You are on page 1of 241

Experimental and Numerical Study on Synthetic Fiber-Reinforced Concrete Pipes

A dissertation presented to

the faculty of

the Russ College of Engineering and Technology of Ohio University

In partial fulfillment

of the requirements for the degree

Doctor of Philosophy

Fouad T. Al Rikabi

August 2020

© 2020 Fouad T.S. Al Rikabi. All Rights Reserved


2

This dissertation titled

Experimental and Numerical Study on Synthetic Fiber-Reinforced Concrete Pipes

by

FOUAD T. AL RIKABI

has been approved for

the Department of Civil Engineering

and the Russ College of Engineering and Technology by

Shad M. Sargand

Russ Professor of Civil Engineering

Mei Wei

Dean, Russ College of Engineering and Technology


3

ABSTRACT

AL RIKABI, FOUAD T., Ph.D., August 2020, Civil Engineering.

Experimental and Numerical Study on Fiber-Reinforced Concrete Pipes

Director of Dissertation: Shad M. Sargand

Synthetic fibers have been recently used in a concrete mixture in an attempt to

produce a new concrete pipe system, cheaper, lighter, and more flexible than

conventional steel reinforced concrete pipes. However, no structural design codes have

been introduced for synthetic fiber reinforced concrete pipes evaluation. Also, there is

little in the literature regarding synthetic fiber applications in the concrete pipes. The

effect of adding two types of synthetic fiber, polypropylene (PP) and polyvinyl alcohol

(PVA) fibers, on the mechanical properties of concrete, including CTE, dynamic modulus

of elasticity, and flexural strength, was investigated. Also, this study focused on the

evaluation of the synthetic fiber reinforced concrete pipes performance in terms of ASTM

requirements for strength, stiffness, and ductility, and developing design tables for

synthetic fiber reinforced concrete pipe similar to those proposed in ASTM C76 standard

using the numerical analysis. The performance of the synthetic fiber reinforced concrete

pipes were evaluated under short- and long-term loading in accordance with ASTM

protocols using different pipe diameters. Fiber dosages ranged from 4.75 to 18 kg/m3 (8

to 30 lb/yd3), and different areas of one steel cage layer were used to reinforce the

concrete pipes. The finite element model of the three-edge bearing test was calibrated and

validated using the experimental results. The linear and non-linear behavior of the

synthetic fiber reinforced concrete material was characterized using the concrete damage
4

plasticity (CDP) model. For input data representing the concrete material properties,

compression strength, tensile strength, and modulus of elasticity were determined for five

fiber dosages 0, 4.75, 6, 7, and 9 kg/m3 (0, 8, 10, 12, and 15 lb/yd3). The results showed

that adding fiber to concrete enhanced the flexural strength, increased flexibility,

decreased the dynamic modulus of elasticity, and increased the CTE. Specimens

reinforced with PP fiber showed more flexural strength and flexibility than those

reinforced with PVA fiber. Using synthetic fiber increased the cracking load (produces

0.3 mm crack width), ultimate load, stiffness, and ductility of concrete pipes. Also, using

synthetic fiber lowered the production cost as the reduction in the steel cage area ranged

from 51 to 100%. Based on the experimental results, a new modified compression model

was adopted to represent the compression behavior of fiber reinforced concrete in the

finite element models. The tension behavior was defined using a model proposed by past

research. A parametric study was conducted on four parameters: pipe diameter, pipe wall

thickness, fiber dosage, and steel cage area. The parametric study results were

summarized in tables that can be used as a reference for a new synthetic fiber reinforced

concrete pipe standard.

Moreover, in response to sustained load, the tested pipes initially exhibited a

linear response, followed by a stable response with a slight increase in deflection over

time. Fiber creep did not significantly increase the crack width or affect the time

dependence of the strain, indicating the fibers adequately transfer the stress in the pipe

wall and limit the crack width.


5

DEDICATION

I dedicate this work to my love and life:


Amina, my wife,
our daughter Layan, our son Rayan
6

ACKNOWLEDGMENTS

I want to thank my advisor Dr. Shad M. Sargand, for his guidance and support

during my research and study. Also, I would like to thank my dissertation committee

members: Dr. Eric P. Steinberg and Dr. Issam Khoury, for their perceptive comments and

support during my graduate study at Ohio University. Also, I would like to express my

sincere gratitude to my sponsor, Higher Committee for Educational Development in Iraq

(HCED), University of Wasit, Iraq.

In addition, I would like to express my gratitude to the Advanced Drainage

Systems staff who facilitated the experimental part of this research, specifically Mr. John

Kurdziel and Mr. Daniel Figola. Also, I would like to thank Foltz Concrete Pipe Co. for

providing the fiber reinforced concrete material and facilitating the concrete pipes short-

term testing, specifically Mr. Shawn Coombs. I would also like to thank Dr. Hussam H.

Hussein, Dr. Anwer K. Al-Jhayyish, and Dr. Waleed Hamid for their help and support

through my graduate study. Finally, I also would like to thank my family and friends for

their support.
7

TABLE OF CONTENTS

Page
Abstract ...........................................................................................................................3
Dedication .......................................................................................................................5
Acknowledgments ...........................................................................................................6
List of Tables...................................................................................................................9
List of Figures ............................................................................................................... 11
Chapter 1: Introduction .................................................................................................. 19
1.1 Research Objectives .................................................................................... 23
Chapter 2 : Literature Review......................................................................................... 27
2.1 Introduction ................................................................................................ 27
2.2 Mechanical Properties of Fiber Reinforced Concrete Material ..................... 27
2.3 Fiber-reinforced Concrete Pipes .................................................................. 28
Chapter 3: Methodology ................................................................................................ 37
3.1 Material Properties of Synthetic Fiber Reinforced Concrete under Freeze-
Thaw Conditions ......................................................................................... 37
3.2 Evaluation of Synthetic Fiber Reinforced Concrete Pipe Performance Using
Three-Edge Bearing Test............................................................................. 47
3.3 Design Proposal for Synthetic Fiber Reinforced Concrete Pipes Using Finite
Element Analysis ........................................................................................ 52
3.4 Thin Walled Synthetic Fiber Reinforced Concrete Pipes ............................. 63
3.5 A New Test Method for Evaluating the Long-Term Performance of Fiber
Reinforced Concrete Pipes .......................................................................... 67
3.6 Long-term Performance of Synthetic Fiber Reinforced Concrete ................. 75
Chapter 4: Results and Discussions ................................................................................ 80
4.1 Introduction ................................................................................................ 80
4.2 Material Properties of Synthetic Fiber Reinforced Concrete under Freeze-
Material Properties ...................................................................................... 80
4.3 Effect of Freeze-thaw Cycles on the Mechanical Properties of Fiber
Reinforced Concrete.................................................................................... 91
8

4.4 Evaluation of Synthetic Fiber Reinforced Concrete Pipe Performance Using


Three-Edge Bearing Test........................................................................... 106
4.5 Design Proposal for Synthetic Fiber Reinforced Concrete Pipes Using Finite
Element Analysis ...................................................................................... 125
4.6 Experimental Investigation of Thin-Wall Synthetic Fiber Reinforced
Concrete Pipes .......................................................................................... 153
4.7 Performance of Thin-wall Synthetic Fiber Reinforced Concrete Pipes Under
Short and Long-term Loading ................................................................... 173
Chapter 5: Conclusions, Summary, and Recommendation ........................................... 213
5.1 Conclusions .............................................................................................. 213
5.2 Performance of Thin-wall Synthetic Fiber Reinforced Concrete Pipes Under
Short and Long-term Loading ................................................................... 218
5.3 Recommendation ...................................................................................... 219
References ......................................................................................................... 220
Appendix I: List of Journal Papers ............................................................................... 236
Appendix II: List of Conference Papers ....................................................................... 239
9

LIST OF TABLES

Page

Table 1. Physical Properties of the Fiber. ....................................................................... 38

Table 2. Mix Proportions of the Concrete. ..................................................................... 38

Table 3. Mix Proportions of the Concrete ...................................................................... 48

Table 4. Details of the Tested Fiber Reinforced Concrete Pipe Specimens. .................... 49

Table 5. ASTM C76 (2015) Pipe Class, Cracking Load, and Ultimate Load .................. 51

Table 6. Details of the Tested Specimens. ...................................................................... 53

Table 7. Physical and mechanical properties of the PP Fiber. ......................................... 54

Table 8. Mix proportions of the concrete used to make the fiber-reinforced pipes. ......... 55

Table 9. Testing results of wire reinforcing used for finite element models of the concrete

pipes. ............................................................................................................................. 63

Table 10. Physical Properties of the Synthetic Fiber....................................................... 64

Table 11. Mix Proportions of the Concrete Mixtures. ..................................................... 64

Table 12. Details of the Tested Specimens. .................................................................... 65

Table 13. ASTM C76 (2015) Class Criteria for Cracking Load and Ultimate Load ........ 66

Table 14. Physical Properties of the Synthetic Fiber....................................................... 76

Table 15. Mix Proportions of the Concrete Mixtures. ..................................................... 76

Table 16. Details of the Tested Pipe Specimens, each having length 1.2 m. ................... 77

Table 17. CTE Measured Using Ohio Method Before Applying Freeze-Thaw Cycles. ... 84

Table 18. CTE Results of Ohio CTE Device and AASHTO T60-00. .............................. 86
10

Table 19. Dynamic Modulus of Elasticity of Concrete Prisms with Different Fiber

Dosages. ........................................................................................................................ 87

Table 20 Flexural Strength of Fiber Reinforced Concrete. ............................................. 89

Table 21. Summary of the Crack and Ultimate Loads of Tested Pipes.......................... 119

Table 22. Stiffness Results at 2% Deflection Ratio. .................................................. 121

Table 23.The average compressive strength, tensile strength, modulus of elasticity and

Poisson’s ratio test results. ........................................................................................... 133

Table 24. FE and experimental results of ultimate strength of the calibrated pipes. ...... 142

Table 25. FE and Experimental Results of Ultimate Strength of the Validate Models. .. 145

Table 26. Design requirements for Class I synthetic fiber reinforced concrete pipes. .... 147

Table 27. Design requirements for class II synthetic fiber reinforced concrete pipes. ... 148

Table 28. Design requirements for class II synthetic fiber reinforced concrete pipes. ... 149

Table 29. Design requirements for Class IV synthetic fiber reinforced concrete pipes. . 150

Table 30. Design requirements for Class V synthetic fiber reinforced concrete pipes. .. 151

Table 31. ASTM C76 (2015) Class Criteria for Cracking Load and Ultimate Load ...... 169

Table 32. Summary of the Crack and Ultimate Loads of Tested Pipe ........................... 170

Table 33. Stiffness of Tested Pipes at deflection ratio of 5%. ....................................... 172

Table 34. Stiffness of Tested Pipes at Deflection Ratio of 3%. ..................................... 172
11

LIST OF FIGURES

Page

Figure 1. Plot of Required Inside Reinforcing Area vs. Design Height of Earth Cover for

Typical Design with Surface Wheel Loads (ASCE15-98). ............................................. 21

Figure 2. a) Typical pipe cracking b); Internal stresses on free body A; c) Internal stresses

on element B. ................................................................................................................ 22

Figure 3. Ohio CTE specimen’s preparation. ................................................................. 39

Figure 4. Ohio CTE device: (a) schematic device details; and (b) specimen installed

inside the chamber. ........................................................................................................ 41

Figure 5. Temperature profile. ....................................................................................... 41

Figure 6. AASHTO-CTE test setup: (a) schematic a mearing setup; and (b) specimen

setup. ............................................................................................................................. 42

Figure 7. Temperature profile used for AASHTO TP60-00 CTE test method. ................ 43

Figure 8. Dynamic modulus of the elasticity test setup. .................................................. 44

Figure 9. Flexural test setup. .......................................................................................... 45

Figure 10. Freeze-thaw cabinet. ..................................................................................... 46

Figure 11. Pipe production: (a) steel cage reinforcement, (b) mixture, and (c) removal of

the jacket. ...................................................................................................................... 50

Figure 12. a) Specimen setup; b) schematic three-edge bearing test setup; c) crack width

indicator. ....................................................................................................................... 51

Figure 13. Polypropylene (PP) fiber used in this research............................................... 54


12

Figure 14. Test setup of (a) Compressive strength (b) Poisson’s ratio, and (c) splitting

tensile strength............................................................................................................... 55

Figure 15. Finite element modeling of the three-edge bearing test. (a) A-A cross-section;

(b) 3D model. ................................................................................................................ 57

Figure 16. Tensile stresses vs. crack mouth opening displacement for fibers, matrix, and

fiber-matrix composite................................................................................................... 59

Figure 17. Tensile stresses-crack mouth opening displacement plots for PP reinforced

concrete material used in the finite element models........................................................ 61

Figure 18. a) Specimen setup; b) schematic three-edge bearing test setup. ..................... 67

Figure 19. FE model of the long-term test setup ............................................................. 69

Figure 20. compressive stress-strain plots used in the FE models. .................................. 70

Figure 21. Tensile stresses-crack mouth opening displacement plots for PP reinforced

concrete material used in the finite element models........................................................ 71

Figure 22. Long-term test setup detail selected based on the FE analysis........................ 73

Figure 23 The damage in the concrete pipe due to the applied. ....................................... 73

Figure 24. The vertical displacement for (a) whole test setup (b) upper beam (c) bottom

beam. ............................................................................................................................. 75

Figure 25. (a) Photograph and (b) schematic diagram of the long-term test setup. .......... 79

Figure 26. Thermal strain-temperature of (a) PP and (b) PVA fiber reinforced specimens.

...................................................................................................................................... 82

Figure 27. Load versus deflection plots of (a) PP and (b) PVA fiber reinforced specimens.

(Note: 1 kN = 0.2248 kip; 1 mm = 0.0394 in). ............................................................... 90


13

Figure 28. The crack width of (a) PP and (b) PVA fiber reinforced specimens. (Note: 1

mm = 0.0394 in). ........................................................................................................... 91

Figure 29. CTE versus freeze-thaw cycles: (a) PVA and; (b) PP fiber reinforced concrete

specimens. (Note: Multiply 1/°C by 5/9 to obtain 1/°F.)................................................. 93

Figure 30. The damage in the meso-structure of PVA fiber reinforced specimen with fiber

dosage of 7 kg/m3 (12 lb/yd3)......................................................................................... 94

Figure 31. Mass of fiber reinforced concrete prisms as a function of freeze-thaw cycles:

(a) PP and (b) PVA fiber reinforcement. ........................................................................ 96

Figure 32. Dynamic modulus of elasticity versus freeze-thaw cycles: (a) PP and (b) PVA

fiber reinforced specimens. (Note: 1 MPa = 0.145 ksi) ................................................... 99

Figure 33. Flexural strength versus freeze-thaw cycles for concrete reinforced with (a)

PVA and (b) PP fiber. (Note: 1 MPa = 0.145 ksi) ........................................................ 102

Figure 34. Load versus deflection of prisms reinforced with PVA fiber with different

dosages before and after applying 300 freeze-thaw cycles: 6 kg/m3 (10 lb/yd3); (b) 7

kg/m3(12 lb/yd3); (c) 9 kg/m3 (15 lb/yd3). (Note: 1 kN = 0.2248 kip; 1 mm = 0.0394 in)

.................................................................................................................................... 104

Figure 35. Load versus deflection of prisms reinforced with PP fiber with different

dosages before and after applying 300 freeze-thaw cycles: 6 kg/m3 (10 lb/yd3); (b) 7

kg/m3(12 lb/yd3); (c) 9 kg/m3 (15 lb/yd3). (Note: 1 kN = 0.2248 kip; 1 mm = 0.0394 in).

.................................................................................................................................... 105
14

Figure 36. Load deflection curves of 600 mm (24 in.) pipe diameter reinforced with: (a)

synthetic fiber only; (b) synthetic fiber with one steel cage layer. (Note: 1 kN = 0.2248

kip; 1 mm = 0.0394 in) ................................................................................................ 110

Figure 37. Load deflection curves of 1200 mm (48 in.) pipes diameter: (a) plain concrete

pipes; (b), pipes reinforced with fiber dosage of 4.75 and 9 kg/m3; (b, c, d, e, and f) pipes

reinforced with fiber dosage of 4.75, 9, 13.5, and 18 kg/m3, respectively, along with one

steel cage layer. (Note: 1 kN = 0.2248 kip; 1 mm = 0.0394 in)..................................... 113

Figure 38. Load deflection curves of 1500 mm pipe diameter reinforced with one steel

cage layer along with fiber dosage of: (a) 9 kg/m3 ;(b) 13.5 kg/m3 ;(c) 18 kg/m3. (Note: 1

kN = 0.2248 kip; 1 mm = 0.0394 in) ............................................................................ 115

Figure 39. Crack development of concrete pipes: 1200 mm (48 in.). ............................ 116

Figure 40. Summary of deflection ratio of fiber-reinforced concrete pipes. .................. 125

Figure 41. Load deflection curves of 600 mm (24 in.) diameter pipe reinforced with (a)

PP fiber only; (b) PP fiber with one steel cage layer. (Note: 1 kN = 0.2248 kip; 1 mm =

0.0394 in). ................................................................................................................... 127

Figure 42. Load deflection curves of 1200 mm (48 in.) diameter pipe: (a) plain concrete

pipes; (b), pipes reinforced with PP fiber dosage of 4.75 and 9 kg/m 3 (8 and 15 lb/yd3);

(c, d, e, and f) pipes reinforced with a steel cage layer and PP fiber dosage of 4.75, 9,

13.5, and 18 kg/m3 (8, 15, 22.5 and 30 lb/yd3), respectively. (Note: 1 kN = 0.2248 kip; 1

mm = 0.0394 in). ......................................................................................................... 130


15

Figure 43. Load deflection curves of 1500 mm (60 in.) diameter pipe reinforced with one

steel cage layer and PP fiber dosage of: (a) 9 kg/m3 (15 lb/yd3);(b) 13.5 kg/m3(22.5

lb/yd3) ;(c) 18 kg/m3 (30 lb/yd3). . (Note: 1 kN = 0.2248 kip; 1 mm = 0.0394 in).......... 132

Figure 44. Comparison of experimental stress-strain curves of concrete cylinders

reinforced PP with dosage of (a) 4.75 kg/m3 (b) 6 kg/m3; (c) 7 kg/m3; and (d) 9 kg/m3

with various models of compressive stress response from the literature........................ 135

Figure 45. compressive stress-strain plots used in the FE models. ................................ 136

Figure 46. Experimental versus calibrated FE load-deflection responses of 1200 mm (48

in.) diameter pipe reinforced with fiber dosage of (a) 9 kg/m3 (15 lb/yd3), (b) 9 kg/m3 (15

lb/yd3) with 510mm2/m (0.24 in2/ft) steel area (c) 13.5 kg/m3 (22.5 lb/yd3)with 510

mm2/m (15 lb/yd3) steel area (d) 18 kg/m3 (30 lb/yd3) with 510mm2/m (0.24 in2/ft) steel

area. . (Note: 1 kN = 0.2248 kip; 1 mm = 0.0394 in). ................................................... 139

Figure 47. Crack propagation (a) experimental results (b) Maximum principal strain

contour at crown and invert (c) Maximum principal strain contour at the crown, invert,

and springline (c,d, and e) cracks propagation. ............................................................. 141

Figure 48. Experimental versus FE load-deflection responses: a) 600 mm (24 in.)

diameter pipe reinforced with 9 kg/m3; (b), (c) and (d) 1500 mm diameter pipe reinforced

with steel area of 510 mm2/m and fiber dosage of 9,13.5 and 18 kg/m 3, respectively.

.(Note: 1 kN = 0.2248 kip; 1 mm = 0.0394 in). ............................................................ 144

Figure 49. Load-deflection responses of 1350 mm (54 in) (B wall) pipe diameter with

different fiber dosage and steel cage areas imposed on (a) 0.3 mm (0.01 in.) and (b)
16

ultimate loads specified in ASTM C76 for pipe strength classes. (Note: 1 kN = 0.2248

kip; 1 mm = 0.0394 in.) . ............................................................................................. 152

Figure 50. Load deflection curves of 1200 mm (48 in.) diameter pipes reinforced with PP

fiber dosage of 9 kg/m3 along with reinforcing steel area of: (a) 5.7 cm2/m (0.27 in2/ft);

(b) 10.2 cm2/m (0.48 in2/ft). (Note: 1 kN = 0.2248 kip; 1 mm = 0.0394 in.). ................ 157

Figure 51. Load deflection curves of 1500 mm (60 in.) diameter pipes reinforced with PP

fiber dosage of 9 kg/m3 along with reinforcing steel area of: (a) 5.7 cm2/m (0.27 in2/ft);

(b) 8.9 cm2/m (0.42 in2/ft). (Note: 1 kN = 0.2248 kip; 1 mm = 0.0394 in.). .................. 158

Figure 52. Strain measurements of pipe with diameter of 1200 mm (48 in) reinforced

with PP fiber dosage of 9 kg/m3(15 lb/yd3) along with steel mesh at amounts indicated in

the legend: (a) inner crown (b) inner springline (c) inner invert (d) outer springline.

