically by dividing the stress in the elastic region by the
corresponding strain.
11.6 Secant Modulus of Elasticity

The secant modulus of elasticity is the ratio of the stress to the


corresponding strain at any given point on the stress-strain curve, or the
slope of the straight line that joins the origin and a selected point on the
actual stress or strain.

Modified 5/20/2003

CIGMAT Recommendations:
The value of E was determined
graphically by dividing the stress in the elastic region by the
corresponding strain.

11.7 Arithmetic Mean

For each series of tests, the arithmetic mean of all values obtained shall be
calculated to three significant figures and reported as the average value
for the particular property in question.
11.8 Standard deviation

The standard deviation shall be calculated as follows and shall be reported


to two significant figures:
s=

nX

n 1

where:
s= estimated standard deviation
X= value of a single observation
N = number of observations and
X = Arithmetic mean of the set of observations.
11.9 Coefficient of variance

Coefficient of variance shall be calculated as follows and reported to two


significant figures: COV= standard deviation/ mean

Modified 5/20/2003

APPENDIX

L/3
specimen

support

L=170 mm

Load Diagram
SAMPLE CALCULATIONS:
Readings recorded:
Specimen dimensions:

Support span
L
(mm)
170

Breadth
b
(mm)
50

Depth
d
(mm)
50

From Instron Machine:

Load reading shall be taken from voltmeter (in volts) and the deflection reading from
LVDT (in volts)
Reading

Load
LVDT
Load Deflection
recorded Recorded difference
Stress
(V)
(V)
(N)
(mm)
(MPa)
Initial
-0.23
13.19
0
0
0
At failure
-0.4
12.9
189.125 0.0048433 0.25721

Strain
(mm/mm)
0
0.0003982

Modified 5/20/2003

Load Conversion factor (l.c.f)

Load (lb) = ( load reading/100 ) * 10 *

5
* 50,000
100

Deflection correction factor (d.c.f)

Deflection (mm) = (LVDT reading difference (V) -0.084637)*25.4/107.92


Stress:

Stress (MPa) = load difference (N) * 170/(50 * 50 * 50)


Strain:

Strain (mm/mm) = 50 * deflection (mm)/(0.21 * 170 * 170)

14

12

stress (MPa)

10

0
0

0 .0 0 0 5

0 .0 0 1

0 .0 0 1 5

0 .0 0 2

0 .0 0 2 5

st r a in ( m m / m m )

Figure - Typical flexural relationship of polymer concrete

This standard is subject to revision at any time by the responsible technical committee and subject to review every five
years and if not revised, either re-approved or withdrawn. Your comments are invited either for revision of this standard or
for additional standards and should be addressed to CIGMAT Headquarters. Your comments will receive careful
consideration at a meeting of the responsible technical committee, which you may attend. If you feel that your comments
have not received a fair hearing you should make your views known to the CIGMAT Committee on Standards, 4800
Calhoun, Houston, TX77204-4791.

Modified 5/20/2003

CIGMAT TESTING STANDARD


CIGMAT PC 5-02

Standard Test Method for Compressive


Strength and stress-strain relationship of
Cylindrical Polymer Concrete Specimens
1994

C I G MAT

Prepared by
C.Vipulanandan and Kallol Sett

Center for Innovative Grout Materials and Technology (CIGMAT)


Department of Civil and Environmental Engineering
University of Houston
Houston, Texas 77204-4003

May 2003

Modified 5/20/2003

Designation: CIGMAT PC 5-02

Standard Test Method for Compressive Strength and stress-strain


relationship of Cylindrical Polymer Concrete Specimens
1. Scope
1.1 This test method covers determination of compressive strength of cylindrical polymer
concrete specimens.
1.2 The values stated in inch-pound units shall be regarded as the standard.
1.3 This standard may involve hazardous materials, operations, and equipment. This
standard does not purport to address all of the safety problems associated with its use. It
is the responsibility of whoever uses this standard to consult and establish appropriate
safety and health practices and determine the applicability of regulatory limitations prior
to use.
2. Referenced Documents
ASTM C 39-Standard Test Method for Compressive Strength of Cylindrical Concrete
Specimens.
ASTM C 31-Standard Practice for Making and Curing Concrete Test Specimens in the
Field.
ASTM C 192-Standard Practice for Making and Curing Concrete Test Specimens in the
Laboratory.
ASTM C 617-Standard Practice for Capping Cylindrical Concrete Specimens
ASTM C 873-Standard Test Method for Compressive Strength of Concrete Cylinders
Cast in Place in Cylindrical Molds.
ASTM E 4-Standard Practices for Load Verification of Testing Machines.
ASTM E 74-Standard Practice for Calibration of Force-Measuring Instruments for
Verifying the Load Indication of Testing Machines.
3. Summary of Method
3.1 This test method consists of applying a compressive axial load to molded cylinders at
a rate, which is within a prescribed range until failure occurs. The compressive strength
of the specimen is calculated by dividing the maximum load attained during the test by
the cross-sectional area of the specimen.
4.Significance and Use
4.1 Care must be exercised in the interpretation of the significance of compressive
strength determinations by this test method since strength is not a fundamental or
intrinsic property of concrete made from given materials. Values obtained will depend on
the size and shape of the specimen, batching, mixing procedures, the methods of