(Note: 1 kN = 0.2248 kip). ........................................................................................... 162

Figure 53. Strain measurements of pipe with diameter of 1500 (60 in.) mm reinforced

with PP fiber dosage of 9 kg/m3 (15 lb/yd3) along with steel mesh at amounts indicated in

the legend: (a) inner crown (b) inner springline (c) inner invert (d) outer springline.

(Note: 1 kN = 0.2248 kip). ........................................................................................... 164

Figure. 54. Crack development in 1200 mm (48 in.) concrete pipe diameter reinforced

with 9 kg/m3 (15 lb/yd3) PP fiber and 5.7 cm2/m (0.27 in2/ft) steel cage: (a) crown; (b)

invert; (c) springline. ................................................................................................... 167

Figure. 55. Crack development in 1200 mm (48 in.) concrete pipe diameter reinforced

with 9 kg/m3 (15 lb/yd3) PP fiber and 10.2 cm2/m (0.48 in2/ft) steel cage: (a) crown; (b)

invert; (c) springline. ................................................................................................... 167


17

Figure. 56. Crack development in 1500 mm (60 in.) concrete pipe diameter reinforced

with 9 kg/m3 (15 lb/yd3) PP fiber and 5.7 cm2/m (0.27 in2/ft) steel cage: (a) crown; (b)

invert; (c) springline. ................................................................................................... 167

Figure 57. Crack development in 1500 mm (60 in.) concrete pipe diameter reinforced

with 9 kg/m3 (15 lb/yd3) PP fiber and 8.9 cm2/m (0.42 in2/ft) steel cage: (a) crown; (b)

invert; (c) springline. ................................................................................................... 168

Figure 58. Load deflection curves of 1200 mm diameter pipes reinforced with PP fiber

dosage of 9 kg/m3 and reinforcing steel area of: (a) 5.7 cm2/m(0.27 in2/ft); (b) 10.2

cm2/m(0.48 in2/ft). (Note: 1 kN = 0.2248 kip; 1 mm = 0.0394 in.) ............................... 175

Figure 59. Load deflection curves of 1500 mm diameter pipes reinforced with PP fiber

dosage of 9 kg/m3 and reinforcing steel area of: (a) 5.7 cm2/m (0.27 in2/ft); (b) 8.9 cm2/m

(0.42 in2/ft). (Note: 1 kN = 0.2248 kip; 1 mm = 0.0394 in.) ......................................... 176

Figure 60. Load vs. displacement plot of 1200 mm diameter concrete pipe reinforced

with steel area of 5.7 cm2/m (0.27 in2/ft) and PP fiber dosage of 9 kg/m3, (a) Stage 1

(40% ultimate load); (b) Stage 2 (50% ultimate load); (c) Stage 3 (70% ultimate load).

(Note: 1 kN = 0.2248 kip; 1 mm = 0.0394 in.). ............................................................ 179

Figure 61. Displacement vs. time plots of 1200 mm (48 in.) diameter concrete pipe

reinforced with steel area of 5.7 cm2/m (0.27 in2/ft) and PP fiber dosage of 9 kg/m3. (a)

Load Stage 1 (b) Load Stage 2 (c) Load Stage 3. (Note: 1 mm = 0.0394 in.). ............... 182

Figure 62. Crack width versus time plots of 1200 mm diameter concrete pipe reinforced

with steel area of 5.7 cm2/m and PP fiber dosage of 9 kg/m3(15 lb/yd3) during (a) Load

Stage 2 (b) Load Stage 3. (Note: 1 mm = 0.0394 in.) ................................................... 185


18

Figure. 63. LVDTs used to measure the crack width. ................................................... 186

Figure 64. Load-strain responses (a) invert and crown (b) springline (c) midway crown-

haunch. ........................................................................................................................ 189

Figure 65. Strain responses as a function of time at (a) springline (b) midway crown-

haunch. ........................................................................................................................ 191

Figure 66. Cracks formed during Load Stage 1 at (a) crown (b) springline (c) invert. ... 193

Figure 67. Crack propagated during Load Stage 2 at (a) crown (b) springline .............. 194

Figure 68. Cracks propagated during Load Stage 3 (a) crown (b) invert (c) springline. 195

Figure 69. Displacement vs. time plots of 1200 mm (48 in) diameter concrete pipe

reinforced with steel area of 5.7 cm2/m (0.27 in2/ft) and PP fiber dosage of 9

kg/m3(15lb/yd3). (a) Load Stage 1 (b) Load Stage 3. (Note:1 mm = 0.0394 in.)............ 198

Figure 70. Crack width versus time plots of 1200 mm diameter concrete pipe reinforced

with steel area of 5.7 cm2/m (0.27 in2/ft) and PP fiber dosage of 9 kg/m3(15lb/yd3). (a)

Load Stage 1 (b) Load Stage 3. (Note:1 mm = 0.0394 in.)............................................ 201

Figure 71. LVDTs used to measure the crack width. .................................................... 202

Figure 72. Creep coefficient (a) Load stage 1 (b) load stage 2 ...................................... 205

Figure 73. Strain responses as a function of time at (a) springline (b) midway crown-

haunch. ........................................................................................................................ 208

Figure 74. Cracks formed during Load Stage 1 at (a) crown (b) springline (c) invert. ... 211

Figure 75. Cracks propagated during Load Stage 3 (a) crown (b) invert (c) springline. 212
19

CHAPTER 1 : INTRODUCTION

Storm sewer system pipes used in the United States are usually made of concrete

reinforced with a steel cage. The cage is located at least 25 mm (1 in.) from the outer face

of the concrete pipe to resist flexural stresses. However, when the pipe diameter exceeds

900 mm (36 in.), other failure mods start to evolve such as radial tension and diagonal

shear. Therefore, the steel cage alone is not sufficient to accommodate for these modes

and cracking of steel reinforced pipe becomes more likely. Additionally, thermal

cracking may also appear on the pipe surface in response to seasonal and daily

temperature fluctuations. Surface cracks admit water and dissolved chemicals into the

concrete pipe wall, resulting in corrosion of the steel reinforcement and deterioration of

the pipe structure. This could be addressed by either increase the steel reinforcement,

which is not a practical solution and difficult to implement, or by incorporating fibers

into concrete mix design. Synthetic fibers have been used to improve the mechanical

properties and performance of concrete pipes in more feasible and economical ways

(Peyvandi et al, 2014; Park et al, 2014 and 2015; Wilson and Abolmaali, 2014; de la

Fuente et al, 2013).

Pipes can be either flexible or rigid based on the relative stiffness of the pipe to

the surrounding soil and the deflection ratio. The most common materials used for

manufacturing flexible pipes around the world are plastic and metal, typically steel. In an

attempt to reduce the cost of flexible pipes and enhance their stiffness, a corrugated wall

profile has been adopted. Corrugated pipes tend to have less wall area compared to rigid

pipes. Also, plastic and metal pipes with corrugated walls have low stiffness. For
20

example, 1200 mm (48 in.) diameter plastic pipe has stiffness as low as 0.124 MPa (50

psi) with deflection up to 5% of their inside diameter, which is quite flexible (ASTM

F2648, 2013). Flexible pipes require special installation procedures, including precise

control of compaction, and backfill levels. When flexible pipes deflect during installation,

compressive stresses in the pipe wall will be generated. By their nature, flexible pipes are

more susceptible to local buckling within the corrugation profile, and to bending strain

from the backfill loads and installation forces. Such issues are not evident in a standard

manufactured concrete pipe, a rigid material.

However, concrete pipes are brittle and experience a small deflection when loaded

before it shows serious damage. rigid pipes such as concrete pipes primarily depend on

their wall strength to carry applied loads. Structural damage of these pipes appears at

vertical deflections of less than 2% of its inside diameter (Plastics Pipe Institute, 2011).

Therefore, design methods for concrete pipes do not take into consideration the ability of

the surrounding soil to carry a portion of the load. As pipe diameter and earth cover

increases, so does the need for increased reinforcing steel area, as additional failure

modes must be countered (flexural, crack, shear), as depicted in Figure 1.


21

Figure 1. Plot of Required Inside Reinforcing Area vs. Design Height of Earth Cover for

Typical Design with Surface Wheel Loads (ASCE15-98).

In some modes, such as shear and radial failures, cracks propagate in diagonal and

radial directions, respectively, so circumferential steel reinforcement is no longer

sufficient. Additionally, shear stirrups must be incorporated into the pipe’s manufacture.

This shear reinforcement is very expensive and difficult to implement within the pipe

wall. Figure 2 shows a typical pipe cracking and internal stresses that may generate due

to different load applications.


22

(a) (b) (c)

Figure 2. a) Typical pipe cracking b); Internal stresses on free body A; c) Internal stresses

on element B.

One approach to overcome these issues in concrete pipes is by adding synthetic

fibers to the concrete mixture (Peyvandi et al, 2014; Park et al, 2014; Wilson and

Abolmaali, 2014; de la Fuente et al, 2013). Synthetic fibers improve the concrete

mechanical properties such as tensile strength, shear strength, flexural strength, and

modulus of elasticity. This improvement is mainly influenced by the size, type,

geometry, dispersion and volume fraction of fibers (Altoubat et al.,2009; Balaguru and

Shah, 1992; Çavdar, 2012, 2013, and 2014; Mehta and Monteiro, 2006; Kuder and Shah,

2010).

Understanding the performance and structural behavior of synthetic fiber

reinforced concrete pipes is in its infancy; no design code has been provided by ASTM or

AASTHTO for these new synthetic fiber concrete pipes. There are few studies in the

literature regarding synthetic fiber reinforced concrete pipes. However, there is


23

uncertainty about the long-term behavior of fiber reinforced concrete pipes under

sustained load, as they could exhibit increased deformation over time due to the

viscoelastic properties of the synthetic fiber; past research has yielded conflicting results.

1.1 Research Objectives

The aim of this study is to improve the performance of the conventional concrete

pipes by incorporating synthetic fiber as a replacement for the conventional steel

reinforcement cage. This aim was met via the following six objectives of this research: 1.

Investigate the effect of temperature change on the mechanical properties of fiber

reinforced concrete, including CTE, dynamic modulus of elasticity, and flexural strength.

2. Evaluation of synthetic fiber reinforced concrete pipe performance using three-edge

bearing test, 3. Design proposal for synthetic fiber reinforced concrete pipes using finite

element analysis, 4. Experimental investigation of thin-wall synthetic fiber reinforced

concrete pipes, 5. A new test method for evaluating the long-term performance of fiber-

reinforced concrete pipes, and 6. Performance of thin-wall synthetic fiber reinforced

concrete pipes under short and long-term loading. A more detailed description of how

each of the above objectives was met follows:

1.1.1 Material Properties of Synthetic Fiber Reinforced Concrete under Freeze-Thaw

Conditions

The objective of this study was to evaluate the effect of PVA and PP fiber

dosages on the CTE, dynamic modulus of elasticity, and flexural strength of fiber

reinforced concrete. Four different dosages of each fiber type, 0, 6, 7, and 9 kg/m 3 ( 0, 8,

10, 12, amd 15 lb/yd3) were evaluated. Test specimens were subjected to 300 freeze-thaw
24

cycles in accordance with ASTM C666/C666M-15 (2015a) to evaluate the fiber

reinforced concrete material performance under extreme weather conditions.

1.1.2 Evaluation of Synthetic Fiber Reinforced Concrete Pipe Performance Using

Three-Edge Bearing Test

The main objective of this study was to investigate the performance of synthetic

fiber reinforced concrete pipes under a three-edge bearing test in accordance with ASTM

C497 (2014). The performance of the concrete pipes was evaluated in terms of failure

mechanism, load carrying capacity, ductility, stiffness, and post-cracking behavior. Three

variables were considered in this research, which are pipe diameter, fiber dosage, and the

amount of steel reinforcement. The results of cracking and ultimate load were compared

with those specified by ASTM C76 (2015).

1.1.3 Design Proposal for Synthetic Fiber Reinforced Concrete Pipes Using Finite

Element Analysis

The objective of this study was to develop a three-dimensional finite element

model to simulate the three-edge bearing test of PP fiber reinforced concrete pipes. This

model was calibrated and validated in terms of the load carrying capacity and load-

deflection response obtained from experimental tests presented by Al Rikabi et al. (2018).

Different pipe wall thicknesses and diameters were used in the model to obtain the

optimum PP fiber dosage with and without steel cage reinforcement to fulfill the strength

requirements specified by ASTM C76 (2014). The results were presented in tables to

simplify the design process of PP fiber reinforced concrete pipe.


25

1.1.4 Experimental Investigation of Thin-wall Synthetic Fiber Reinforced Concrete

Pipes

The main objective of this study was to develop a new concrete pipe system, which

is lighter, cheaper, flexible, and less sensitive to installation type. The fiber-reinforced

concrete pipes evaluated had diameters of 1200 and 1500 mm (48 and 60 in.) with

respective wall thickness of 50 and 63 mm (2 and 2.5 in.), reduced to 50% of that specified

in ASTM C76 (2015) for wall A. They were instrumented with strain gauges and string

potentiometers then subjected to the three-edge bearing test in accordance to ASTM C497

(2015) to evaluate the new fiber-reinforced concrete pipe system in terms of mode of

failure, load carrying capacity, and flexibility. In order to assure that synthetic fiber will

significantly enhance the concrete pipe strength, fiber content of 9 kg/m3(15 lb/yd3) was

used along with different steel cage areas).

1.1.5 Performance of Thin-wall Synthetic Fiber Reinforced Concrete Pipes Under

Short and Long-term Loading

The main objective of this study was to evaluate the PP fiber reinforced concrete

pipe performance subjected to short and long-term loading. Concrete pipes with a

diameter of 1200 and 1500 mm (48 and 60 in.) with respective wall thickness of 50 and

63 mm (2 and 2.5 in.), reduced to 50% of that specified in ASTM C76 (2015) for wall A,

were tested under short and long-term load. To get a better understanding of the pipes’

behavior subjected to short and long-term loads, pipes were instrumented using high-

resolution strain gauges and string potentiometers. All pipes were reinforced using fiber
26

dosage of 9 kg/m3 (15 lb/yd3) along with different steel areas to assure significant

enhancement can be contributed by the synthetic fiber to the concrete pipe strength.
27

CHAPTER 2 : LITERATURE REVIEW

2.1 Introduction

This chapter presents a comprehensive literature review on experimental and

numerical analyses of fiber reinforced concrete material. This chapter is divided into two

sections covering the laboratory investigation on synthetic fiber reinforced concrete

material and incorporating synthetic fiber in concrete pipes mixture to minimize the

conventional cage area.

2.2 Mechanical Properties of Fiber Reinforced Concrete Material

For decades, synthetic fibers have been found to enhance the mechanical

properties of concrete, including flexural strength, tensile strength, shear strength,

ductility, shrinkage, and creep. The change in temperature between day and

night subjects concrete structures to daily cycles of stress. Freeze-thaw cycles occur when

the freezing point of water lies between the daily extremes, and the increased stress from

freezing enlarges surface cracks. Under existing applied loads, these cracks propagate

deeper inside the concrete structure and allow the water and chemical agents to migrate

inside the concrete (Gergely and Lutz, 1968). When chemical agents, such as those used

for highway de-icing, reach the steel reinforcement, the ensuing corrosion can lead to

concrete spalling and loss of strength. Richardson et al. (2012) performed an

experimental study on the durability of the air-entrained concrete and Type 1

polypropylene reinforced concrete. Three groups of air entrained concrete cubic

specimens of nominal dimension 150 mm (reinforced with 0.9 kg/m3 and 1.8 kg/m3 of

Type 1 polypropylene fiber, and plain concrete,) were tested. The researchers measured
28

mass loss, relative pulse velocity, and final compressive strength of the specimens after

applying 300 freeze-thaw cycles. The added fiber significantly improved the durability

and compressive strength over that of plain concrete. Yun et al. (2013) conducted an

experimental study to investigate the effect of freeze-thaw cycles on the flexural behavior

of fiber reinforced concrete. A 1.5% combination volume fraction of polypropylene and

polyvinyl alcohol fiber was used to reinforce the plain concrete. Results showed that

using this combination of synthetic fibers increased the durability of concrete undergoing

freeze-thaw cycles. Zhang and Li (2013) investigated the effect of adding polypropylene

fiber on the workability and durability of concrete composite containing fly ash and silica

fume. Low volume fractions of polypropylene fiber, 0.06%, 0.08%, 0.1%, and 0.12%,

were used. The researchers applied 300 freeze-thaw cycles on the specimens. The

durability of concrete composite improved with the addition of polypropylene fiber.

However, the volume fraction of PP is not recommended in other studies since it is

not sufficient to enhance the tensile behavior of the fiber reinforced

concrete. Çavdar (2014) evaluated the mechanical properties of fiber reinforced concrete

subjected to 100 freeze-thaw cycles using carbon, polypropylene, aramid, polyvinyl

alcohol, and glass fibers at volume fractions of 0.0%, 0.4%, 0.8%, and 1.2% for each

type. All fiber reinforced concrete specimens exhibited higher resistance to freeze-thaw

cycles than the plain concrete.

2.3 Fiber-reinforced Concrete Pipes

Understanding the performance and structural behavior of synthetic fiber

reinforced concrete pipes is in its infancy; no design code has been provided by ASTM or
29

AASTHTO for these new synthetic fiber concrete pipes. There are few studies in the

literature regarding synthetic fiber reinforced concrete pipes.

2.3.1 Design Proposal for Synthetic Fiber Reinforced Concrete Pipes

Recently, several researchers have tried to improve the performance of concrete

pipes by incorporating synthetic fibers to augment or replace steel cage reinforcement

(Park et al, 2016; Peyvandi et al., 2013 and 2014; Wilson and Abolmaali, 2014; de la

Fuente et al, 2013). Synthetic fibers contribute significantly in enhancing mechanical

properties of concrete, including flexural strength, shear strength, tensile strength, and

dynamic modulus of elasticity. The size of this enhancement is mainly proportional to

fiber type, dosage, geometry, and dispersion (Balaguru and Shah, 1992; Çavdar, 2014;

Mehta and Monteiro, 2006; Kuder and shah, 2010). Although synthetic fibers have been

used as alterative reinforcement for the steel cage in concrete pipes, no design standard

has been introduced. Past research has tried to determine the optimum fiber dosage to

meet strength requirements specified by the ASTM standards for specific pipe diameter

and wall thickness. Peyvandi et al. (2013 and 2014) performed an experimental study to

investigate the performance of polyvinyl alcohol (PVA) fiber reinforced concrete pipes.

The researchers tested two categories of concrete pipes with diameter of 675 mm (27 in.),

one reinforced with fiber only, and the second including one layer of steel cage. Findings

showed that using fiber at 0.5% to 0.75% by volume offered the desired balance between

economy and strength effects. De la Fuente. (2013) tested pipes diameter of 1000 mm

(39 in.) reinforced with polypropylene (PP) fiber using crushing test. The researchers

used three dosages of fiber, 3, 4.5, and 6 kg/m3 (5, 7.5, and 10 lb/yd3). Results were used
30

to validate a numerical model for mechanical analysis of pipes. Design tables for the PP

fiber reinforced concrete pipes with diameter ranging from 300 to 1000 mm (12 to 39 in.)

with two wall thicknesses were created. The designed pipes were reinforced only with

fiber at dosage ranging from 3 to 9 kg/m3 (5 to 15 lb/yd3). However, using fiber only as

sole reinforcement may not be enough to accommodate both flexural loads and other

failure modes such as radial and shear failure as pipe diameter and ASTM strength

requirement of higher classes increase. Also, the researchers assumed that the

compression strength of PP fiber dosage increases with fiber dosage increase. However,

past research showed that increasing fiber dosage had a negative impact on the

compressive strength (Zhang et al., 1999; Huang, 2001; Puertas et al., 2003; Choi and

Yuan, 2005; Çavdar, 2014). Wilson et al. (2014) investigated the performance of PP fiber

reinforced concrete pipes with diameter ranging from 375 mm to 900 mm (15 to 36 in.)

under three-edge bearing load. The researchers used fiber dosages of 2.3, 3.6, 4.7, 6, 7,

9.5, and 10.6 kg/m3(3.87, 6, 7.9, 10.1, 16, and 17.86 lb/yd3). The results indicated that

using PP fiber is a suitable replacement for the steel reinforcement cage. Also, the

researchers recommended further studying the effect of added fiber on the pipes with a

diameter larger than 900 mm (36 in.). Park et al. (2016) studied the effect of inclusion PP

fiber dosages ranging from 1.2 to 14.3 kg/m3 (2 to 24 lb/yd3) on the thin-walled concrete

pipes. The pipes with a diameter ranging from 760 to 3050 mm (30 to 122 in) were tested

using the three-edge bearing test. In addition to the PP fiber, pipes were reinforced with a

steel cage area reduced to 35% - 50% of the steel cage area specified by ASTM C76.