Modified 5/20/2003

sampling, molding, and fabrication and the age, temperature, and moisture conditions
during curing.
4.2 This test method may be used to determine compressive strength of cylindrical
specimens prepared and cured in accordance with ASTM Methods C 31, C 42, and C
192, Practice C 617, and Test Method C 873.
4.3 The results of this test method may be used as a basis for quality control of polymer
concrete proportioning, mixing, and placing operations; determination of compliance
with specifications; control for evaluating effectiveness of admixtures and similar uses.
5. Apparatus
5.1 Testing MachineThe testing machine shall be of a type having sufficient capacity
and capable of providing the rates of loading prescribed in 7.5. CIGMAT TESTING
MACHINE: 400 kips capacity Tinius Olsen universal testing machine
5.1.1 DesignThe design of the machine must include the following features:
5.1.2 The machine must be power operated and must apply the load continuously rather
than intermittently, and without shock. If it has only one loading rate (meeting the
requirements of 7.5), it must be provided with a supplemental means for loading at a rate
suitable for verification. This supplemental means of loading may be power or hand
operated.
5.1.3 The space provided for test specimens shall be large enough to accommodate, in a
readable position, an elastic calibration device which is of sufficient capacity to cover the
potential loading range of the testing machine.
NOTE 1The type of elastic calibration device most generally available and most
commonly used for this purpose is the circular proving ring.
5.1.4 AccuracyThe accuracy of the testing machine shall be in accordance with the
following provisions:
5.1.4.1 The percentage of error for the loads within the proposed range of use of the
testing machine shall not exceed + 1.0 % of the indicated load.
5.1.4.2 The accuracy of the testing machine shall be verified by applying five test loads in
four approximately equal increments in ascending order. The difference between any two
successive test loads shall not exceed one third of the difference between the maximum
and minimum test loads.
5.1.4.3 The test load as indicated by the testing machine and the applied load computed
from the readings of the verification device shall be recorded at each test point. Calculate
the error, E, and the percentage of error, Ep, for each point from these data as follows:
E=A-B
Ep = 100(A - B)/B
where:
A = load, lbf (or N) indicated by the machine being verified, and
B = applied load, lbf (or N) as determined by the calibrating device.
5.1.4.4 The report on the verification of a testing machine shall state within what loading
range it was found to conform to specification requirements rather than reporting a
blanket acceptance or rejection. In no case shall the loading range be stated as including

Modified 5/20/2003

loads below the value which is 100 times the smallest change of load that can be
estimated on the load-indicating mechanism of the testing machine or loads within that
portion of the range below 10 % of the maximum range capacity.
5.1.4.5 In no case shall the loading range be stated as including loads outside the range of
loads applied during the verification test.
5.1.4.6 The indicated load of a testing machine shall not be corrected either by
calculation or by the use of a calibration diagram to obtain values within the required
permissible variation.
5.2 The testing machine is equipped with two steel bearing blocks with hardened faces
(Note 2), one of which is a spherically seated block that will bear on the upper surface of
the specimen, and the other a solid block on which the specimen shall rest. Bearing faces
of the blocks shall have a minimum dimension at least 3 % greater than the diameter of
the specimen to be tested. Except for the concentric circles described below, the bearing
faces shall not depart from a plane by more than 0.001 in. (0.025 mm) in any 6 in. (152
mm) of blocks 6 in. in diameter or larger, or by more than 0.001 in. in the diameter of any
smaller block; and new blocks shall be manufactured within one half of this tolerance.
When the diameter of the bearing face of the spherically seated block exceeds the
diameter of the specimen by more than 1/2 in. (13 mm), concentric circles not more than
l/32 in. (0.8 mm) deep and not more than 3/64 in. (1.2 mm) wide shall be inscribed to
facilitate proper centering.
NOTE 2It is desirable that the bearing faces of blocks used for compression testing of
concrete have a Rockwell hardness of not less than 55 HRC.
5.2.1 Bottom bearing blocks shall conform to the following requirements:
5.2.1.1 The bottom bearing block is specified for the purpose of providing a readily
machinable surface for maintenance of the specified surface conditions (Note 3). The top
and bottom surfaces shall be parallel to each other. The block may be fastened to the
platen of the testing machine. Its least horizontal dimension shall be at least 3 % greater
than the diameter of the specimen to be tested. Concentric circles as described in 5.2 are
optional on the bottom block.
5.2.1.2 Final centering must be made with reference to the upper spherical block. When
the lower bearing block is used to assist in centering the specimen, the center of the
concentric rings, when provided, or the center of the block itself must be directly below
the center of the spherical head. Provision shall be made on the platen of the machine to
assure such a position.
5.2.1.3 The bottom bearing block shall be at least 1 in (25 mm) thick when new, and at
least 0.9 in. (22.5 mm) thick after any resurfacing operations.
NOTE 3If the testing machine is so designed that the platen itself can be readily
maintained in the specified surface condition, a bottomblock is not required.
5.3 Load Indication:
5.3.1 If the load of a compression machine used in concrete testing is registered on a dial,
the dial shall be provided with a graduated scale that can be read to at least the nearest
0.1% of the full scale load. The dial shall be readable within 1 % of the indicated load at
any given load level within the loading range. In no case shall the loading range of a dial
be considered to include loads below the value that is 100 times the smallest change of
load that can be read on the scale. The scale shall be provided with a graduation line

Modified 5/20/2003

equal to zero and so numbered. The dial pointer shall be of sufficient length to reach the
graduation marks; the width of the end of the pointer shall not exceed the clear distance
between the smallest graduations. Each dial shall be equipped with a zero adjustment that
is easily accessible from the outside of the dial case, and with a suitable device that at all
times until reset, will indicate to within 1 % accuracy the maximum load applied to the
specimen.
5.3.2 If the testing machine load is indicated in digital form, the numerical display must
be large enough to easily read. The numerical increment must be equal to less than 0.10
% of the full-scale load of a given loading range. In no case shall the verified loading
range include loads less than the minimum numerical increment multiplied by 100. The
accuracy of the indicated load must be within 1.0 % for any value displayed within the
verified loading range. Provision must be made for adjusting indicator true zero at zero
load. There shall be provided maximum load indicator at all times until reset.
5.4 Strain Measurement - Extensometer shall be used to measure the axial strain of the
specimen.
6. Specimens
6.1 Specimens shall not be tested if any individual diameter of a cylinder differs from any
other diameter of the same cylinder by more than 2 %.
6.2 Neither end of compressive test specimens when tested shall depart from
perpendicularity to the axis by more than 0.5 (approximately equivalent to 1/8 in. in 12
in. (3 mm in 300 mm)). The ends of compression test specimens that are not plane within
0.002 in. (0.050 mm) shall be capped in accordance with Practice C 617 or they may be
sawed or ground to meet that tolerance. The diameter used for calculating the crosssectional area of the test specimen shall be determined to the nearest 0.01 in. (0.25 mm)
by averaging two diameters measured at right angles to each other at about midheight of
the specimen.
6.3 The number of individual cylinders measured for determination of average diameter
may be reduced to one for each ten specimens or three specimens per day, whichever is
greater, if all cylinders are known to have been made from a single lot of reusable or
single-use molds which consistently produce specimens with average diameters within a
range of 0.02 in. (0.51 mm). When the average diameters do not fall within the range of
0.02 in. or when the cylinders are not made from a single lot of molds, each cylinder
tested must be measured and the value used in calculation of the unit compressive
strength of that specimen. When the diameters are measured at the reduced frequency, the
cross-sectional areas of all cylinders tested on that day shall be computed from the
average of the diameters of the three or more cylinders representing the group tested that
day.
6.4 The length shall be measured to the nearest 0.05D when the length to diameter ratio is
less than 1.8, or more than 2.2, or when the volume of the cylinder is determined from
measured dimensions.