Certain fiber dosages were found to enhance the performance of the tested pipes.
31

2.3.2 Thin Walled Synthetic Fiber Reinforced Concrete Pipes

Understanding of the structural performance of synthetic fiber reinforced concrete

pipes is still in its infancy. No codes or standards have been introduced. The ASTM C76

standard includes synthetic fiber as a possible replacement for steel reinforcement cage

without presenting any design. Few studies have been conducted on concrete pipes

incorporating synthetic fibers. Peyvandi et al. (2013 and 2014) conducted an

experimental study on PVA fiber reinforced concrete pipes. Concrete pipes of diameter

of 675 mm (ASTM C76 (2012) Class IV, C-wall pipe) reinforced with different fiber

dosages into groups were tested under the three-edge bearing test. In the first test group,

the pipes were reinforced with fiber only; for the second group, pipes had additional steel

reinforcement. Investigators found that using fiber content of 0.5 to 0.75% enhanced the

load capacity and ductility of tested pipes. Also, they reported that this fiber dosage offers

a balance between strength and economy. Fuente et al. (2014) investigated the

performance of polypropylene (PP) fiber reinforced concrete pipe of diameter 1000 mm

(39 in.) with wall thickness of 80 mm (3 in.). The pipe was reinforced at fiber dosages of

3, 4.5, and 6 kg/m3 (5, 7.5, and 10 lb/yd3). Researchers concluded that using synthetic

fiber is a suitable replacement for the steel cage. They recommended further investigation

of the PP fiber reinforced concrete pipe performance.

Park et al. (2014) performed an experimental study to investigate the performance

of synthetic fiber reinforced concrete pipes using the three-edge bearing test. Concrete

pipes with a diameter of 600 and 900 mm (24 and 36 in. [Class III Wall-B]) were

reinforced using three fiber dosages of 3.5, 4.8, and 7 kg/m3 (5.9, 8, and 11.8 lb/yd3)
32

along with one steel cage layer. Their results showed pipes could fulfill the strength

requirement specified by ASTM C76 (2015) Class III with a fiber dosage of 4.8 and 7

kg/m3 (4.8 and 11.8 lb/yd3) for pipes of diameter 600 and 900 (24 and 36 in.),

respectively. Wilson et al. (2014) evaluated the structural performance of synthetic fiber

reinforced concrete pipes using three-edge bearing test. The pipes tested had diameters

ranging from 375 to 900 mm (15 to 36 in.) reinforced with fiber dosages of 2.3, 3.6, 4.7,

6, 7, 9.5, and 10.6 kg/m3 (3.87, 6, 7.9, 10.1, 16, and 17.86 lb/yd3). The pipes were

manufactured in two production sites using different manufacturing equipment. It was

concluded that each pipe diameter has an optimal fiber dosage satisfying the strength

requirements specified by ASTM C76 (2015). Also, the optimum fiber dosage depends

on cementitious material content, local aggregate, and manufacturing plant equipment.

The authors also recommended studying the effect of adding synthetic fiber to pipes with

diameters exceeding 900 mm (36 in.). Park et al., (2016) studied the structural

performance of thin-walled concrete pipe, including synthetic fiber using the three-edge

test. Fiber dosages ranging from 1.2 to 14.3 kg/m3 (2 to 24 lb/yd3) combined with steel

cage area reduced to 35 to 50% of conventional steel area required by ASTM C76 were

used in pipes of diameter 760 to 3050 mm (30 to 122 in). It was concluded that these

pipes had enhanced load carrying capacity and shear resistance compared to standard

pipes without fiber reinforcement. Also, the inclusion of fiber enhanced the ductility as

the most tested pipes withstood deflection between 3 to 5% of the inside diameter.

In the literature, there is a clear consensus that the addition of synthetic fibers

makes concrete pipes more flexible than traditional concrete, and the steel reinforcement
33

can be reduced or even eliminated. However, most researchers focused on using the same

wall thickness specified by ASTM C76 (2015). Decreasing the wall thickness of the

concrete pipe while adding an optimum fiber dosage may result in a significant

enhancement in the concrete pipe ductility, flexibility, and load transfer to the

surrounding soil while maintaining residual strength under deflection. Like flexible pipes,

fiber reinforced concrete pipes may be able to take advantage of interaction with the

surrounding soil to carry part of the applied load with less sensitivity to the installation

process. Compared to plastic and metal pipes, the fiber reinforced concrete pipes would

have more stiffness, have a simple design, be lightweight, be less sensitive to backfill

compaction and cost less to install. Moreover, fiber reinforced concrete pipes are very

strong in compression, which minimizes the possibility of thrust failure. However, most

researchers focused on using the same wall thickness specified by ASTM C76 (2015).

Decreasing the wall thickness of the concrete pipe along with adding an optimum fiber

dosage may result in a significant enhancement in the concrete pipe ductility, flexibility,

and load transfer between the pipe and surrounding soil, and maintain residual strength

under deflection.

2.3.3 Long-term Performance of Synthetic Fiber Reinforced Concrete

The problems associated with concrete pipe rigidity, shear failure, and radial

failure can be mitigated using synthetic fiber to supplement or replace the conventional

steel reinforcement (Park et al., 2016; Peyvandi et al., 2013 and 2014; Wilson and

Abolmaali, 2014; de la Fuente et al, 2013). For decades, synthetic fibers have been found

to enhance the mechanical properties of concrete, including flexural strength, tensile


34

strength, shear strength, ductility, shrinkage, and creep. The size of this improvement

depends on the dosage, geometry, dispersion, and size of fiber (Altoubat et al., 2009;

Balaguru and Shah, 1992; Çavdar, 2013 and 2014; Kuder and shah, 2010).

The understanding of the structural performance of synthetic fiber reinforced

concrete pipe is still in its infancy, and no design methodology currently exists to

correlate short-term tests with long-term field performance criteria for specific

engineering designs and manufacturing fiber dosage and distribution. In the literature,

there is a clear consensus that incorporating synthetic fiber enhances the short-term

structural performance of concrete pipes. However, there is uncertainty about the long-

term behavior of fiber reinforced concrete pipes under sustained load, as they could

exhibit increased deformation over time due to the viscoelastic properties of the synthetic

fiber; past research has yielded conflicting results. Park et al. (2014) performed an

experimental study to investigate the structural performance of synthetic fiber reinforced

concrete pipes under short-term and long-term loads. For the short-term test, concrete

pipes with a diameter of 600 and 900 mm (24 and 36 in. [lass III Wall-B]) were tested

under a three-edge bearing load. The pipes were reinforced with fiber dosages of 3.5, 4.8,

and 7 kg/m3 (5, 7.5, and 10 lb/yd3) plus one steel cage layer. Fiber dosages 4.8 and 7

kg/m3 along with one steel layer achieved the strength requirements specified by ASTM

Class III for 600- and 900-mm (24 and 36 in.) diameter pipe, respectively. The pipes

were then tested under sustained load by applying a cracking load, defined as a load

producing a crack 0.3 mm (0.01 in.) wide. The fibers hold the cracks from expanding and

prevented the pipe from collapsing. Pujadas et al. (2017) studied the long-term behavior
35

of pre-cracked concrete prisms reinforced with polypropylene (PP) fiber. Under sustained

load, the prisms exhibited large deformation and cracks in comparison with those

reinforced with steel fiber. The investigators recommend using steel reinforcement along

with PP fiber for concrete reinforcement as there is still uncertainty about the long-term

behavior of PP fiber reinforced concrete. Babafemi and Boshoff (2013, 2015, and 2016)

performed an experimental study to investigate the long-term performance of PP fiber

prisms under uniaxial tensile and flexural loading. Researchers reported that specimens

showed increases in the crack widths due to fiber creep. Also, they found that the

increase in the crack widths under flexural load was much less than that under uniaxial

tensile load. Bernard (2010) studied the post-crack creep characteristics of a shotcrete

panel reinforced with synthetic and steel fibers. The researchers found specimens

reinforced with synthetic fiber showed higher creep deformation than those reinforced

with steel fiber. Also, they reported that creep deformation of synthetic fiber reinforced

concrete specimens depends on many factors, including fiber content, fiber embossment

along their length, method of manufacture, and applied load value. Mackay and Trottier

(2004) conducted an experimental study to evaluate the post-crack creep behavior of

concrete prisms reinforced with synthetic and steel fibers and found there was not a big

difference in the creep deflection between synthetic and steel fiber reinforced concrete

prisms. They concluded that synthetic fiber could be used as a replacement for steel

reinforcement without adverse effect due to fiber creep. Boshoff et al (2009) performed

an experimental study to evaluate the behavior of the tensile creep of synthetic fiber

reinforced pre-cracked specimens, single fiber pull-out creep embedded in the concrete
36

matrix, and single fiber creep under different load levels. Researchers reported that

synthetic fiber did not contribute to the resistance to tensile creep of pre-cracked

specimens. Also, they concluded that the increase in the tensile creep is due to the

initiation of new cracks and the increase in crack widths caused by the fiber pull-out over

time.
37

CHAPTER 3 : METHODOLOGY

The laboratory tests of the fiber reinforced concrete material in the laboratory are

described, including the apparatus, materials, specimen preparations, and the test

procedures. Also, the laboratory tests of concrete pipes reinforced with different fiber

dosages under short and long term loading are described, including materials, pipes

production, curing, testing, instrumentation, and data collection. Then, a three-

dimensional finite element model was created along with different fiber contents and

steel cage areas to create design tables as those introduced by ASTM C76.

3.1 Material Properties of Synthetic Fiber Reinforced Concrete under Freeze-Thaw

Conditions

3.1.1 Materials

The properties of synthetic fiber used in this study are listed in Table 1. The PP

and PVA fiber have different dimensions (length and diameter) and material properties,

which affect the bond area between concrete and fiber. The tensile strength and Young’s

modulus are higher for PVA than PP. Also, both PP and PVA have a different CTE than

conventional concrete (ACI 209R-92, 2008).


38

Table 1. Physical Properties of the Fiber.

Properties Polypropylene fiber (PP) Polyvinyl alcohol fiber (PVA)

Specific Gravity 0.91 1.3

Length, mm (in.) 54 (2.1) 31.75 (1.25)

Diameter, μm (in.) 806 (0.032) 660 (0.026)

Tensile Strength, N/mm2 585 (85) 800 (116)

Young modulus, N/mm2 4000 (580) 23000 (3336)

Chemical Resistance Excellent Excellent

Coefficient of thermal expansion, 10-6 /°C 70 to100a (38.9 to 55.5) 100b (55.5)

(/°F)
a (Tripathi, 2002), b (Mark, 2009).

This research compared plain concrete to concrete with PP and PVA fibers at

concentrations of 6, 7, and 9 kg/m3( 8, 10, 12, and 15 lb/yd3). The fibers were mixed with

other concrete constituents using the mix design in Table 2. Air entrainment was used to

produce 2.5% air content. The cement included 30% Fly ash. The aggregate consisted of

North Carolina 78M stone, which was crushed granite with most pieces ranging between

9 to 12 mm long. The water-cement ratio used in this experiment was 0.32.

Table 2. Mix Proportions of the Concrete.

Cement, Fly ash Fly ash, Fine Coarse Water, PP fiber Water-reducing

kg/m3 (lb/yd3) (%) kg/m3 aggregate, aggregate, kg/m3(lb/yd3 kg/m3(lb/y admixture mL

(lb/yd3) kg/m3(lb/yd3 kg/m3(lb/yd3) ) d3) (gal)

309 (520) 30% 92.7 (156) 1129 (1903) 752 (1268) 128 (216) 9 (15) 2,235 (0.6)
39

3.1.2 Specimen Preparations

According to ASTM C1609/C1609M-15 (2015b), ASTM C215-14 (2014), and

ASTM C666/C666M-15 (2015a) test protocols, prism-shaped specimens with nominal

size 75×100×405mm (3×4×16 in.) were prepared to measure modulus of rupture,

dynamic modulus of elasticity, and loss of mass due to freeze-thaw cycles. For the CTE

tests, two different specimen sizes were used: a) concrete disks with 150 mm (6 in.)

diameter and 38 mm (1.5 in.) thickness for the Ohio CTE test method (see Figure 3 ) and

b) concrete cylinders with nominal diameter 100 mm and 200 mm height for AASHTO

specification AASHTO TP60-00 (2007). All specimens were poured on three separate

days and were steam cured.

Figure 3. Ohio CTE specimen’s preparation.


40

3.1.3 Test procedure

Coefficient of Thermal Expansion

The CTE values of the concrete specimens were measured using the Ohio CTE

test method reported by Akentuna et al. (2017) and Kim et al. (2015) and using the

AASHTO TP60-00 (2007) standard. The Ohio CTE device shown in Figure 4 was

calibrated using a cross square invar validated with 6061 aluminum and 316 stainless

steel disks. After drying test specimens at 60°C (140 ⁰F) for 48 hours, to ensure no effects

from moisture content, the specimen was placed in the chamber and held at 60°C (140 ⁰F)

for one hour. The temperature was then decreased linearly at 20°C/hr (68 ⁰F/hr), as shown

Figure 5 until reaching a minimum of -60°C (-76 ⁰F) after six hours, at which point the

temperature was held constant for an hour. Deformation and temperature data were

collected every 60 seconds.


41

(a) (b)

Figure 4. Ohio CTE device: (a) schematic device details; and (b) specimen installed

inside the chamber.

80
60
40
Temprature (°C)

20
0
-20
-40
-60
-80
0 2 4 6 8
Time (hour)

Figure 5. Temperature profile.

Figure 6 shows the test setup for the AASHTO TP60-00 (2007) test.

The AASHTO test device was calibrated using a 304 stainless steel cylinder. The

concrete specimens were saturated for 48 hours before testing in a water bath. The length
42

of each specimen was measured three times, and the average value recorded. The test

specimen was placed in the frame with the LVDT at the center of the specimen. Then, the

frame with the specimen was placed in the controlled temperature bath to maintain the

full saturated condition during the test. The temperature profile used in this test is shown

in Figure 7. The temperature was set to 10 °C for one-half hour, the temperature then was

raised from 10°C (50 ⁰F) to 50°C (122 ⁰F) over 2 hours, and then kept constant one-half

hour.

(a) (b)

Figure 6. AASHTO-CTE test setup: (a) schematic a mearing setup; and (b) specimen

setup.
43
60

50

Temprature (°C)
40

30

20

10

0
0 1 2 3 4 5
Time (hour)

Figure 7. Temperature profile used for AASHTO TP60-00 CTE test method.

Dynamic Modulus of Elasticity

The dynamic modulus of elasticity was measured according to ASTM C215-14

(2014). Determining the dynamic modulus of elasticity requires measuring the mass and

fundamental transverse frequency of the test specimens. A piezoelectric accelerometer

having a frequency range from 100 to 10000 Hz was attached to the concrete prism using

a special base glued at 25 mm (1 in.) distance from the end of the test specimen, as shown

in the schematic in Figure 8. The piezoelectric accelerometer was connected to the high-

speed data acquisition with the sampling rate of 65000 samples/sec. A special hammer

with a head weight of 23 g (0.05 lb) was used to vibrate the specimens. Data collected

from the accelerometer was analyzed using the Fourier series method to calculate the

fundamental transverse frequency. Equation 2 was used to calculate the dynamic of

elasticity.

Dynamic E = 0.9464 (L3T/bt3) Mn2 (2)


44

where M is mass of specimen (kg), n is fundamental transverse frequency (Hz), L

= length of specimen (m), t, b = dimensions of cross section of prism (m), and T is the

correction factor.

Figure 8. Dynamic modulus of the elasticity test setup.

Flexural Strength

The flexural performance was evaluated by measuring the modulus of rupture,

deflection at mid-span, and post-cracking behavior in accordance with ASTM

C78/C78M-15 (2015b). The deflection was measured at the mid-span using an LVDT

installed on the loading plate, as shown in Figure 9. A displacement rate of 0.1 mm/min

was employed in this test.


45

Figure 9. Flexural test setup.

3.1.4 Freeze-thaw Procedure

The Freeze-thaw test was completed in accordance with ASTM C666/666M-15

(2015a) Procedure A and B. The temperature was varied between 4°C to -18°C (39 to 0

F), with the cycle time about 4 to 5 hours. For procedure A, a freeze-thaw cabinet

(Humboldt MFG. Co. DBA, Logan Freeze-Thaw Mfg. Co. UT) was used to apply cycles

of freezing and thawing, as shown in Figure 10. This cabinet has 18 stainless steel prism

molds, and one mold was used to carry the pilot specimen with the control bulb of the

control cycle and thermometers. Also, two thermocouples were installed on the surface,

and the center of the pilot specimen to monitor the temperature at these locations and

ensure the difference did not exceed 10 °C (50 ⁰F). S-shaped brass wires were placed

under all specimens to allow water movement underneath and prevented rapid heat-

transfer between the specimen and the bottom surface of the mold, which sits on the

freezing plate of the test machine. The resistance of fiber reinforced concrete to the

freeze-thaw cycles was evaluated by measuring the dynamic modulus of elasticity, loss in
46

mass, and flexural strength. The change in the dynamic modulus of elasticity and loss of

mass due freeze thaw cycles was obtained every 30 cycles based on ASTM C666/666M-

15 (2015a) standard, while flexural strength was evaluated after 100, 200, and 300 freeze-

thaw cycles. All specimens were soaked in the water bath for 14 days prior to the test,

and the mass measured with a precision of 0.001 kg in accordance with ASTM

C666/666M-15 (2015a).

Figure 10. Freeze-thaw cabinet.

ASTM C666/666M (2015a) Procedure B was employed for CTE specimens since

CTE specimens did not fit at freeze-thaw cabinet molds. The CTE specimens were frozen

using a different chamber (Cincinnati Sub-Zero products Inc. Mosteller, Cincinnati,

Ohio) and thawed using water. The CTEs were calculated after 0, 100, 200, and 300

freeze-thaw cycles. All the CTE test specimens were air-dried at 60°C (140 ⁰F) for 48

hours before testing since the moisture level of the test specimens has an impact on the

CTE value (Sellevold and Bjøntegaard 2006; Al-Ostaz 2007; ACI 209R-92, 2008).
47

3.2 Evaluation of Synthetic Fiber Reinforced Concrete Pipe Performance Using Three-

Edge Bearing Test

3.2.1 Experimental Procedure

The three-edge bearing test was conducted in accordance with ASTM C497

(2014) on fiber-reinforced concrete pipes with diameters, 600, 1200, and 1500 mm (24,

48, 60 in.). Polypropylene (PP) synthetic fiber with dosages of 4.75, 9, 13.5, and 18

kg/m3 ( 8, 15, 22.5, and 30 lb/yd3) were used. The fiber-reinforced concrete pipes tested

included two types: those with a minimum amount of steel reinforcing cage embedded in

a single layer augmenting the fiber reinforcement and those with only the fiber

reinforcement and no steel. The performance of the fiber reinforced concrete pipe system

was measured in terms of strength requirements (ASTM C76 (2015)), ductility, stiffness,

and post-cracking behavior.

3.2.2 Material Properties

The synthetic fiber used in this study was produced from a blend of 100% virgin

polypropylene resins. This fiber type is monofilament with an embossed surface. The

distance from the embossed peak to the lowest point is from 0.085 to 0.095 mm (0.0033

to 0.037 in.). This fiber has a length of 54 mm (2.1 in.) with an equivalent diameter of

0.82 mm (0.032 in.). The maximum tensile strength that can be carried by this fiber is

585 MPa. The mechanical and physical properties of synthetic fiber used in this study are

listed in Table 1.

Four different dosages of PP fiber, 4.75, 9, 13.50, and 18 kg/m 3(8, 15, 22.5, and

30 lb/yd3), were used. Each fiber dosage was mixed with other concrete constituents
48

using the mix design listed in Table 3. Air concrete of 2.5 % was assured in the mix by

using air-entraining additives. The coarse aggregate used in this study was crashed

granite North Carolina 78M stone. Most pieces of aggregate ranged from 9 to 12 mm

(0.35 to 0.5 in.) long. A water to-cement ratio of 0.32 with zero slump was utilized in this

experiment.

Table 3. Mix Proportions of the Concrete

Cement, Fly ash Fly ash Fine aggregate Coarse aggregate Water

kg/m3 (lb/yd3) (%) kg/m3 (lb/yd3) kg/m3 (lb/yd3) kg/m3 (lb/yd3) kg/m3 (lb/yd3)

309 (520) 30% 92.7 (156) 1129 (1903) 752 (1267) 128 (215)

3.2.3 Specimen Preparation

The concrete pipes were constructed using zero slump mixture and Hawkeye

equipment. In this manufacturing procedure, cylindrical inner and outer metal jackets are

situated in a manner to form the outside and inside diameter of the concrete pipe. A steel

ring is then used to hold the reinforcement cage (see Figure 11(a)). This system is placed

over a vibrating table to allowed concrete material to be compacted during the pouring

process, as shown in Figure 11(b). The fresh compacted concrete with the jackets are

moved carefully to the curing environment. To expedite the curing process, the outside

jacket is removed, and the exposed surface of the pipe is subject to a stem, as shown

Figure 11 (c). Thirty-six concrete pipes of the diameter of 600 mm (24 in.) Wall-B,

1200mm Wall-B, and 1500 mm Wall-C were tested. All pipes were 2.4 m long, except

some 1200 mm (48 in.) diameter pipes were 1.2 m long. Steel reinforcement, when
49

present, was placed in a single layer at 25 mm (1 in.) from the inside diameter. Details of

the pipes used in this study are listed in Table 3.

Table 4. Details of the Tested Fiber Reinforced Concrete Pipe Specimens.

Series no. Inside Wall Fiber dosage kg/m3 Steel cage


Number of
diameter, Thickness, mm (lb/yd3): volume cm2/linear m
pipes tested
mm (in.) (in.) fraction (%) (in2/ft)

1 600 (24) 75 (3) 9 (15):1 0 2

9(15):1 1.5 (0.07) 2

18(30):2 0 2

18(30):2 1.5(0.07) 1

2 1200 (48) 125 (5) 0:0 0 3

4.75(8):0.5 0 3

4.75(8):0.5 5.1 (0.24) 2

9(15):1 0 3

9(15):1 5.1 (0.24) 3

13.5(22.5):1.5 5.1(0.24) 3

18(30):2 5.1(0.24) 3

3 1500 (60) 169 (6.5) 9(15):1 5.1(0.24) 3

13.5(22.5):1.5 5.1(0.24) 3

18(30):2 5.1(0.24) 3
50

(a) (b) (c)

Figure 11. Pipe production: (a) steel cage reinforcement, (b) mixture, and (c) removal of

the jacket.

3.2.4 Test Procedure

The hydraulic testing machine shown in Figure 12 (a) was used to perform the

three-edge bearing test. The test pipe was supported using two incompressible plastic

strips extended along its full length. In order to assure that the applied load is uniformly

distributed along the pipe length, a hard rubber strip fastened to a hard wood member was

used, as shown in Figure 12(a, and b). The crack was measured manually using Feeler

gauge, as shown in Figure 12(c). The deflection was measured at three different locations

along the pipe length using three string potentiometers. The hydraulic testing equipment

and displacement transducers were connected to a computer-based data acquisition

system to record the applied load and the vertical displacement. In order to compare the

pipe strength results with the strength requirements specified by ASTM C76 (2015), the

measurements of the load corresponding to 0.3 mm (0.01 in.) crack width (called the

cracking load) and the ultimate load were divided by the inside diameter of the pipe and

the length of the pipe. The strength requirements specified by ASTM C76 (2015) are
51

presented in Table 5. Based on the three-edge bearing test results, the classification of the

concrete pipe (Class I to Class V) was determined.

(a) (b) (c)

Figure 12. a) Specimen setup; b) schematic three-edge bearing test setup; c) crack width

indicator.

Table 5. ASTM C76 (2015) Pipe Class, Cracking Load, and Ultimate Load

ASTM C76 class Cracking load, Ultimate load,

kN/m/m (kip/ft/ft) kN/m/m (kip/ft/ft)

I 40 (0.8) 60 (1.2)

II 50 (1) 70 (1.5)

III 65 (1.35) 100 (2)

IV 100 (2) 150 (3)

V 140 (3) 175 (3.75)


52

3.3 Design Proposal for Synthetic Fiber Reinforced Concrete Pipes Using Finite

Element Analysis

3.3.1 Fiber Reinforced Concrete Pipes

Full-scaled reinforced concrete pipes by PP fiber were tested under three edge

bearing tests. The pipes with a diameter of 600, 1200, and 1500 mm (24, 48, and 60 in.)

were tested in two categories. In the first category, the pipes were reinforced with only

PP fiber. In the second category, the pipes were also reinforced with a single-layer steel

cage. The details or pipes tested in previous research by the authors (Al Rikabi et al.

2018) are listed in Table 6. The following fiber dosages were used to replace the steel

cage reinforcement in traditional concrete pipes: 4.75, 9, 13.5 and 18 kg/m 3, ( 8, 15, 22.5,

and 30 lb/yd3) equivalent to 0.52, 1, 1.5, and 2 % of concrete volume (Vf).


53

Table 6. Details of the Tested Specimens.

Series Inside Wall Fiber dosage, Steel cage Number

no. diameter, Thickness, kg/m3 (lb/yd3): volume cm2/linear m of pipes

mm (in.) mm (in.) fraction (%) (in2/ft) tested

1 600 (24) 75 (3) 9 (15):1 0 2

9(15):1 1.5 (0.07) 2

18(30):2 0 2

18(30):2 1.5(0.07) 1

2 1200 (48) 125 (5) 0:0 0 3

4.75(8):0.5 0 3

4.75(8):0.5 5.1 (0.24) 2

9(15):1 0 3

9(15):1 5.1 (0.24) 3

13.5(22.5):1.5 5.1(0.24) 3

18(30):2 5.1(0.24) 3

3 1500 (60) 169 (6.5) 9(15):1 5.1(0.24) 3

13.5(22.5):1.5 5.1(0.24) 3

18(30):2 5.1(0.24) 3

3.3.2 Materials

The fiber used in this study was 100% virgin PP resin monofilament with an

embossed surface. The physical and mechanical properties of this type of fiber are listed

in Table 6. This type of fiber has a length of 54 mm (2.1) and an equivalent diameter of

0.82 mm (0.032 in.). The maximum tensile strength of this type of fiber is 585 MPa (85

ksi). Different dosages of the fiber were mixed with other concrete components using the
54

mix design listed in Table 3, a zero-slump mix with water to cement ratio of 0.32. Air

entraining additives were used to assure 2.5 % of air content in the concrete mix. Crushed

granite North Carolina 78M stone of length 9 to 12 mm was used as coarse aggregate.

Figure 13. Polypropylene (PP) fiber used in this research.

Table 7. Physical and mechanical properties of the PP Fiber.

Properties Polypropylene PP) fiber

Specific gravity 0.91

Length, mm (in.) 54 (2)

Equivalent diameter, μm (in.) 820 (0.032)

Tensile strength, N/mm2 (ksi) 585 (85 )

Young’s modulus, N/mm2 (ksi) 4000 (580)

Chemical resistance Excellent


55

Table 8. Mix proportions of the concrete used to make the fiber-reinforced pipes.

Cement, Fly ash Fly ash Fine aggregate Coarse aggregate Water

kg/m3 (lb/yd3) (%) kg/m3 (lb/yd3) kg/m3 (lb/yd3) kg/m3 (lb/yd3) kg/m3 (lb/yd3)

309 (520) 30% 92.7 (156) 1129 (1903) 752 (1267) 128 (215)

Measurements were made on the compressive strength, modulus of elasticity,

Poisson’s ratio, and tensile strength of the fiber reinforced concrete threat fiber dosages

of 0 (plain concrete), 4.75, 6, 7, and 9 kg/m3( 0, 8, 10, 12, and 15 lb/yd3). Compressive

strength and modulus of elasticity were obtained in accordance to ASTM C39 (2015) and

ASTM C469 (2015), respectively, as shown in Fig. 3(a). To obtain Poisson’s ratio, two

strain gauges were affixed horizontally and vertically on the concrete cylinder, as shown

in Fig. 3(b). Tensile strength was obtained in accordance with ASTM C496 (2015), as

depicted in Fig. 3(c).

(a) (b) (c)

Figure 14. Test setup of (a) Compressive strength (b) Poisson’s ratio, and (c) splitting

tensile strength.
56

3.3.3 Finite Element Model of the Three-edge Bearing Test

A three-dimensional finite element model of the three-edge bearing test was

created using finite element software ABAQUS/CAE Version 6.12-3. This model

consisted of five components: concrete pipe, steel reinforcement cage, two lower strips,

upper strip, and rigid member, as shown in Figure 15. The upper and lower rigid

members had dimensions of 100×50 (4×2 in.) and 175×50 mm (3×2 in.), respectively.

The lower strips had cross-section dimensions of 50 ×25 mm each and 75 mm separation.

The pipe, rigid members, and rubber strips were modeled using 8-node linear solid

elements with reduced integration and hourglass control. The steel cage reinforcement

was modeled using 2-node linear truss elements (T3D2) constrained and embedded inside

the concrete pipe, as depicted in Figure 15 (a). The steel reinforcement cage was

distributed along the pipe length at a spacing of 63.5 mm (2.5 in.), as recommended by

the manufacturer. The steel reinforcement cage layer was placed 25 mm from the inside

pipe surface. The lower strips were attached to the concrete pipe and bottom rigid

member using tie constraints. The same constraint was used to attach the upper strip to

the upper rigid member and the concrete pipe. The upper rigid member was constrained

in the horizontal direction to restrict the load to the vertical direction. The lower rigid

member was fixed at the bottom. The load was uniformly applied using the upper rigid

member. Also, the pipe diameter and wall thickness varied in the parametric study.
57

(a) (b)

Figure 15. Finite element modeling of the three-edge bearing test. (a) A-A cross-section;

(b) 3D model.

3.3.4 Concrete Model

In the ABAQUS software, two constitutive models can be used to model the

concrete. These models are concrete smeared crack (CSC) and concrete damage plasticity

(CDP). The CSC model has convergence problems when the steel reaches the yield stage

(Ahn, 2011; Chen and Graybeal, 2011), and was therefore not used. The concrete damage

plasticity (CPD) model has the ability to model the nonlinear response of the concrete

material. CPD assumes that concrete material experiences two modes of failure: tensile

cracking and compressive crushing. In the latter, the concrete material experiences elastic

behavior until the yield point is reached; beyond that, the concrete exhibits strain
58

hardening followed by softening. For the tensile behavior, the tension response of the

concrete material can be represented using three types of modeling, stress-strain data,

stress crack mouth opening displacement data, and yield stress-fracture energy data.

Since the PP fibers start to contribute after the first crack, the stress crack mouth opening

displacement data would be the best model to represent the post cracking behavior of

fiber reinforced concrete under tension. In addition to the compression and tension

response models, CPD includes five parameters: the dilation angle in degrees (ѱ, default

36.67°), the flow potential eccentricity (ϵ, default 0.1), the ratio of the second stress

invariant on the tensile meridian to that on the compressive meridian such that the

maximum principal stress is negative (Kc, default 0.67), a viscosity parameter (μ, 0.001),

and the ratio of the initial equibiaxial compressive yield stress to the initial uniaxial

compressive yield stress (σb0/σc0, default 1.16).

3.3.5 Uniaxial Modelling of Fiber Reinforced Concrete Compression Behavior

Several researchers have introduced analytical models to describe the

compressive strength behavior of fiber-reinforced concrete (Soroushian and Lee 1989;

Ezeldin and Balaguru 1992; Barros and Figueiras 1999; Mansur et al. 1999; Nataraja et

al. 1999). However, most of the research adapted plain concrete models by adding a

dimensionless factor to the compressive strength and ultimate strain. This dimensionless

factor was obtained based on how incorporating steel fiber increases the compressive

strength of plain concrete. However, past research showed that using synthetic fiber

reduces the compressive strength of concrete as fiber dosage is increased. This is

attributed to the high ductility of PP fiber compared to that of other concrete constituents,
59

which may create a discontinuity in the concrete matrix and result in a reduction of

compressive strength (Zhang et al., 1999; Huang, 2001; Puertas et al., 2003; Choi and

Yuan, 2005, Çavdar 2014).

3.3.6 Uniaxial Modelling of Fiber Reinforced Concrete Tensile Behavior

The contribution of concrete matrix dominates the tensile strength until the first

crack and diminishes in the post cracking region as fibers start to dominate the response.

The combined response of matrix and fiber under tensile load was described by Voo and

Foster (2003) using stress-crack opening displacement relationship, as shown in Figure

16.

Figure 16. Tensile stresses vs. crack mouth opening displacement for fibers, matrix, and

fiber-matrix composite.

This research modeled the post-cracking behavior of PP fiber reinforced concrete

after its ultimate tensile strength was reached using fracture energy cracking criteria. The

model was defined by providing the ABAQUS with tensile stress-crack opening
60

displacement values. The tension behavior of fiber-reinforced concrete followed the

Variable Engagement Model proposed by Voo and Foster (2003). In this model, for a

given crack mouth opening displacement (w), the composite tensile strength (σ(w)) is

equal to the sum of stresses carried by the concrete matrix (σc(w)) and stresses (σf (w))

carried by fibers, as represented in equation (1).

𝜎(w)=𝜎𝑐 (𝑤) + 𝜎𝑓 (𝑤) Equation 1

The softening behavior of plain concrete beyond its maximum tensile strength can

be expressed using equation with

𝜎𝑐 (w)=𝑐1 𝑓𝑐𝑡 𝑒 𝑐2𝑤

where fct is the tensile strength of plain concrete, c1 is a parameter representing

any benefit of fibers on the maximum tensile strength of plain concrete (typically c1=1),

and c2 is a parameter controlling the steepness of the softening behavior of plain concrete.

The c2 parameter is mainly affected by fiber volume (Vf) and other concrete constituents

(Voo and Foster, 2003 and 2004; Htut, and Foster 2012; Mohamed et al., 2016). Ng et al.

(2012) proposed to calculate c2 for concrete with maximum aggregate size of 10 mm

using the following equation:


20
𝑐2 = (1+100𝑉 Equation 2
𝑓)

The stresses (σf (w)) carried by the fibers can be calculated using the following

expression (Voo and Foster, 2003 and 2004; Htut, and Foster 2012; Mohamed et al.,

2016).

1 αe∙w 2∙w 2 l
σf (w) = π tan−1 ( ) (1 − ) ∙ df ∙ τb ∙ Vf Equation 3
df lf f
61

where τb is bond strength between fiber and concrete components, and α is the

engagement factor. The bond strength of single fiber (τ) for the synthetic fiber can be

calculated using the following expression (Wang et al., 1987):

df σuf
τ=
4Lc

where σuf is the maximum fiber tensile strength, and Lc is the maximum length of

fiber that can be pulled out without experiencing rupture. According to Li et al. (1990), in

order for fiber reinforcing effects on the concrete behavior to be optimum, fiber length

should twice the Lc. Therefore, Lc was assumed to be half the fiber length. Figure 17

shows the tension stress-crack opening curves used in the finite (FE) element models

based on a model proposed by Voo and Foster.

5
6 kg/m3 7 kg/m3
9 kg/m3 13.5 kg/m3
4 18 kg/m3
Tensile stress (MPa)

0
0 1 2 3 4 5 6 7 8 9 10
CMOD (mm)

Figure 17. Tensile stresses-crack mouth opening displacement plots for PP reinforced

concrete material used in the finite element models.


62

3.3.7 Other Models

In order to simulate the steel cage so that it reaches its ultimate strength, the steel

cage was defined in terms of elastic and plastic behavior. The elastic behavior was

defined using the Poisson’s ratio and modulus of elasticity. The plastic behavior was

modeled using a stress-strain (f-ε) relation proposed by Mirza and MacGregor (1981) for

deformed and plain reinforcing wire, which is used as steel reinforcement in concrete

pipes. This relationship was expressed using the following equation:

𝑓 𝑓 𝑓 20
𝜀 = 𝐸 + (𝜀𝑢 − 𝐸 ) (𝑓 ) Equation 4
𝑠 𝑠 𝑢

where Es, εu, and fu are modulus of elasticity, ultimate strain, and ultimate stress of

plain reinforcing wire, respectively. The ultimate strain was calculated using the

following expression:

1/2
𝜀𝑢 = 0.0042 An

where An is the nominal area of the wire reinforcing in (mm2).

The ultimate stress results of different wire reinforcing sizes were provided by the

concrete pipe manufacturer and listed in Table 9. The upper and lower strips in the

experiment were modeled as hard rubber with a modulus of elasticity of 4350 MPa (630

ksi) and Poisson’s ratio of 0.4 (AASHTTO LRFD, 2014). The rigid steel members of the

apparatus were modeled with a modulus of elasticity of 200 GPa (29000 ksi) and

Poisson’s ratio of 0.3. The elastic properties of concrete material were defined based on

the experimental results of this study.


63

Table 9. Testing results of wire reinforcing used for finite element models of the concrete

pipes.

Diameter, Area, Ultimate stress,


Wire size
mm (in.) mm (in )
2 2
MPa (ksi)
W2.5 4.53 (0.17) 16.10 (0.025) 672.2 (97.5)
W3 4.95 (0.19) 19.35 (0.030) 620.7 (90.0)
W3.5 5.36 (0.21) 22.60 (0.035) 633 (91.8)
W4.5 6.08 (0.24) 29.00 (0.045) 563.5 (81.7)
W7 7.59 (0.30) 45.16 (0.070) 698.2 (101.2)
W8 8.11 (0.32) 51.6 (0.080) 607.6 (88.1)

3.4 Thin Walled Synthetic Fiber Reinforced Concrete Pipes

3.4.1 Materials

Monofilament polypropylene (PP) fiber used as a replacement for steel reinforcement

cage in the concrete pipes is 100 virgin polypropylene resins, depicted in Fig. 3. This type

of fiber has a length of 54 mm (2.1 in.) and equivalent diameter of 0.82 mm. The physical

and mechanical properties of this type of fiber are presented in Table 10. A fiber dosage

of 9 kg/m3 was mixed with other concrete constituents using the mix design illustrated in

Table 11. Air entrained admixture with a dosage of 28 ml/m3 was used to produce

air content of 2.5% in the mixture. Crushed granite with length ranging from 9 to 12 mm

was used as a coarse aggregate. It should be noted that a self-consolidating concrete

mixture was used for this application due to the thickness of the wall forms. This mix is

typical of a precast concrete mix, which uses high water reducing admixtures to maintain

a very low water-cement ratio. The result is a high strength concrete with a relatively

water content when compared to standard wet-cast ready-mix concrete


64

Table 10. Physical Properties of the Synthetic Fiber.

Properties Polypropylene PP) fiber

Specific gravity 0.91

Length, mm (in.) 54 (2)

Equivalent diameter, μm (in.) 820 (0.032)

Tensile strength, N/mm2 (ksi) 585 (85 )

Young’s modulus, N/mm2 (ksi) 4000 (580)

Chemical resistance Excellent

Table 11. Mix Proportions of the Concrete Mixtures.

Cement, Fly ash Fly ash, Fine Coarse Water, PP fiber Water-reducing

kg/m3 (%) kg/m3 aggregate, aggregate, kg/m3 kg/m3 admixture

(lb/yd3) (lb/yd3) kg/m3 kg/m3(lb/yd3) (lb/yd3) (lb/yd3) mL (gal)

(lb/yd3)

309 (520) 30% 92.7 (156) 1129 752 (1268) 128 (216) 9 (15) 2,235 (0.6)

(1903)

3.4.2 Specimen Preparation

All concrete pipes were produced using the wet cast method. A cylindrical outer

jacket forms the outside diameter of the concrete pipe. A steel ring was used to hold the

steel cage in place inside the cylindrical outer jacket. The concrete with the selected fiber

dosage was poured into the pipe mold using a cement mixer truck. Two concrete pipes

with a diameter of 1200 mm (48 in.), wall thickness 50 mm (2 in.), and length 1.22 m (4

ft.) were fabricated. Both pipes were reinforced with a 9 kg/m 3 (15 lb/yd3) dosage of PP
65

fiber. There was one layer of steel cage reinforcement located 25 mm (1 in.) from the

inside diameter of each pipe; one pipe had steel reinforcement at 5.7 cm2/m (0.27 in2/ft)

and the other had 10.2 cm2/m (0.48 in2/ft). Two pipes were prepared with a diameter of

1500 mm (60 in.), wall thickness 63 mm (2.5 in.), and length 1.2 m. The fiber dosage and

steel reinforcement details were the same as for the smaller pipes, except the steel content

in the pipes was 5.7 cm2/m and 8.9 cm2/m (0.27 and 0.42 in2/ft). The details of the pipes

used in this study are summarized in Table 12.

Table 12. Details of the Tested Specimens.

Fiber dosage kg/m3


Inside diameter, mm Wall Thickness, mm Steel cage,
(lb/yd3): volume
(in.) (in.) cm2/m (in2/ft)
fraction (%)
1200 (48) 50 (2) 9 (15):1 5.7 (0.27)

9 (15):1 10.2 (0.48)

1500 (60) 64 (2.5) 9 (15):1 5.7 (0.27)

9 (15):1 8.9 (0.42)

3.4.3 Test Procedure

The three-edge bearing test was conducted using the apparatus shown in Figure

18(a) in accordance with ASTM C497 (2015). The load was applied uniformly using a

hard rubber strip affixed to a hardwood member at the crown of the pipe. The pipe was

supported using two hard rubber strips fastened to a steel member. The pipes were

instrumented with two linear string potentiometers to measure the vertical displacement

at the crown, and two linear string potentiometers to measure the horizontal displacement
66

at the spring line. Surface strains for the pipe were measured using uniaxial electrical

resistance gauges with a length of 60 mm (2.36 in.). Four of these strain gauges,

designated SG#1, SG#2, SG#3, and SG#4, were affixed on the concrete pipe wall at the

crown, invert, and spring line, as shown in Figure 18 (b). The load was applied at a rate

of 43.8 kN/linear meter of pipe per minute (3000 lbf/linear foot of pipe per minute) using

a hydraulic testing machine. The data were recorded using a computer-based data

acquisition system. The strength results were compared with those specified by ASTM

C76 (2015) to determine the class of tested pipe, as indicated in Table 13.