Modified 5/20/2003

7. Conditioning
7.1 ConditioningThe test specimen shall be conditioned at 232C and 505% relative
humidity for not less than 40 hours prior to test.
7.2 Test conditionsThe test shall be conducted in standard laboratory atmosphere of
232C and 505% relative humidity.
8. Procedure
8.1 Placing the SpecimenPlace the plain (lower) bearing block, with its hardened face
up, on the table or platen of the testing machine directly under the spherically seated
(upper) bearing block. Wipe clean the bearing faces of the upper and lower bearing
blocks and of the test specimen and place the test specimen on the lower bearing block.
Carefully align the axis of the specimen with the center of thrust of the spherically seated
block. As the spherically seated block is brought to bear on the specimen, rotate its
movable portion gently by hand so that uniform seating is obtained.
8.2 Rate of LoadingLoad shall be applied continuously and without shock.
8.2.1 For testing machines of the screw type, the moving head shall travel at a rate of
approximately 0.05 in.(1.3 mm)/min when the machine is running idle. For hydraulically
operated machines, the load shall be applied at a rate of movement (platen to crosshead
measurement) corresponding to a loading rate on the specimen within the range of 20 to
50 psi/sec (0.14 to 0.34 MPa/s). The designated rate of movement shall be maintained at
least during the latter half of the anticipated loading phase of the testing cycle.
8.2.2 During the application of the first half of the anticipated loading phase a higher rate
of loading shall be permitted.
8.2.3 No adjustment shall not be made in the rate of movement of the platen at any time
while a specimen is yielding rapidly immediately before failure.
8.3 Load shall be applied at a constant rate until the specimen fails. Load and strain
readings shall be recorded during the process.
9. Calculation
Compressive strength of the specimen shall be calculated by dividing the maximum load
carried by the specimen during the test by the average cross-sectional area determined as
described in Section 6 and express the result to the nearest psi (6.9 kPa).
Sample Calculation:
Specimen dimensions: length = L; diameter = D.
Displacement reading shall be taken from extensometer (in volts) and load reading from
Tinius Olsen machine (in pounds).

Modified 5/20/2003

1st
reading
2nd
reading

load
reading
(lb)
0

extensome load (N)


ter reading
(V)
-140.8
0

displacem
ent (in)

stress
(MPa)

strain (%)

1.995860

300

-141.1

1.995788

1.17713

0.003594

1335

load (N) =load reading*4.45.(1)


length (in) =2.0296+0.00023963*extensometer reading..(2)
displacement (in) =initial length-current length....(3)
stress (MPa) =4*load/(3.1415926*diameter*diameter)(4)
strain (%) =100*displacement/2...(5)

70

Stress (MPa)

60
50
40
30
20
10
0
0

0.5

1
1.5
2
2.5
Strain (%)
Figure - Typical compressive stress-strain relationship of polymer concrete

10. Report
10. 1 The report shall include the following:
10.1.1 Identification number,
10.1.2 Diameter (and length, if outside the range of 1.8D to 2.2D), in inches or
millimetres,
10.1.3 Cross-sectional area, in square inches or square centimetres,

Modified 5/20/2003

10.1.4 Maximum load, in pounds-force or newtons,


10.1.5 Compressive strength calculated to the nearest 10 psi or 69 kPa,
10.1.6 Type of fracture, if other than the usual cone,
10.1.7 Defects in either specimen or caps.

This standard is subject to revision at any time by the responsible technical committee and must be reviewed every five years and if
not revised, either reapproved or withdrawn. Your comments are invited either for revision of this standard or for additional standards
and should be addressed to CIGMAT Headquarters. Your comments will receive careful consideration at a meeting of the responsible
technical committee, which you may attend. If you feel that your comments have not received a fair hearing you should make your
views known to the CIGMAT Committee on Standards, 4800 Calhoun., Houston, TX77204-4791.

Modified 5/20/2003

Amplitude

(a)

accelerometer
support
impact point
0

Frequency (kHz)

12.8

Amplitude

(b)

accelerometer
support
impact point
0

Frequency (kHz)

12.8

Amplitude

(c)

accelerometer
support
impact point
0

Frequency (kHz)

12.8

Figure 1 - Typical frequency spectrum (X-f) for 6% GFRPC cylinders


(a) Longitudinal mode (b) Transverse mode and (c) Torsional mode

Amplitude

(a)

accelerometer
support
impact point
0

Time (millisecond)

31.2195

Amplitude

(b)

accelerometer
support
impact point
0

Time (millisecond)

31.2195

Amplitude

(c)

accelerometer
support
impact point
0

Time (millisecond)

31.2195

Figure 2 Typical time decay (X-t) relationships for 6% GFRPC cylinders


(a) Longitudinal mode (b) Transverse mode and (c) Torsional mode

Figure 3 Typical set up for impact resonance testing

You might also like