Table 13. ASTM C76 (2015) Class Criteria for Cracking Load and Ultimate Load

ASTM C76 class Cracking load, Ultimate load,

kN/m/m (kip/ft/ft) kN/m/m (kip/ft/ft)

I 40 (0.8) 60 (1.2)

II 50 (1) 70 (1.5)

III 65 (1.35) 100 (2)

IV 100 (2) 150 (3)

V 140 (3) 175 (3.75)


67

(a) (b)

Figure 18. a) Specimen setup; b) schematic three-edge bearing test setup.

3.5 A New Test Method for Evaluating the Long-Term Performance of Fiber

Reinforced Concrete Pipes

3.5.1 Finite Element Model of Long-Term Test Setup

A three-dimensional finite element model was created using FE software Abaqus

Version 6.14-3 to simulate the long-term test setup. This model consists of four

components: steel frame, concrete pipe, wood strips, and steel reinforcement cages

embedded in the concrete pipe wall, as shown in Figure 19. The steel frame comprises of

a steel column, steel angles used as column bracing, a cantilever beam used for applying

the sustained load, a pin connection connecting the steel column and the cantilever beam,

a loading beam used to apply the load along pipe length, and a lower beam used as a

support for the concrete pipe. The pin connection was made of semicircular steel plates
68

with a 1 in thickness connected using a 3 in diameter steel rode. To measure the applied

load and avoid eccentricity, a circular steel plate with a thickness of 3 in and semicircular

head, representing the load cell, was placed between the loading and cantilever beams.

The wood strips were affixed to the upper and lower beams as specified by ASTM C497

(2015b). The steel frame, wood strips, and concrete pipe were modeled using 8-node

linear solid elements with reduced integration and hourglass control. The steel

reinforcement cage embedded in the concrete pipe was modeled as a 2-node truss

element (T3D2) constrained and embedded inside the concrete pipe. The upper and lower

wood strips were attached to the concrete pipe using the surface to surface tie constraint.

For the pin connection, the surface to surface interaction was utilized to model the

interface between steel rode and surrounding hole face. The interaction properties were

defined using tangential behavior. This behavior was characterized using penalty friction,

with a friction coefficient of 0.7 (Sullivan, 1988). Also, the interface between the load

cell and cantilever beam was defined in terms of normal and tangential behavior. The

normal behavior was modeled as hard contact, in which no penetration between the load

cell and cantilever beam could occur. Penalty fraction was assigned to the tangential

behavior with a friction coefficient of 0.7 (Sullivan, 1988).


69

Figure 19. FE model of the long-term test setup

3.5.2 Fiber Reinforced Concrete Material Model

In the ABAQUS software, two constitutive models can be used to model the

concrete material. These models are concrete smeared crack (CSC) and concrete damage

plasticity (CDP). The CSC model has convergence problems when the steel reaches the

yield stage (Ahn, 2011; Chen, 2012) and was, therefore, not used. The concrete damage

plasticity (CDP) model has the ability to model the nonlinear response of the concrete

material. CDP assumes that concrete material experiences two modes of failure: tensile

cracking and compressive crushing. In the latter, the concrete material experiences elastic

behavior until the yield point is reached; beyond that, the concrete exhibits strain

hardening followed by softening. This model was defined using the compressive stress-
70

strain curve proposed in previous research by the authors (Al Rikabi et al., 2020), as

shown in Figure 20.

50
9 kg/m3
40

Stress (MPa) 30

20

10

0
0 0.002 0.004 0.006 0.008 0.01 0.012
Strain (mm/mm)

Figure 20. compressive stress-strain plots used in the FE models.

For the tensile behavior, the tension response of the concrete material can be

represented using three types of modeling, stress-strain data, stress crack mouth opening

displacement data, and yield stress-fracture energy data. Since the PP fibers start to

contribute after the first crack, the stress crack mouth opening displacement data would

be the best model to represent the post cracking behavior of fiber-reinforced concrete

under tension. The tension behavior of fiber-reinforced concrete followed the Variable

Engagement Model proposed by (Voo and Foster, 2003 and 2004), as shown in Figure

21.
71

5
9 kg/m3
4

Tensile stress (MPa)


3

0
0 1 2 3 4 5 6 7 8 9 10
CMOD (mm)

Figure 21. Tensile stresses-crack mouth opening displacement plots for PP reinforced

concrete material used in the finite element models.

In addition to the compression and tension response models, CDP includes five

parameters: the dilation angle in degrees (ѱ, default 36.67°), the flow potential

eccentricity (ϵ, default 0.1), the ratio of the second stress invariant on the tensile meridian

to that on the compressive meridian such that the maximum principal stress is negative

(Kc, default 0.67), a viscosity parameter (μ, 0.001), and the ratio of the initial equibiaxial

compressive yield stress to the initial uniaxial compressive yield stress (σb0/σc0, default

1.16).

3.5.3 Finite Element Results

A series of finite element models were developed to assess the state of stresses

within the pipe wall and other frame components. The models were created using

different steel member sections to ensure that the new test setup is easy to install, able to

transfer the load, complied with requirements specified by ASTM C497 (2015). Figure
72

22 shows the steel sections adopted for the experimental test procedure based on the FE

analysis. It can be seen even though large sections were selected for the column and

cantilever beam, steel plates were welded along their length to eliminate any section

rotation while applying the load.

Figure 23 shows the damage occurred ae the pipe wall due to the load attached to

the end of the cantilever beam. The failure occurred at the inner invert, inner crown, and

outer springline where the maximum flexural stresses were expected. The results herein

indicate that the new test setup successfully transferred the load to the pipe. Also, to

ensure that the steel frame will not influence the long-term pipe deflection, the

deformation at the steel members was investigated. According to ASTM C497 (2015),

the wood-faced steel beam of such dimensions that deflections under maximum load

should not be greater than L⁄720 of the specimen length. Figure 24 shows the vertical

displacement throughout the wood faced beam depth. It can be seen that the difference

between the displacement at the top and bottom of the beam is almost zero, fulfilling the

deflection requirement specified by ASTM C497.


73

Figure 22. Long-term test setup detail selected based on the FE analysis.

Figure 23 The damage in the concrete pipe due to the applied.


74

(a)

250

200
Steel beam depth (mm)

150

100

50

0
-12.5237 -12.5237 -12.5238 -12.5238 -12.5238 -12.5238
Vertical displacement

(b)
75

350

300

250

Steel beam depth (mm)


200

150

100

50

0
0.000000 -0.000010 -0.000020 -0.000030 -0.000040
Vertical displacement (mm)

(c)

Figure 24. The vertical displacement for (a) whole test setup (b) upper beam (c) bottom

beam.

3.6 Long-term Performance of Synthetic Fiber Reinforced Concrete

3.6.1 Material Properties

Monofilament polypropylene (PP) fiber used as a replacement for steel

reinforcement cage in the concrete pipes is 100 virgin polypropylene resins. This type of

fiber has a length of 54 mm and an equivalent diameter of 0.82 mm (0.032 in.). The

physical and mechanical properties of this type of fiber are presented in Table 14. A fiber

dosage of 9 kg/m3 (15 lb/yd3) was mixed with other concrete constituents using the mix

design illustrated in Table 15. Air entrained admixture with a dosage of 28 ml/m3 was

used to produce air content of 2.5% in the mixture. Crushed granite with length ranging

from 9 to 12 mm (0.35 to 0.5 in.) was used as a coarse aggregate. It should be noted that
76

a self-consolidating concrete mixture was used for this application due to the thickness of

the wall forms. This mix is typical of a precast concrete mix, which uses high water-

reducing admixtures to maintain a very low water-cement ratio.

Table 14. Physical Properties of the Synthetic Fiber.

Properties Polypropylene PP) fiber

Specific gravity 0.91

Length, mm (in.) 54 (2)

Equivalent diameter, μm (in.) 820 (0.032)

Tensile strength, N/mm2 (ksi) 585 (85 )

Young’s modulus, N/mm2 (ksi) 4000 (580)

Chemical resistance Excellent

Table 15. Mix Proportions of the Concrete Mixtures.

Cement, Fly ash Fly ash, Fine Coarse Water, PP fiber Water-reducing

kg/m3 (%) kg/m3 aggregate, aggregate, kg/m3 kg/m3 admixture

(lb/yd3) (lb/yd3) kg/m3 kg/m3 (lb/yd3) (lb/yd3) mL (gal)

(lb/yd3) (lb/yd3)

309 (520) 30% 92.7 (156) 1129 (1903) 752 (1268) 128 (216) 9 (15) 2,235 (0.6)

3.6.2 Specimen Preparation

A wet cast method was used to produce all concrete pipes. Cylindrical inner and

outer metal jackets were used to form the inside and outside diameter of the concrete

pipe, and a steel ring located at the end of the mold was used to hold the steel
77

reinforcement mesh in place. The concrete mixture, including selected fiber dosage was

poured in the pipe molds using a concrete mixing truck. After pouring, the pipes were

steam cured. Concrete pipes with a diameter of 1200 and 1500 mm (48 and 60 in.) with a

length of 1.2 m (4 ft) and wall thicknesses of 50 and 63 mm (2 and 2.5 in.) were

produced. All pipes had a PP fiber dosage 9 kg/m3 or 1% by volume, while the area of

steel reinforcement differed as shown in Table 16. The steel reinforcement consisted of a

single layer embedded 25 mm from the inside diameter. The details of the pipes used in

this study are presented in Table 16.

The PP fiber dosage was selected based on the experimental investigation that

indicated there was a performance versus manufacturing limit that must be balanced

when utilizing fibers. Once fiber content exceeds 2% by volume, it is not able to be

feasibly produced in a concrete pipe operation without the material balling-up and

blocking form distribution. On the other hand, fiber content less than 0.5% by volume

was too widely distributed to provide any tensile benefit. For these reasons, the optimal

economic-performance volume content of 1% was selected.

Table 16. Details of the Tested Pipe Specimens, each having length 1.2 m.

Inside diameter, Wall thickness, Fiber dosage, Steel mesh,


Pipe No.
mm (in.) mm (in.) kg/m (lb/yd )
3 3
cm2/m (in2/ft)
1 1200 (48) 50 (2) 9 (15) 5.7 (0.27)

2 9 (15) 10.2 (0.48)

3 1500 (60) 64 (2.5) 9 (15) 5.7 (0.27)

4 9 (15) 8.9 (0.42)


78

3.6.3 Test Procedure

The long-term test setup is shown in Figure 25. The pipe was supported at the

invert using two wood strips fastened to a rigid steel member. The load was applied along

the top of the pipe using a wood strip attached to a steel beam member. The sustained

load was applied using a concrete block attached to the end of the cantilever beam. The

load was applied incrementally until the selected load level was reached by relaxing a

hydraulic jack placed under the cantilever beam at a distance of 1m from the beam-

column joint. The load on the pipe was monitored with a load cell attached to the steel

beam cross member below the cantilever beam. The horizontal and vertical

displacements of the pipe interior wall were measured using string potentiometers

installed along the pipe length. Strains were measured using high-resolution strain

gauges placed on the inside of the pipe at the invert, crown, springline, and midway

between haunch and crown, and on the outside of the pipe at the latter two positions, as

indicated by the notation “SG” in the diagram in Figure 25 (b). The pipe was subjected

to 40% of the ultimate load for 30 days in Stage 1, after which the load was removed,

then replaced with a new load at 50% of the ultimate load for another 30 days in Stage

2, and then the load was removed and replaced at 70% of the ultimate load in Stage 3 for

another 30 days. Crack widths were measured using linear variable

transducers (LVDTs) and a feeler gauge. The pipe in the long-term test reported here is

a 1200 mm (48 in.) diameter concrete pipe of length 1.2 m (4 ft) and wall thickness 50

mm (2 in.) with PP fiber reinforcement at a dosage of 9 kg/m3 and a steel reinforcement

mesh with area 5.7 cm2/m (0.27 in2/ft) (Pipe No. 1 in Table 16).
79

(a)

(b)

Figure 25. (a) Photograph and (b) schematic diagram of the long-term test setup.
80

CHAPTER 4 : RESULTS AND DISCUSSIONS

4.1 Introduction

This chapter provides a discussion of the results obtained during the experimental

tests and numerical simulations of this research. The first section discusses the effect of

adding fiber on the mechanical properties of the concrete material. The next section

discusses the fiber reinforced concrete pipes performance under three-edge bearing load.

Then, the numerical analysis of different pipe diameters reinforced with different fiber

contents and steel areas are discussed. The validating and calibrating finite element

models used to understand synthetic fiber performance are discussed. Then, the

performance of synthetic fiber reinforced concrete pipes with reduced wall thickness are

discussed. Finally, thin-walled synthetic fiber reinforced concrete pipes under sustained

load were discussed.

4.2 Material Properties of Synthetic Fiber Reinforced Concrete under Freeze-Material

Properties

4.2.1 Effect of Adding Fiber on the Mechanical Properties of Concrete

Coefficient of Thermal Expansion

The thermal strain-temperature plots for all air-dried specimens tested using the

Ohio method for the range -60°C to 60°C (-76 to 140 °F) are shown in Figure 26. The

thermal-strain behavior of PP and PVA fiber reinforced concrete is almost linear within

the range from -40 to 60 °C (-40 to 140 °F), and nonlinear from -40 to -60 °C (-40 to -76

°F). The nonlinear behavior may be attributed to the high difference between the coarse

aggregate and hydrated cement past CTEs, which may lead to differential movement
81

between coarse aggregate and the surrounding paste due to the bond failure at large

temperature change (Neville 2011). Also, this behavior may be related to the high CTE

value of synthetic fibers compared to the CTEs of other concrete constituents (Kim et al.,

2008; Sellevold and Bjøntegaard, 2006; Mark, 2009; Tripathi, 2002). For both fiber

types, an increase in dosage led to an increase in the thermal strain, as shown in the

steeper slope in the graph; this increase was higher for PVA fiber. From Figure 26 (a), the

extreme thermal deformations for three dosages of PP fiber ranged from -625 to -656με

at -60°C (-76 °F) and between 722 to 810με at +60°C. Also, for PVA fiber reinforced

concrete specimens, thermal strains varied from -623 to -784με at -60°C (-76 °F), and

between 746 to 908με at +60°C (140 °F), as shown in Figure 26 (b). Comparing fiber

reinforced concrete with plain concrete showed that adding 9 kg/m3(15 lb/yd3) of fiber

increased thermal strain by about 12.62% for PP and 26.32% for PVA fiber reinforced

concrete specimens at 60°C (140 °F). At -60°C (-76 °F), this increase was 16.15% for PP

and 46.35% for PVA fiber reinforced concrete specimens at a temperature of -60°C (-76

°F), respectively. The increase in fiber content led to an increase in the thermal strains of

concrete-fiber matrix, as shown in Fig. 5(a and b). This behavior can be attributed to the

high deformation of synthetic fibers (high CTE, see Table 1) compared to other concrete

constituents, which increases the deformation of fiber reinforced concrete. The PVA fiber

reinforced specimens experienced higher thermal strain than the PP fiber reinforced

specimens (see Figure 26). The difference in thermal strain of the PVA and PP specimens

may be attributed to the properties of each fiber type such as fiber configuration and CTE

value, as listed in Table 1.


82

1000
Plain concrete
800 PP-6 kg/m3
PP-7 kg/m3
600 PP-9 kg/m3
400
Strain (με) 200
0
-200
-400
-600
-800
-1000
-100 -50 0 50 100
Temperature (°C)

(a)

1000
Plain concrete
800 PVA-6 kg/m3
PVA-7 kg/m3
600 PVA-9 kg/m3
400
200
Strain (με)

0
-200
-400
-600
-800
-1000
-100 -50 0 50 100
Temperature (°C)

(b)

Figure 26. Thermal strain-temperature of (a) PP and (b) PVA fiber reinforced specimens.
83

Table 17 summarizes the results of the CTE measurements obtained using the Ohio

method for each material over the temperature ranges: the full range -60°C to 60°C (-76

to140 °F), the lower half range -60°C to 0°C(-76 to 0 °F), and the upper half range 0°C to

60°C (0 to 140 °F). For all three temperature ranges, the CTE increases with fiber dosage,

except in the lower half range where the CTE decreases for PVA dosage of 7 kg/m3 (12

lb/yd3)from that of 6 kg/m3(10 lb/yd3). The increase in CTE with fiber dosage is larger for

PVA fiber than for PP fiber, particularly at the highest dosage, where the increase is 16.73%

for concrete with PP fiber over the full range -60°C to 60°C (-76 to140 °F), and 34.52%

for concrete with PVA fiber. Therefore, the effect of fiber on the concrete structures may

be considered in the design stage of fiber reinforced concrete structures.


84

Table 17. CTE Measured Using Ohio Method Before Applying Freeze-Thaw Cycles.

CTE (10-6 /°C)* Multiply 1/°C by 5/9 to obtain 1/°F

Temp. from COV % Temp. COV % Temp. from COV %


Fiber dosage
Fiber type -60 to 60°C from 0 to 60°C
(kg/m3)
-60 to

0°C

No Fiber 0 10.40 4.17 8.36 8.10 12.00 5.43

PVA 6 11.32 5.39 10.06 4.26 12.59 9.90

PVA 7 11.68 1.37 9.29 9.34 13.66 5.04

PVA 9 13.99 1.37 12.64 4.98 15.07 3.22

PP 6 11.27 8.80 9.98 8.6 12.35 8.13

PP 7 11.56 0.20 10.34 5.50 12.41 3.40

PP 9 12.14 0.76 10.43 0.07 13.48 0.66

* Multiply 1/°C by 5/9 to obtain 1/°F.

The CTEs were also measured for all concrete fiber dosages using the AASHTO

TP60-00 (2007) standard. For convenience in comparing the AASHTO T60-00 (2007)

results with the Ohio method results, the average slope of the strain-temperature curve for

the temperature range from 10°C to 50°C (50 to122 °F) in accordance with AASHTO

T60-00 (2007) was calculated. The CTE results of the two methods are compared in

Table 18. The Ohio method consistently gave greater values than the AASHTO method

for all concrete fiber dosages because the AASHTO method used samples saturated with

water compared to the oven-dried condition of specimens in the Ohio method. According

to ACI 209R-92 (2008) and Neville (1995), increased moisture content reduces the CTE
85

value of normal strength concrete. ACI 209R-92 (2008) specifies corrections to the CTEs

determined from saturated concrete specimens. Therefore, the AASHTO TP60-00 (2007)

CTE values should be corrected for the expected degree of saturation of the concrete

member. In Table 7, the AASHTO method CTE is 13.94% lower than the Ohio method

for concrete without fiber. Also, the CTEs of the specimens reinforced with 9 kg/m3(15

lb/yd3) of PVA and PP are 23. 53% and 17.96% is lower than the Ohio CTE method,

respectively.

The Ohio CTE values represent the severest air-dry condition that the concrete

superstructure may experience. Due to the lack of environmental data regarding fiber

reinforced concrete structures under severe weather conditions, the values of the Ohio CTE

test for the thermal coefficient of expansion may be used for design purposes. The findings

of the Ohio CTE test method also highlighted the significance of using a severe temperature

range to evaluate the thermal effects on the fiber reinforced concrete.


86

Table 18. CTE Results of Ohio CTE Device and AASHTO T60-00.

Fiber type Fiber dosage (kg/m3) CTE (10-6 /°C)* Temp. from 10 to 50°C

Ohio CTE device AASHTO T60

No Fiber 0 11.77 10.33

PVA 6 12.22 10.70

PVA 7 13.56 11.27

PVA 9 14.17 11.47

PP 6 12.52 10.29

PP 7 12.51 11.3

PP 9 13.66 11.58

* Multiply 1/°C by 5/9 to obtain 1/°F.

Dynamic Modulus of Elasticity

The dynamic modulus of elasticity was measured for each material according to

the ASTM C215-14 (2014) method described earlier before applying freeze-thaw cycles,

with the results listed in Table 19. Since the fibers have a low modulus of elasticity

compared to the conventional concrete, the fiber reinforced concrete was expected to

have a lower modulus of elasticity than the plain concrete. The PVA fiber reinforced

specimens had elastic modulus values that decreased by 3.8%, 8.6%, and 13% for fiber

dosages 6, 7, and 9 kg/m3 (10, 12, and 15 lb/yd3), respectively compared to plain

concrete; for PP fiber reinforced concrete the corresponding decreases were 1.7%, 4.5%,

and 16%. Therefore, adding fiber increased the flexibility of the concrete material. As a

result, specimens reinforced with synthetic fiber endure high deformations before

showing structural damage compared to the plain concrete. The results of the present
87

study concur with the findings from past research by Çavdar (2014) showing that adding

synthetic fibers decrease the dynamic modulus of elasticity.

Table 19. Dynamic Modulus of Elasticity of Concrete Prisms with Different Fiber Dosages.

Fiber dosage, Fiber type Average Dynamic modulus of elasticity ,


COV (%)
kg/m3 (lb/yd3) MPa (ksi)

0 No Fiber 44602 (6469) 2.55

6 (10) PVA 42899 (6222) 3.01

7 (12) 40755 (5911) 2.58

9 (15) 38707 (5614) 0.82

6 (10) PP 43851 (6360) 1.15

7 (12) 42617 (6181) 1.98

9 (15) 37432 (5429) 3.32

Flexural Performance

The flexural strength results for plain and fiber reinforced concrete specimens are

listed in Table 20. Adding fiber increased the flexural strength of the concrete

significantly in an amount proportional to the fiber type and dosage. This improvement in

flexural strength was expected since concrete is a brittle material. Adding fiber enhanced

the flexural strength of the concrete because of the fibers ability to bridge micro and

macro cracks (Banthia and Sappakittipakorn, 2007). The flexural increased by 15%, 24%,

and 43% for fiber dosages 6, 7, and 9 kg/m3 (10, 12, and 15 lb/yd3),, respectively

compared to plain concrete; for PP fiber reinforced concrete the corresponding increases

were 17.3%, 17.4%, and 51%. PP fiber reinforcement provided more improvement than
88

PVA fiber reinforcement due to the differences in the geometrical properties of the fibers,

such as length and diameter. PP fiber has surface area more than PVA fiber based on

physical properties listed in Table 1 . Therefore, it was expected that PP fiber specimens

to have more bending tensile strength compared to the PVA fiber reinforced specimens as

well as plain concrete. Figure 27 illustrates the load-deflection curves of PP (Figure 27a)

and PVA (Figure 27b) fiber reinforced concrete prisms with different fiber dosages. The

PP reinforced concrete showed better load capacity and post-cracking load performance

than the PVA fiber reinforced specimens. This can be explained by the greater ductility

of PP fiber compared to PVA fiber. At 3 mm deflection, which is the serviceability

deflection specified by ASTM C1609/1609M-15 (2015b), most PVA fibers had pulled

out from the concrete while PP fibers were still engaged. This behavior is mainly

attributed to the configuration of each fiber type. PP fiber have more surface area than

PVA fiber, resulting in high bond strength between PP fiber and surrounding concrete

matrix. This led to larger cracks in the PVA fiber reinforced specimens than the PP fiber

reinforced specimens, which can be seen in Figure 28, which shows the crack width at 3

mm deflection for PP (see Figure 28a) and PVA (see Figure 28b) fiber reinforced

concrete prisms at a dosage of 7 kg/m3 (12lb/yd3).


89

Table 20 Flexural Strength of Fiber Reinforced Concrete.

Fiber dosage, Average Flexural strength, COV (%)


Fiber type
kg/m3 (lb/yd3) MPa (psi)

0 No Fiber 5.43 (787.55) 2.25

6 (10) PVA 6.25 (906.48) 2.98

7 (12) 6.76 (980.45) 4.36

9 (15) 7.76 (1125.49) 3.15

6 (10) PP 6.37 (923.89) 5.12

7 (12) 6.92 (924.76) 3.55

9 (15) 8.20 (1189.31) 5.56


90
20
Plain concrete
PP-6 kg/m3
PP-7 kg/m3
16 PP-9 kg/m3

12

Load (kN)
8

0
0 1 2 3 4 5
Deflection (mm)

(a)
20
Plain concrete
PVA-6 Kg/m3
PVA-7 Kg/m3
16 PVA-9 Kg/m3

12
Load (kN)

0
0 1 2 3 4 5
Deflection (mm)

(b)

Figure 27. Load versus deflection plots of (a) PP and (b) PVA fiber reinforced specimens.

(Note: 1 kN = 0.2248 kip; 1 mm = 0.0394 in).


91

(a) (b)

Figure 28. The crack width of (a) PP and (b) PVA fiber reinforced specimens. (Note: 1

mm = 0.0394 in).

4.3 Effect of Freeze-Thaw Cycles on the Mechanical Properties of Fiber Reinforced

Concrete

Repeated freezing and thawing of concrete initiates micro cracking in concrete

that breaks contact between concrete constituents and causes loss of structural integrity.

Micro cracking is expected to impact various material properties of the concrete, while

the reinforcing fibers are expected to counteract these effects by holding concrete pieces

together.

Coefficient of Thermal Expansion

The effect of freeze-thaw cycles on the CTE of fiber reinforced and plain concrete

is shown in Figure 29. The CTE specimens were tested every 100 cycles. The plain

concrete showed a decrease in the CTE value by 15% after 200 freeze-thaw cycles; after
92

300 cycles, the specimens had deteriorated too much to measure the CTE. Micro cracks

due to freezing-thawing cycles break the bond between the concrete constituents, and

then causing the concrete specimens to lose their integrity to expand/contract as unit

structure. Thus, the plain concrete showed a decrease in the CTE.

The CTE of PVA fiber reinforced concrete specimens increased after 300 cycles

by 12% for fiber dosage of 7 kg/m3 and 14.5% for fiber dosage of 9 kg/m3. The CTE of

the 6 kg/m3 PVA fiber dosed cement declined by 1.8% compared to the CTE at zero

cycle, which suggests that this fiber dosage was not sufficient to bridge all microcracks

initiated due to freeze-thaw cycles. The CTE of PP fiber reinforced concrete specimens

increased by 12%, 4.4%, and 14.7% with fiber dosages of 6, 7, and 9 kg/m 3 (10, 12, and

15 lb/yd3),, respectively. Therefore, the higher CTE in fiber reinforced concrete after 300

freeze-thaw cycles can be attributed to the higher CTE of the fiber materials bridging the

microcracks that cause the specimens to expand/contract with high thermal strain, as

shown in Figure 30. Based on these results, the optimum fiber dosage was found to 7

kg/m3 (12 lb/yd3) for both fiber types, which shows the least increase in the CTE.
93

16

14

12

CTE (10-6/°C)
10

4
Plain concrete
2 PVA-6 kg/m3
PVA-7 kg/m3
0
0 50 100 150 200 250 300
Number of freeze-thaw cycles

(a)

16

14

12
CTE (10-6/°C)

10

4
Plain concrete
2 PP-6 kg/m3
PP-7 kg/m3
0
0 50 100 150 200 250 300
Number of freeze-thaw cycles

(b)

Figure 29. CTE versus freeze-thaw cycles: (a) PVA and; (b) PP fiber reinforced concrete

specimens. (Note: Multiply 1/°C by 5/9 to obtain 1/°F.)


94

Cracks in the interface between


coarse aggregate and cement
mortar

Cracks in adhering mortar

Figure 30. The damage in the meso-structure of PVA fiber reinforced specimen with fiber

dosage of 7 kg/m3 (12 lb/yd3).

Mass Lost

Figure 31 illustrates the change in the concrete prism masses over 300 freeze-

thaw cycles as a fraction of the original specimen mass. Both PP and PVA fiber

reinforced concrete prisms showed different percentages of mass loss after applying 300

cycles. PP fiber reinforced specimens showed loss in the mass by 0.41%, 0.88%, and

4.90% at fiber dosage of 6 kg/m3, 7 kg/m3, and 9 kg/m3(10, 12, and 15 lb/yd3),,

respectively, while PVA fiber reinforced specimens lost 1.50%, 0.06%, and 0.16%,

respectively. Plain concrete showed a noticeable decrease in mass, especially after the

120 freeze-thaw cycles, eventually losing 74% of its mass after 300 cycles. This massive

reduction in the plain concrete mass can be attributed to water expansion when it freezes

in the concrete matrix pores (Gruebl 1980; Korhonen 2002; Setzer and Liebrecht 2002).

This mechanism exerts pressure on the pore walls. When the applied pressure exceeds the

tensile strength of the concrete material, micro cracks start to propagate causing an
95

increase in the pore volume of the concrete matrix. As a result, the deterioration increases

after each cycle causing the specimens to lose their integrity and losing more mass. On

the other hand, fiber reinforced specimens showed better performance due to fiber

presence connecting the concrete constituents compared with the plain concrete

specimens. Therefore, Synthetic fiber increased the durability of the concrete material

under these freeze-thaw conditions.


96

1.1
1
0.9
0.8
Normalized mass 0.7
0.6
0.5
0.4 Plain concrete
PP-6 Kg/m3
0.3 PP-7 Kg/m3
PP-9 Kg/m3
0.2
0 50 100 150 200 250 300
Number of freeze-thaw cycles

(a)

1.1
1
0.9
0.8
Normalized mass

0.7
0.6
0.5
0.4 Plain concrete
PVA-6 Kg/m3
0.3 PVA-7 Kg/m3
PVA-9 Kg/m3
0.2
0 50 100 150 200 250 300
Number of freeze-thaw cycles

(b)

Figure 31. Mass of fiber reinforced concrete prisms as a function of freeze-thaw cycles:

(a) PP and (b) PVA fiber reinforcement.


97

Dynamic Modulus of Elasticity

Figure 32 shows the effect of freeze-thaw cycles on the dynamic modulus of

elasticity of concrete specimens. For the plain concrete specimens, it was difficult to

measure the dynamic modulus of elasticity after 120 freeze-thaw cycles due to specimen

disintegration. The dynamic modulus of elasticity of PP fiber reinforced concrete prisms

decreased by about 12%, 9%, and 51% at fiber dosages of 6, 7, and 9 kg/m3(10, 12, and

15 lb/yd3), respectively, after 300 freeze-thaw cycles. Since the relationship between the

dynamic modulus of elasticity and the unit mass is proportional, there was also a decrease

in the dynamic modulus of elasticity. Another reason for the dynamic modulus reduction

is that when the specimen undergoes more freeze-thaw cycles, the fundamental transverse

frequency decreases as a result of material deterioration

The steep decrease in the dynamic modulus of elasticity for concrete with PP fiber

at the dosage of 9 kg/m3(15 lb/yd3) is due to the increase in the surface area of the mortar

matrix, which could lead to reduced workability of the concrete mixture (Karahan and

Atiş, 2011; Salih and Al-Azaawee, 2008; Raghavan et al. 1998; Zhong, 2007). The

reduction in the concrete mixture workability will increase the number of permeable

voids, which leads to tensile stresses in the surrounding matrix when water freezes in the

voids. If these stresses exceed the maximum concrete tensile strength, microcracking will

occur (Neville and Brooks, 1993). Adding fiber to the concrete mixture will reduce

bleeding and segregation during concrete mixing and create a grid structure supporting

concrete material structure along with the aggregate (Bagherzadeh et al. 2011), which

creates more air voids and can improve the durability of the concrete. Also, Huang (1997
98

and 2001) and Karahan and Atiş (2011) stated that fiber inclusion in the concrete material

increased the number of large pores and water absorption. This implies that the increase

in water absorption is due to voids caused by fiber addition. However, at a PP fiber

dosage of 9 kg/m3, the limited space between fibers made interface areas overlap each

other, which has a detrimental effect on the freeze-thaw durability and leads to a

reduction in the dynamic modulus of elasticity.

PVA fiber reinforced specimens showed a reduction in dynamic modulus of

elasticity by 5.4%, 1.6%, and 1 % for fiber dosages of 6, 7, and 9 kg/m3, respectively,

after 300 freeze-thaw cycles. Comparing the dynamic modulus of elasticity for both fiber

types showed that the PVA fiber reinforced specimens experienced less reduction than PP

fiber reinforced specimens, especially, with fiber dosage of 9kg/m3 (15 lb/yd3), after 300

freeze-thaw cycles. This mainly could be related to fiber configuration. The PP fiber has

a length and diameter greater than the PVA fiber, which could lead to more water

absorption (Bagherzadeh et al. 2012). As a result, the PP fiber reinforced specimens

could experience more micro-cracks due to the freeze-thaw cycles than the PVA fiber

reinforced specimens, which may reduce the dynamic modulus of elasticity. The best

performance was found to be with using fiber dosage of 9 and 7 kg/m3 (12 and 15 lb/yd3),

for PVA and PP fiber reinforced specimens, respectively, which maintained the highest

dynamic modulus of elasticity after 300 freeze-thaw cycles.


99

50000
45000

Dynamic Modulus of elasticity (MPa)


40000
35000
30000
25000
20000
15000
10000 Plain concrete
PP-6 Kg/m3
5000 PP-7 Kg/m3
PP-9 Kg/m3
0
0 50 100 150 200 250 300
Number of freeze-thaw cycles

(a)

50000
45000
Dynamic Modulus of elasticity (MPa)

40000
35000
30000
25000
20000
15000
Plain concrete
10000
PVA-6 Kg/m3
5000 PVA-9 Kg/m3
PVA-7 Kg/m3
0
0 50 100 150 200 250 300
Number of freeze-thaw cycles

(b)

Figure 32. Dynamic modulus of elasticity versus freeze-thaw cycles: (a) PP and (b) PVA

fiber reinforced specimens. (Note: 1 MPa = 0.145 ksi)


100

Flexural Performance

Figure 33 shows that synthetic fiber minimized the reduction in the flexural

strength by freezing and thawing compared with plain concrete. After 300 freeze-thaw

cycles, the flexural strength of PP fiber reinforced specimens decreased by 36 %, 19%,

and 78% for fiber dosages of 6, 7, and 9 kg/m3(10, 12, and 15 lb/yd3),, respectively. As

the fiber dosage increased from 6 to 7kg/m3, the flexural strength resistance to the freeze-

thaw cycle increased by about 50%. Using low fiber, lower than 7kg/m3, may not be

enough to link all micro cracks propagated due to freeze-thaw cycles. Specimens with 9

kg/m3 showed lower resistance to the freeze-thaw effects compared to lower dosages.

Adding fiber to the concrete material increases the entrained air voids, while the fibers

tend to block capillary pores in concrete specimens, which reduces water ingress

(Richardson, 2003). As a result of these processes, less water absorption and better

freeze-thaw resistance are expected. However, according to the United States Department

of Agriculture (USDA), natural resources conservation service (USDA, 1976), once air

content increases past 12% there is a significant decrease in durability. Richardson et al.

(2012) noted that adding 0.9 kg/m3 of PP fiber increased air concrete by 1.5%. The

maximum PP fiber dosage of 9 kg/m3 (15 lb/yd3), employed in this study is ten times that

was used by Richardson et al. (2012), which could produce air content over 12%.

Increasing the fiber dosage while maintaining the same water-to-cement ratio will

decrease the workability of the concrete mixture due to the increase in the surface area

and increase the possibility of creating new air voids. When water goes inside theses

voids and solidifies at the freeze phase, high stresses will be generated within the
101

concrete, leading to damage. For PVA fiber reinforced concrete, increased fiber dosage

improved the resistance to the freeze-thaw cycles until the optimum fiber was reached.

After this point, a reduction in the flexural strength was observed. The reduction in the

flexural strength of PVA fiber reinforced concrete was 33%, 22%, and 27% at fiber

dosages of 6, 7, and 9 kg/m3(10, 12, and 15 lb/yd3), respectively. The best MOR

performance under freeze-thaw conditions appeared at fiber dosage of 7 kg/m3 for PP and

PVA fibers, which suggests there is an optimum fiber dosage. It is recommended,

however, that the optimum fiber dosage be investigated further for each fiber type.
102

9
8
7
6

Flexural strength (MPa) 5


4
3
2 Plain concrete
PVA-6 kg/m3
1 PVA-7 kg/m3
PVA-9 kg/m3
0
0 50 100 150 200 250 300
Number of cycles

(a)

9
8
7
Flexural strength(Mpa)

6
5
4
3
Plain concrete
2
PP-6 kg/m3
1 PP-7 kg/m3
PP-9 kg/m3
0
0 50 100 150 200 250 300
Number of cycles

(b)

Figure 33. Flexural strength versus freeze-thaw cycles for concrete reinforced with (a)

PVA and (b) PP fiber. (Note: 1 MPa = 0.145 ksi)


103

Figure 34 and Figure 35 depict the load deflection response of prisms reinforced

with PVA and PP fibers before and after applying 300 freeze-thaw cycles, respectively.

The deflection increased linearly with applied load until the first crack. After the first

crack, an abrupt decrease in the applied load was observed. The magnitude of the load

reduction after the first crack for 300 freeze-thaw cycles load-deflection response was

small compared to the zero-cycle load reduction. This mainly can be attributed to the

existence of micro cracks propagated after 300 cycles. Beyond the first crack, the PVA

fiber reinforced specimens showed a deflection hardening response followed by a steep

decrease in the applied due to fiber pull-out or rapture. This steep decrease in the load

deflection response was greater after 300 cycles than zero cycle. This may be related to

the weakness in bond strength between fiber and surrounding matrix due to freeze-thaw

cycles. On the other hand, PP fiber reinforced specimens showed a deflection hardening

behavior after the first crack even after 300 cycles. This type of fiber has a surface area

greater than PVA fiber, which increased the bond strength between fiber and concrete

components. Therefore, the possibility of failure under certain loads for PVA fiber

reinforced specimens is greater than PP fiber reinforced specimens. In general, fiber

reinforced concrete showed consistently better performance compared to plain concrete

with improved resistance to freeze-thaw cycles. Past research by Allan and Kukacka

(1995) indicated adding fiber did not significantly improve freeze-thaw durability, while

this study shows using synthetic fiber enhanced concrete durability against the freeze-

thaw cycles. Also, this study contributes to the literature by providing more data on

freeze-thaw resistance of concrete containing PP and PVA fibers.


104

16
14 PVA-6 Kg/m3-0C
12 PVA-6 Kg/m3-300C

Load (kN)
10
8
6
4
2
0
0 1 2 3 4 5
Deflection (mm)

(a)

16
PVA-7 Kg/m3-0C
14 PVA-7 Kg/m3-300C
12
Load (kN)

10
8
6
4
2
0
0 2 4
Deflection (mm)

(b)

16
PVA-9 Kg/m3-0C
14 PVA-9 Kg/m3-300C
12
Load (kN)

10
8
6
4
2
0
0 1 2 3 4 5
Deflection (mm)

(c)

Figure 34. Load versus deflection of prisms reinforced with PVA fiber with different

dosages before and after applying 300 freeze-thaw cycles: 6 kg/m3 (10 lb/yd3); (b) 7

kg/m3(12 lb/yd3); (c) 9 kg/m3 (15 lb/yd3). (Note: 1 kN = 0.2248 kip; 1 mm = 0.0394 in)
105

15

Load (kN)
10

5
PP-6 Kg/m3-0C
PP-6 Kg/m3-300C
0
0 1 2 3 4 5
Deflection (mm)

(a)

15

10
Load (kN)

5
PP-7 Kg/m3-0C
PP-7 Kg/m3-300C
0
0 1 2 3 4 5
Deflection (mm)

(b)

15

10
Load (N)

PP-9 Kg/m3-0C
5 PP-9 Kg/m3-300C

0
0 1 2 3 4 5
Deflection (mm)

(c)

Figure 35. Load versus deflection of prisms reinforced with PP fiber with different

dosages before and after applying 300 freeze-thaw cycles: 6 kg/m3 (10 lb/yd3); (b) 7

kg/m3(12 lb/yd3); (c) 9 kg/m3 (15 lb/yd3). (Note: 1 kN = 0.2248 kip; 1 mm = 0.0394 in).
106

4.4 Evaluation of Synthetic Fiber Reinforced Concrete Pipe Performance Using Three-

Edge Bearing Test

4.4.1 Load-deflection Response

Load-deflection curves for concrete pipes of diameter 600, 1200, and 1500 mm

(24, 48, and 60 in.) are shown in Figure 36, Figure 37, and Figure 38, respectively. A

label with four numerals separated by hyphens was created to distinguish each tested

pipe. For example, in the label 600-9-1.5-1, the first numeral represents the pipe diameter

in mm (600 mm (24 in.) ), the second numeral represents the fiber dosage in kg/m3 (9

kg/m3), the third numeral represents the amount of steel reinforcement in cm2/m (1.5

cm2/m [0.07 in2/ft]; 0 indicates no steel), and the fourth numeral represents the specimen

number of that type (1 denotes the first of two pipes with these properties.). Due to

differences in pipe lengths used in this study, the applied load was normalized by

dividing the load by the ratio of the pipe length to the shortest pipe length, 1.2 m. The

vertical displacement represents the average readings of three displacement transducers

installed along the tested pipe.

For the 600 mm (24 in.), concrete pipe diameter without steel reinforcement,

Figure 36(a) shows the load-deflection behavior of pipes reinforced with fiber dosage 9

and 18 kg/m3. Pipes reinforced with fiber dosage of 9 kg/m3 showed a linear behavior

until the first crack and then deflection softening behavior. With a fiber dosage of 18

kg/m3, the 600 mm (24 in.) pipe diameter had a significant enhancement in load

deflection response; as there was deflection hardening after the first crack. Also, the

average first crack and the ultimate loads increased by 8.22% and 26.80%, respectively,
107

for the higher fiber dose. Figure 36(b) shows that including 18 kg/m3 fiber dosage with

steel reinforcement improved the load-defection behavior as average cracking, and

ultimate loads increased by 11.0% and 4.95% compared to the pipe reinforced with a

one-layer steel cage and fiber dosage of 9 kg/m3. In addition, increasing fiber dosage

limited the instability regions after the first crack due to the increase in the concrete

matrix strength due to the increased number of fibers resisting crack propagation. For the

1200 mm (48 in.) pipe diameter, Figure 37(b) shows the load deflection behavior of

using synthetic fiber as sole reinforcement at the dosages of 4.75 and 9 kg/m3 (8 to 15

lb/yd3) as well as a plain concrete pipe shown in Figure 37(a). It can be seen that the

cracking load (produces 0.3 mm crack width) of plain concrete pipe coincides with its

ultimate load. This due to the brittle behavior of plain concrete. The cracking load also

matches the ultimate load of pipes tested with fiber dosage of 4.75 and 9 kg/m3 as sole

reinforcement; however, these load values for the pipes with fiber were considerably

larger than for the plain concrete pipes: 21.21% (4.98%) and 31.0% (10.0%) larger

cracking (ultimate) loads. The tested pipes exhibited a softening deflection behavior after

reaching its cracking load. Using synthetic fiber as sole reinforcement enhanced the load

capacity, ductility, and post-cracking behavior of tested pipes due to the fibers’ tendency

to resist the propagated cracks. The load deflection curves depicted in Figure 37(c - e)

represent the 1200 mm (48 in.) concrete pipe diameter behavior reinforced using one

steel cage reinforcement layer with fiber dosages of 4.75, 9, 13.5, and 18 kg/m3.

Increasing fiber dosage enhanced the cracking, ultimate load, post-cracking behavior, and

the ductility of the pipes with the steel reinforcement. The cracking load of concrete pipe
108

reinforced with one steel cage layer and fiber dosage of 9, 13.5, and 18 kg/m3 increased

by 3.70, 8.60, and 29.15%, respectively, compared to 1200 mm (48 in.) pipes diameter

reinforced with fiber dosage of 4.75 kg/m3 while the ultimate load increased by 14.77,

16.83, and 43.27%, respectively. The effect of increasing fiber dosage on the response of

the 1200 mm (48 in.) pipes under the three-edge bearing test was more effective than the

effect of increasing fiber dosage on the 600 mm (24 in.) diameter pipe.

For 1500 mm pipe diameter, pipes were reinforced with synthetic fiber dosages of

9, 13.5, and 18 kg/m3 along with one-layer steel cage, as shown in Figure 38(a-c).

Increasing fiber dosage enhanced cracking and ultimate load. Cracking load increased by

15.60 and 19.55% with fiber dosage of 13.5 and 18 kg/m3, respectively. Ultimate load

increased by 26.48 and 33.07% with fiber dosage of 13.5, and 18 kg/m3, respectively.

The effect of incorporating synthetic fiber on both cracking and ultimate load was more

noticeable than concrete pipe diameter of 600 m and 1200 mm as pipe thickness

increased to 1500 mm, which confirms the observed trend of the increased fiber dosage

to yield greater improvement resistance to applied loads for larger diameter pipes.

All specimens reinforced with synthetic fiber surpassed 2% deflection of their

inside diameter with high load capacity. This indicates that buried synthetic fiber

reinforced concrete pipe could interact with the surrounding soil and carry more load than

expected. The inclusion of synthetic fiber enhanced the cracking load, ultimate load,

flexibility, stability, and ductility for all tested pipes. The size of this enhancement is

related to two factors. The first factor is the pipe diameter. For small diameter pipes, the

primary failure mode is flexural control, which means that fibers resist flexural cracks.
109

As concrete pipe diameter increases, different failure mechanisms become critical such as

radial and diagonal failures, and due to fiber dispersion in the concrete pipe wall, more

fibers are engaged to resist these failure modes. The second factor that affects the amount

of enhancement is the pipe wall thickness. As pipe wall thickness increases, better fiber

dispersion is expected, and then fiber more efficiently resists the applied load. Thus,

concrete pipes with a diameter of 1200 mm and 1500 mm (48 and 60 in.) exhibited better

performance in terms of load carrying capacity than the 600 mm (24 in.) pipes. More

studies are required to show the effect of using fiber dosages greater than 18 kg/m3 (30

lb/yd3) on load carrying capacity.


110

350
600-9-0-1
300 600-9-0-2
600-18-0-1
600-18-0-2
250

Load (kN)
200

150

100

50 2% deflection

0
0 5 10 15 20 25 30 35 40 45
Deflection (mm)

(a)

350
600-9-1.5-1
600-9-1.5-2
300 600-18-1.5-1

250
Load (kN)

200

150

100

50
2% deflection
0
0 5 10 15 20 25 30 35 40 45
Deflection (mm)

(b)

Figure 36. Load deflection curves of 600 mm (24 in.) pipe diameter reinforced with: (a)

synthetic fiber only; (b) synthetic fiber with one steel cage layer. (Note: 1 kN = 0.2248

kip; 1 mm = 0.0394 in)


111

350
1200-0-0-1
300 1200-0-0-2
1200-0-0-3
250

Load (kN.)
200

150

100

50

0
0 1 2 3 4 5
Deflection (mm)

(a)

350
1200-4.75-0-1
1200-4.75-0-2
300 1200-4.75-0-3
1200-9-0-1
250 1200-9-0-2
1200-9-0-3
Load (kN.)

200

150

100 2% deflection ratio

50

0
0 5 10 15 20 25 30 35 40 45
Deflection (mm)

(b)
112

350
1200-4.75-5.1-1
300 1200-4.75-5.1-2

250

Load (kN.)
200

150

100

50 2% deflection ratio

0
0 5 10 15 20 25 30 35 40 45
Deflection (mm)

(c)

350
1200-9-5.1-1
1200-9-5.1-2
300
1200-9-5.1-3

250
Load (kN.)

200

150

100
2% deflection ratio
50

0
0 5 10 15 20 25 30 35 40 45
Deflection (mm)

(d)
113

350
1200-13.5-5.1-1
1200-13.5-5.1-2
300
1200-13.5-5.1-3
250

Load (kN)
200

150

100

50 2% deflection ratio

0
0 5 10 15 20 25 30 35 40 45
Deflection (mm)

(e)

350
1200-18-5.1-1
1200-18-5.1-2
300 1200-18-5.1-3

250
Load (kN)

200

150

100
2% deflection ratio
50

0
0 5 10 15 20 25 30 35 40 45
Deflection (mm)

(f)

Figure 37. Load deflection curves of 1200 mm (48 in.) pipes diameter: (a) plain concrete

pipes; (b), pipes reinforced with fiber dosage of 4.75 and 9 kg/m3; (b, c, d, e, and f) pipes

reinforced with fiber dosage of 4.75, 9, 13.5, and 18 kg/m3, respectively, along with one

steel cage layer. (Note: 1 kN = 0.2248 kip; 1 mm = 0.0394 in).


114

350
1500-9-5.1-1
300 1500-9-5.1-2
1500-9-5.1-3
250

Load (kN.) 200

150 2%

deflection ratio

100

50 2% deflection ratio

0
0 5 10 15 20 25 30 35 40 45
Deflection (mm)

(a)

350
1500-13.5-5.1-1
1500-13.5-5.1-2
300
1500-13.5-5.1-3
250
Load (kN)

200

150

100

50 2% deflection ratio

0
0 5 10 15 20 25 30 35 40 45
Deflection (mm)

(b)
115

350

300

250

Load (kN)
200

150

100
1500-18-5.1-1
50 1500-18-5.1-2 2% deflection ratio
1500-18-5.1-3
0
0 5 10 15 20 25 30 35 40 45
Deflection (mm)

(c)

Figure 38. Load deflection curves of 1500 mm pipe diameter reinforced with one steel

cage layer along with fiber dosage of: (a) 9 kg/m3 ;(b) 13.5 kg/m3 ;(c) 18 kg/m3. (Note: 1

kN = 0.2248 kip; 1 mm = 0.0394 in)

4.4.2 Failure Mechanism

The crack opening was monitored while applying load to measure the crack width

and observe fiber failure. The main failure mechanism observed for the test pipes was a

flexural failure. This mode was characterized by longitudinal cracks at the inside face of

the crown and invert as well as along outside face of the spring lines, as shown in Figure

39. As it can be seen, the largest crack was at the highest flexural tension stresses zone.

Also, longitudinal cracks with a small width were observed. These multiple cracking

patterns resulted in the pipe actually deflecting more than a non-fiber reinforced pipe.

The negative of this benefit is the increase in strain on the fibers, which are time-
116

dependent materials. A greater discussion of this critical design issue is a principle

interest in developing a structural design criterion for this product. Other failure

mechanisms such as radial and diagonal cracks associated with large concrete pipes

diameter were observed at high load rates. This can be attributed to fiber ability to

distribute the stresses to a greater degree.

Figure 39. Crack development of concrete pipes: 1200 mm (48 in.).

4.4.3 ASTM Strength Requirements

According to ASTM C76 (ASTM 2015), concrete pipes are classified into five

Classes, numbered I through V, based on the strength required for each class, as listed in

Table 5. It should be mentioned that pipe diameters of 600 mm (24 in.) and 1200 mm

appear in all classes except Class I. In order to validate the strength of fiber reinforced

concrete pipes, the strength results (loads that produce 0.3 mm crack width and ultimate
117

load) were compared to those specified by ASTM C76 (2015). The results of cracking

load and ultimate load for all tested concrete pips are summarized in

Table 21. The average coefficient of variation (COV) ranged from 0.73% to

28.81% for cracking load, and from 1.73% to 17.91% for ultimate load. Based on these

results, concrete pipes with a diameter of 600 mm (24 in.) reinforced with fiber dosage

of 9 and 18 kg/m3 (15 and 30 lb/yd3) as only reinforcement met the cracking load and

ultimate load requirements specified by ASTM C76 (ASTM 2015) as Class II and III.

Adding these fiber dosages eliminated the use of steel reinforcement without

compromising the strength requirements specified by ASTM standard. Another

advantage of using fiber reinforcement is the reduction of the total cost as the fiber is

considerably cheaper than steel reinforcement. In order to meet strength requirements of

Class IV and Class V that demand high strength requirements, minimum steel

reinforcement of one layer of cage with 1.5 cm2/m (0.07 in2/ft) area placed at 25.4 mm

from the inside diameter was used. Concrete pipes with a diameter of 600 mm (24 in.)

reinforced with steel cage of one layer and fiber dosage of 9 and 18 kg/m3 met the

strength requirements of Class IV. These fiber dosages with minimum steel

reinforcement helped reduce 85.84% of the steel reinforcement area compared with the

conventional steel reinforcement of concrete pipe for the same class. However, the

cracking load results did not meet the cracking load specified for Class V even though the

ultimate load of fiber reinforced concrete pipes exceeded Class V’s ultimate load.

Concrete pipes with a diameter of 1200 mm (48 in.) reinforced purely with synthetic
118

fiber dosage of 4.79 and 9 kg/m3 met the strength requirements for any class. Therefore,

using one steel cage layer along with synthetic fibers was necessary. The results of crack

and ultimate load of 1200 mm (48 in.) diameter pipes reinforced with fiber dosage of

4.75, 9, and 13.5 kg/m3 along with one steel cage layer exceeded the strength

requirements of pipes Class II and Class III. Moreover, strength results of 1200 mm (48

in.) diameter concrete pipes reinforced with 18 kg/m3 fulfilled the strength requirements

of ASTM C76 (ASTM 2015) Class II, III, and IV. The reduction in steel reinforcement

area associated with incorporating synthetic fiber and one steel cage layer was 51.02%,

63.07%, and 78.76% for Class II, III, and IV, respectively. Since only fiber reinforcement

was not enough for 1200 m pipe diameter to meet the stranded specifications under all

classes, concrete pipes with a diameter of 1500 mm were reinforced with synthetic fiber

as well as one steel cage layer with area of 5.1 cm2/m. The strength results of 1500 mm

pipes diameter reinforced with 13.5, and 18 kg/m3 along with one steel cage layer

satisfied the strength requirements of pipes Class I, II, III, and IV. Moreover, pipes

reinforced with 9 kg/m3 along with one steel cage layer satisfied the strength

requirements of pipes Class I, II, III. The reduction in steel reinforcement cage area

associated with incorporating synthetic fiber with one steel cage layer was 57.14%,

63.63%, 73.91%, and 85% for Class I, II, III, and IV, respectively.

The outcomes reported previously concur with past research (Peyvandi et al.,

2013 and 2014; de la Fuente et al., 2014; Park et al., 2014; Wilson and Abolmaali, 2014)

that using synthetic fiber can be a suitable replacement for the reinforcement steel cage in

precast concrete pipes. However, it should be mentioned that the results reported above
119

could be different from manufacturer to another. For example, comparing the results of

600 mm (24 in.) reported by (Wilson et al., 2014) shows a difference in the fiber dosage

that achieves strength requirements from manufacturing plant to another. This can be

attributed to differences in the material used in the mix design such coarse aggregate, w/c

ratio, etc. The achieved strength is a function of the synthetic fiber content, pipe wall

thickness, and pipe diameter.

Table 21. Summary of the Crack and Ultimate Loads of Tested Pipes

Cracking
COV Ultimate load COV
No Pipe load Average Average
(%) (kN/m/m) (%)
(kN/m/m)
1 600-9-0-1 115.29 116.35
118.33 3.64 118.86 3.0
2 600-9-0-2 121.38 121.38
3 600-9-1.5-1 123.39 229.20
120.32 3.6 232.55 2.03
4 600-9-1.5-2 117.26 235.90
5 600-18-0-1 121.19 154.43
128.06 7.58 150.72 3.48
6 600-18-0-2 134.93 147.00
7 600-18-1.5-1 133.54 133.54 - 244.10 244.10 -
8 1200-0-0-1 46.85 58.77
9 1200-0-0-2 46.85 47.88 3.75 54.99 57.03 3.34
10 1200-0-0-3 49.96 57.34
11 1200-4.75-0-1 61.31 66.77
12 1200-4.75-0-2 58.72 58.05 6.26 58.72 59.87 10.68
13 1200-4.75-0-3 54.13 54.13
14 1200-4.75-5.1-1 79.42 117.98
79.01 0.73 119.44 1.73
15 1200-4.75-5.1-2 78.60 120.90
16 1200-9-0-1 66.24 66.24
17 1200-9-0-2 58.87 62.73 5.89 58.87 62.73 5.89
18 1200-9-0-3 63.08 63.08
19 1200-9-5.1-1 80.09 123.77
81.96 2.09 137.09 9.89
20 1200-9-5.1-2 83.44 150.88
120

21 1200-9-5.1-3 82.34 136.61


22 1200-13.5-5.1-1 86.07 119.40
23 1200-13.5-5.1-2 86.84 85.83 1.33 140.06 139.55 14.26
24 1200-13.5-5.1-3 84.59 159.18
25 1200-18-5.1-1 74.84 150.65
26 1200-18-5.1-2 133.22 102.04 28.81 201.87 171.14 15.84
27 1200-18-5.1-3 98.04 160.90
28 1500-9-5.1-1 71.85 100.33
29 1500-9-5.1-2 105.07 90.31 18.73 143.68 121.24 17.91
30 1500-9-5.1-3 94.00 119.71
31 1500-13.5-5.1-1 99.99 148.47
32 1500-13.5-5.1-2 106.58 104.38 3.64 151.61 153.34 3.87
33 1500-13.5-5.1-3 106.58 159.96
34 1500-18-5.1-1 100.59 147.54
35 1500-18-5.1-2 101.79 107.97 10.90 174.34 161.34 8.32
36 1500-18-5.1-3 121.54 162.12
(Note: 1 kN = 0.2248 kip; 1 m = 3.28 ft;1 mm = 0.0394 in).

4.4.4 Stiffness of Synthetic Fiber Reinforced Concrete Pipes

Based on most US standards, pipes are considered flexible if they can deflect more

than 2% of their inside diameter. Most of the tested fiber reinforced pipes exhibited

deflection of more than 2% of their inside diameter with superior stiffness and retained

strength compared to other kinds of flexible pipes such as corrugated metal pipes and

high-density polyethylene pipes specified by AASHTO M294 (AASHTO 2011). The

results of all concrete pipes stiffness at 2% deflection are summarized in Table 22. The

COV of stiffness results ranged from 0.44 to 21.17%. Stiffness was calculated by

dividing the applied load by the corresponding deflection and the length of the pipe.
121

Table 22. Stiffness Results at 2% Deflection Ratio.

No Pipe Stiff. /length (kN/m/m) Average (kN/m/m) COV (%)

1 600-9-0-1 - 5204.25 -
2 600-9-0-2 5204.25
3 600-9-1.5-1 10379.91 10347.61 0.44
4 600-9-1.5-2 10315.31
5 600-18-0-1 7362.31 7144.68 4.31
6 600-18-0-2 6927.05
7 600-18-1.5-1 10578.41 10578.41 -
8 1200-0-0-1* - - -
9 1200-0-0-2* -
10 1200-0-0-3* -
11 1200-4.75-0-1* - 961.21 4.69
12 1200-4.75-0-2 993.1
13 1200-4.75-0-3 929.31
14 1200-4.75-5.1-1 5961.34 5909.10 1.25
15 1200-4.75-5.1-2 5856.85
16 1200-9-0-1* - 2125.10 -
17 1200-9-0-2* -
18 1200-9-0-3* 2125.1
19 1200-9-5.1-1 5921.01 6455.01 8.62
20 1200-9-5.1-2 7030.92
21 1200-9-5.1-3 6413.09
22 1200-13.5-5.1-1 5911.06 6621.16 11.45
23 1200-13.5-5.1-2 6533.01
24 1200-13.5-5.1-3 7419.41
25 1200-18-5.1-1 7167.06 7898.89 11.26
26 1200-18-5.1-2 8889.13
27 1200-18-5.1-3 7640.47
28 1500-9-5.1-1 4675.25 5825.66 21.17
29 1500-9-5.1-2 7127.91
30 1500-9-5.1-3 5673.81
31 1500-13.5-5.1-1 5747.01 6997.32 15.64
32 1500-13.5-5.1-2 7463.88
33 1500-13.5-5.1-3 7781.07
34 1500-18-5.1-1 7122.9 7923.35 9.39
35 1500-18-5.1-2 8593.21
36 1500-18-5.1-3 8053.93

*Unavailable data at 2% deflection ratio. (Note: 1 kN = 0.2248 kip; 1 m = 3.28 ft).


122

Inclusion of fiber with zero-slump concrete mixture improved the stiffness of

concrete pipes significantly due to fibers’ tendency to bridge cracks, and as result increased

load capacity. The pipes with a diameter of 600 mm (24 in.) reinforced with synthetic fiber

only exhibited a gain in the stiffness by 37.29% when fiber dosage increased from 9 kg/m3

to 18 kg/m3. This gain was less for pipes reinforced with one steel cage layer in addition to

synthetic fiber, where the gain was only 2.23% for the same increase in fiber dosage. This

is because for small diameter pipes, the primary mode of failure was flexural, and the

ductility of steel cage governed the failure behavior. The effect of increasing fiber dosage

on the stiffness of 1200 mm (48 in.) diameter concrete pipes showed greater improvement

in the pipe stiffness when fiber dosage increased. With synthetic fiber only, increasing fiber

dosage from 4.75 to 9 kg/m3 resulted in 121.09% gain in the stiffness. Moreover, for pipes

reinforced with synthetic fiber and one steel cage layer, using fiber dosage of 9, 13.5, and

18 kg/m3 (15, 22.5 and 30 lb/yd3) increased stiffness by 9.24, 12.05, and 33.67%,

respectively, compared to pipe reinforced with a steel cage layer and fiber dosage of 4.75

kg/m3. The effect of fiber dosage on the stiffness of 1500 mm pipe diameter was more

noticeable as strength requirements increased and more modes of failure came into play.

The stiffness of pipes with this diameter increased by 20.11% and 36.01% with fiber

dosage of 13.5 and 18 kg/m3, respectively, compared to pipes reinforced with fiber dosage

of 9 kg/m3. The effect of addition of fiber on the stiffness of concrete pipes of diameters

1200 and 1500 mm was more pronounced than for those of diameter 600 mm (24 in.).

Figure 40(a, b, and c) shows the ratio of ultimate deflection to the 2% deflection of

the pipe diameter (ΔUltimate/Δ2%) that the concrete pipes experienced under three-edge
123

bearing test. When (ΔUltimate/Δ2%) is 1 or greater, it means that the ultimate pipe stiffness is

equal or greater than its stiffness at 2% deflection, which means these concrete pipes have

more bending strength to resist the exerted load (Park, Y., 2015). In addition, the

surrounding soil could add support to the pipe to resist load. All fiber concrete pipes with

diameter of 600 and 1200 mm reinforced with synthetic fiber and a steel cage layer had

(ΔUltimate/Δ2%) greater than 1. The ratio of (ΔUltimate/Δ2%) for concrete pipe diameters of 1500

mm ranged from 0.63 to 1, only one had a ratio of one.

Incorporating synthetic fiber had a significant impact on the stiffness and retained

bending strength of all tested pipes. The effect of using synthetic fiber along with one

steel cage layer on concrete pipe stiffness was more pronounced for inner diameters of

1200 and 1500 mm (48 and 60 in.) than for 600 mm (24 in.). This difference is attributed

to the thickness of the pipe wall and the different failure modes associated with each pipe

diameter. modes associated with each pipe diameter.


124

2.2
5 1 600-9-0-1
2 7 2 600-9-0-2
1.8 3 6 3 600-9-1.5-1
1.6 4 4 600-9-1.5-2
5 600-18-0-1
1.4 6 600-18-0-2
Δult/Δ2%

1.2 7 600-18-1.5-1
1
0.8
0.6
0.4
0.2
0 1 2
600 mm (24 in.) pipe diameter

(a)

2.2
8 1200-0-0-1
2 9 1200-0-0-2
1.8 26 10 1200-0-0-3
1.6 21 23 11 1200-4.75-0-1
19 27 12 1200-4.75-0-2
1.4 20 24 25
22 13 1200-4.75-5.1-1
Δult/Δ2%

1.2 13 14 1200-4.75-5.1-2
14 15 1,200-4.75-5.1-3
1
16 1200-9-0-1
0.8 17 1200-9-0-2
0.6 18 1200-9-0-3
19 1200-9-5.1-1
0.4 20 1200-9-5.1-2
0.2 18 21 1200-9-5.1-3
8 9 1011 12 151617
0 22 1200-13.5-5.1-1
23 1200-13.5-5.1-2
1200 mm (48 in.) pipe diameter 24 1200-13.5-5.1-3
25 1200-18-5.1-1
26 1200-18-5.1-2
(b) 27 1200-18-5.1-3
125

2.2
28 1500-9-5.1-1
2
29 1500-9-5.1-2
1.8 30 1500-9-5.1-3
1.6 31 1500-13.5-5.1-1
32 1500-13.5-5.1-2
1.4
33 1500-13.5-5.1-3
Δult/Δ2%

1.2 34 1500-18-5.1-1
32
1 29 35 1500-18-5.1-2
33 35 36 1500-18-5.1-3
0.8 36
31 34
28 30
0.6
0.4
0.2
0
28 29 30 31 32 33 34 35 36 37 38
1500 mm (60 in.) pipe diameter

(c)

Figure 40. Summary of deflection ratio of fiber-reinforced concrete pipes.

4.5 Design Proposal for Synthetic Fiber Reinforced Concrete Pipes Using Finite

Element Analysis

4.5.1 Three-Edge Bearing Test Results

The load-deflection responses of all tested pipes are shown in Figs. 8-10. Each

pipe was given a label with four numerals. For example, in the label 600-9-150-1, the

first numeral represents the pipe diameter in mm (600 mm (24 in.) ), the second numeral

represents the fiber dosage in kg/m3 (9 kg/m3), the third numeral represents the amount of

steel reinforcement in mm2/m (150mm2/m; 0 indicates no steel), and the fourth numeral

represents the specimen number of that type (1 denotes the first of two pipes with these

properties). Incorporating PP fiber as a replacement for steel cage enhanced the first

crack and ultimate load capacity of tested pipes. As fiber dosage increased, both the first
126

crack and ultimate load increased. However, the size of the enhancement in the first crack

and ultimate loads capacities was influenced by the diameter and thickness of tested

pipes. For small diameter pipes, the failure mode was flexural, where the fibers were

effective in resisting only the flexural loads. As pipe diameter increased, other failure

modes such as radial and shear came into play, and the random distribution and

orientation of the fibers help the larger diameter pipes withstand these stresses as well.

Therefore, increasing fiber dosage on the large diameter pipes had a greater effect on

ultimate load results than for the smaller diameter pipes. For example, for the 600 mm

(24 in.) diameter pipes, increasing the fiber dosage from 9 to 18 kg/m3 resulted in a

4.95% gain in the ultimate load. In contrast, the ultimate load of the 1500 mm diameter

pipe increased by 23.34% for the same fiber dosages.

Because the pipes reinforced only with PP fiber could not meet the ASTM

strength requirements as both pipe diameter and ASTM Class strength requirements

increased (Al Rikabi, 2017), additional reinforcement in the form of a single steel mesh

layer located 25 mm (in.) from the inside diameter were tested. These pipes showed a

superior stiffness and could withstand deflection ratios of 2% of the pipe diameter.
127

350
600-9-0-1
600-9-0-2
300
600-18-0-1
250

Load (kN)
200

150

100

50

0
0 10 20 30 40
Deflection (mm)

(a)

350
600-9-150-1
600-9-150-2
300
600-18-150-1
250
Load (kN)

200

150

100

50

0
0 10 20 30 40
Deflection (mm)

(b)

Figure 41. Load deflection curves of 600 mm (24 in.) diameter pipe reinforced with (a)

PP fiber only; (b) PP fiber with one steel cage layer. (Note: 1 kN = 0.2248 kip; 1 mm =

0.0394 in).
128

350
1200-0-0-1
1200-0-0-2
300 1200-0-0-3

250

Load (kN.) 200

150

100

50

0
0 1 2 3 4 5
Deflection (mm)

(a)

350
1200-4.75-0-1
1200-4.75-0-2
300 1200-4.75-0-3
1200-9-0-1
250 1200-9-0-2
1200-9-0-3
Load (kN.)

200

150

100

50

0
0 5 10 15 20 25 30 35 40 45
Deflection (mm)

(b)
129

350
1200-4.75-510-1
300 1200-4.75-510-2

250

Load (kN.)
200

150

100

50

0
0 5 10 15 20 25 30 35 40 45
Deflection (mm)

(c)

350
1200-9-510-1
1200-9-510-2
300 1200-9-510-3
250
Load (kN.)

200

150

100

50

0
0 5 10 15 20 25 30 35 40 45
Deflection (mm)

(d)
130

350
1200-13.5-510-1
1200-13.5-510-2
300 1200-13.5-510-3
250

Load (kN)
200

150

100

50

0
0 10 20
Deflection (mm)30 40

(e)

350
1200-18-510-1
300 1200-18-510-2
1200-18-510-3
250
Load (kN)

200

150

100

50

0
0 5 10 15 20 25 30 35 40 45
Deflection (mm)

(f)

Figure 42. Load deflection curves of 1200 mm (48 in.) diameter pipe: (a) plain concrete

pipes; (b), pipes reinforced with PP fiber dosage of 4.75 and 9 kg/m 3 (8 and 15 lb/yd3); (c,

d, e, and f) pipes reinforced with a steel cage layer and PP fiber dosage of 4.75, 9, 13.5,

and 18 kg/m3 (8, 15, 22.5 and 30 lb/yd3), respectively. (Note: 1 kN = 0.2248 kip; 1 mm =

0.0394 in).
131

350

300

Load (kN.) 250

200

150

100

1500-9-510-1
50
1500-9-510-2
1500-9-510-3
0
0 10 20 30 40 50
Deflection (mm)

(a)

350

300

250
Load (kN)

200

150

100
1500-13.5-510-1
50 1500-13.5-510-2
1500-13.5-510-3
0
0 5 10 15 20 25 30 35 40 45
Deflection (mm)

(b)
132

350

300

250

Load (kN)
200

150

100
1500-18-510-1
50 1500-18-510-2
1500-18-510-3
0
0 5 10 15 20 25 30 35 40 45
Deflection (mm)

(c)

Figure 43. Load deflection curves of 1500 mm (60 in.) diameter pipe reinforced with one

steel cage layer and PP fiber dosage of: (a) 9 kg/m3 (15 lb/yd3);(b) 13.5 kg/m3(22.5

lb/yd3) ;(c) 18 kg/m3 (30 lb/yd3). . (Note: 1 kN = 0.2248 kip; 1 mm = 0.0394 in).

4.5.2 Material Properties

In order to provide the FE model in ABAQUS with concrete material properties, values for

compressive strength, tensile strength, modulus of elasticity, and Poisson’s ratio were

obtained experimentally and are shown in Table 6. Increasing PP dosage reduced the

compressive strength and modulus of elasticity of test specimens, while Poisson’s ratio and

splitting tensile strength increased. The trend in compressive strength agrees with past

research ( Zhang et al., 1999; Huang, 2001; Puertas et al., 2003; Choi and Yuan, 2005).

Therefore, in order to validate models mentioned previously used for PP fiber reinforced

concrete compression behavior, the dimensionless factor added to compressive strength


133

and strain of plain concrete was modified using a regression analysis of the compressive

strength results.

Table 23.The average compressive strength, tensile strength, modulus of elasticity and

Poisson’s ratio test results.

Fiber dosage, Splitting tensile


Compressive Modulus of
strength Poisson’s
Mix no. strength at 28 days, elasticity,
kg/m 3
strength at 28 days, ratio
MPa (ksi) MPa (ksi)
(lb/yd )3
MPa (ksi)
1 0 3.58 60.70 (8.860) 36618 (5311) 0.13
2 4.75 (8) 3.56 53.35 (7.73) 35925 (5210) 0.142
3 6 (10) 3.59 51.79 (7.51) 34155 (4954) 0.185
5 7 (12) 3.74 50.22 (7.28) 31064 (4505) 0.192
6 9 (15) 3.93 47.35 (6.86) 30333 (4400) 0.253

The stress-strain curves for the fiber reinforced concrete cylinders were compared

to the models given in past research (Soroushian and Lee 1989; Ezeldin and Balaguru 1992;

Barros and Figueiras 1999; Nataraja et al. 1999), as shown in Figure 44(a, b, and c). As

can be seen from Figure 44, the modified stress-strain relationship proposed by Nataraja et

al. (1999) showed the most similar behavior to the experimental stress-strain response in

this research. Therefore, this model was adopted to describe the compression behavior of

synthetic fiber reinforced concrete. This model includes the following relationships.

𝜀
𝛽 (𝜀 )
𝜎 𝑝𝑓
=
𝑓𝑐𝑓 𝜀 𝛽
𝛽 − 1 + (𝜀 )
𝑝𝑓

with
134

𝑓𝑐𝑓 = 𝑓𝑐 − 19.907 (𝑅𝐼)

𝜀𝑝𝑓 = 𝜀𝑐𝑜 + 965 × 10−6 (𝑅𝐼)

𝛽 = 0.5811 + 1.93 (𝑅𝐼)−0.7406

𝑊𝑓 𝐿𝑓
𝑅𝐼 =
𝐷𝑓

Where fc is the compressive strength of plain concrete, εco is the strain the peak compressive

strength of plain concrete, RI is the reinforcing index, Wf is the fiber weight percentage in

the mixture, Lf is the fiber length, and Df is the fiber diameter. Figure 45shows the

compressive stress-strain plots used in the FEM based on the modified Nataraja et al.

(1999) model.
135

Figure 44. Comparison of experimental stress-strain curves of concrete cylinders

reinforced PP with dosage of (a) 4.75 kg/m3 (b) 6 kg/m3; (c) 7 kg/m3; and (d) 9 kg/m3

with various models of compressive stress response from the literature.


136

60
4.75 kg/m3 6 kg/m3
7 kg/m3 9 kg/m3
50 13.5 kg/m3 18 kg/m3

40

Stress (MPa)
30

20

10

0
0 0.002 0.004 0.006 0.008 0.01 0.012
Strain (mm/mm)

Figure 45. compressive stress-strain plots used in the FE models.

The combined reduction in the modulus of elasticity and an increase in Poisson’s

ratio indicated that the synthetic fibers increased the flexibility of the concrete material.

This may be attributed to the ductile behavior of synthetic fiber compared to the plain

concrete since fiber has a lower modulus of elasticity than plain concrete. Furthermore,

adding synthetic fiber to the concrete mixture enhanced tensile strength. The Poisson’s

ratio for fiber dosages over 9 kg/m3 (15 lb/yd3) was calculated using a regression analysis

of the Poisson’s ratio results, yielding Equation 6.

𝜈𝑓 = 𝜈𝑜 + 0.174 (𝑅𝐼 ) (6)

where 𝜈𝑓 is the Poisson’s ratio of fiber reinforced concrete, and 𝜈𝑜 is the Poisson’s ratio

of plain concrete.
137

4.5.3 Finite Element Results and Discussion

Model Calibration

The finite element (FE) model was calibrated using the results of concrete pipes

with a diameter of 1200 mm (48 in.) in terms of load-deflection response, ultimate load

capacity, and failure mode. surface-to-surface tie constraint was used to attach the

concrete pipe model to the lower and upper strips. The concrete pipe external surface was

selected as a master surface while the lower and upper striped surfaces were selected as

slave surfaces. Concrete pipes reinforced with dosages of PP fiber 9, 13.5, and 18 kg/m3

(15, 22.5, and 30 lb/yd3) along with one steel layer area of 510 mm2/m were simulated.

Also, concrete pipe with a diameter of 1200 mm (48 in.) reinforced with fiber dosage of

9 kg/m3 as sole reinforcement was calibrated. To calibrate the response of fiber

reinforced concrete pipes, a tensile strength parameter was considered. According to past

research (Mobasher, 2011; Blazejowski, 2012), this parameter along with modulus of

elasticity have the most significant impact on the FE results. After several trials, a good

agreement was obtained between experimental and FE results as the specimens showed

the same load-deflection response, as shown in Figure 46.


138

300
1200-9-0-1
1200-9-0-2
250 1200-9-0-3
FE

200

Load (kN)
150

100

50

0
0 10 20 30 40 50
Deflection (mm)

(a)

300

250

200
Load (kN)

150

100
1200-9-510-1
1200-9-510-2
50 1200-9-510-3
FE
0
0 10 20 30 40 50
Deflection (mm)

(b)
139

300

250

200

Load (kN)
150

100
1200-13.5-510-1
50 1200-13.5-510-2
1200-13.5-510-3
FE
0
0 10 20 30 40 50
Deflection (mm)

(c)

300

250

200
Load (kN)

150

100
1200-18-510-1
50 1200-18-510-2
1200-18-510-3
FE
0
0 5 10 15 20 25 30 35 40 45
Deflection (mm)

(d)

Figure 46. Experimental versus calibrated FE load-deflection responses of 1200 mm (48

in.) diameter pipe reinforced with fiber dosage of (a) 9 kg/m3 (15 lb/yd3), (b) 9 kg/m3 (15

lb/yd3) with 510mm2/m (0.24 in2/ft) steel area (c) 13.5 kg/m3 (22.5 lb/yd3)with 510

mm2/m (15 lb/yd3) steel area (d) 18 kg/m3 (30 lb/yd3) with 510mm2/m (0.24 in2/ft) steel

area. . (Note: 1 kN = 0.2248 kip; 1 mm = 0.0394 in).


140

The same mode of failure was observed in both the experiments and in the

corresponding finite element simulation, as shown in Figure 47 where both the tested and

simulated pipes experienced a progressive flexural failure that started at the invert and the

crown (see Figure 47(b)) as ultimate tensile strength of the pipe was reached, followed by

cracking at the spring lines (see Figure 47(c)).


141

Figure 47. Crack propagation (a) experimental results (b) Maximum principal strain

contour at crown and invert (c) Maximum principal strain contour at the crown, invert,

and springline (c,d, and e) cracks propagation.


142

Furthermore, the ultimate strengths of the calibrated pipes were compared to those

experimentally tested. The results were summarized in Table 24. As can be seen, the

difference between experimental and FE ultimate load ranged from -0.2 to -8.8 %, making

the FE model more conservative than the experiment.

Table 24. FE and experimental results of ultimate strength of the calibrated pipes.

Experimental ultimate
FE ultimate load,
Pipe designation load, Error (%)
kN
kN

1200-9-0 89.2 88.9 -0.2

1200-9-5.1 190.3 181.6 -4.5

1200-13.5-5.1 205.4 204.8 -0.3

1200-18-5.1 246.8 224.9 -8.8

(Note: 1 kN = 0.2248 kip; 1 mm = 0.0394 in)

Model Verification

The calibrated models of 1200 mm (48 in.) diameter concrete pipe reinforced

with steel and fiber at selected dosages were validated using the experimental results of

1500 mm diameter pipes reinforced with the same fiber dosage and steel area. Also, the

calibrated model of 1200 mm (48 in.) diameter pipe reinforced with fiber dosage of 9

kg/m3 (15 lb/yd3) was validated using experimental results of 600 mm (24 in.) diameter

pipe with the same fiber dosage. The calibrated models were validated in terms of load-

deflection response and ultimate load capacity. A good agreement was observed between
143

experimental and FE load-deflection response observed in elastic and post-cracking

stages, as shown in Figure 48.

350
600-9-0-1
600-9-0-2
300 FE

250
Load (kN)

200

150

100

50

0
0 5 10 15 20 25 30 35 40 45 50
Deflection (mm)

(a)

350

300

250
Load (kN.)

200

150

100
1500-9-510-1
1500-9-510-2
50
1500-9-510-3
FE
0
0 5 10 15 20 25 30 35 40 45 50
Deflection (mm)

(b)
144

350

300

250

Load (kN)
200

150

100
1500-13.5-510-1
1500-13.5-510-2
50
1500-13.5-510-3
FE
0
0 5 10 15 20 25 30 35 40 45 50
Deflection (mm)

(c)

350

300

250
Load (kN)

200

150

100
1500-18-510-1
1500-18-510-2
50 1500-18-510-3
FE
0
0 5 10 15 20 25 30 35 40 45 50
Deflection (mm)

(d)

Figure 48. Experimental versus FE load-deflection responses: a) 600 mm (24 in.)

diameter pipe reinforced with 9 kg/m3; (b), (c) and (d) 1500 mm diameter pipe reinforced

with steel area of 510 mm2/m and fiber dosage of 9,13.5 and 18 kg/m3, respectively.

.(Note: 1 kN = 0.2248 kip; 1 mm = 0.0394 in).


145

Moreover, the ultimate load of FE models was compared to the corresponding

experimental results and summarized in Table 8. The ultimate load predicted by the FE

model fell in the conservative side with values less than the experiment. With this

validation, the parametric study including different pipe diameters and thickness used the

same fiber dosages as the validated pipes.

Table 25. FE and Experimental Results of Ultimate Strength of the Validate Models.

Pipe designation Experimental ultimate


FE ultimate load Difference
load,
kN (kip) (%)
kN (kip)

600-90-0 83.7 (18.82) 83.1 (18.68) -0.7

1500-9-5.1 229.1 (51.5) 222.7 (50.06) -2.7

1500-13.5-5.1 275.6 (61.96) 262.3 (58.97) -4.8

1500-18-5.1 286 (64.3) 285.3 (64.14) -0.2

Parametric Study

In an attempt to provide engineers with tables that can be used for design

purposes, concrete pipes with different diameters (D) and wall thicknesses (h) were

simulated to determine the ultimate strength. Based on the ultimate strength results, pipes

were classified as Class I through Class V as specified by ASTM C76 (2015). As

mentioned previously, a significant enhancement in maximum load capacity and post-

cracking behavior of tested pipes using fiber dosage of 9 kg/m3 (15 lb/yd3) (1% by

volume) or more. Also, it was observed that using synthetic fiber as sole reinforcement

did not comply with ASTM C76 (2015) specifications as pipe diameter and strength
146

requirements increase. Therefore, pipes were simulated using fiber dosage over 9 kg/m 3

(15 lb/yd3) with and without additional steel cage reinforcement. The steel cage

reinforcement was simulated as a single layer located 25 mm (1 in.) from the inside

surface of the pipe.

The outcomes of the parametric study are summarized in Tables 26-30. For Class

I Wall A pipes, using a fiber dosage of 9 kg/m3 (15 lb/yd3) was sufficient to fulfill the

strength requirement for pipes with diameters ranging from 300 to 750 mm (12 to 30 in.).

Pipes with diameters beyond this range were reinforced with an additional steel cage

meeting the minimum requirements specified by ASTM C76, 150 mm 2/m (0.07 in2/ft), to

fulfill the strength requirements of Class I. Pipes with type B wall achieved the strength

requirement of Class I using fiber dosage of 9 kg/m3, except for pipes with a diameter of

1200 mm (48 in.), which required additional steel cage reinforcement to comply with

strength requirements. Pipes with type C wall reinforced with fiber dosage of 9 kg/m3 (15

lb/yd3)fulfilled the strength requirements of Class I and II. In order to achieve the

strength requirements of higher classes, additional reinforcement from either fiber or steel

cage was needed.


147

Table 26. Design requirements for Class I synthetic fiber reinforced concrete pipes.

Wall A Wall B Wall C


D, mm (in.)
h steel area Wf h steel area Wf h steel area Wf
mm mm2/m kg/m3 mm mm2/m (kg/m3) mm mm2/m (kg/m3)
300 44 - 9 50 - 9 69 - 9
375 47 - 9 57 - 9 75 - 9
450 50 - 9 63 - 9 82 - 9
525 57 - 9 69 - 9 88 - 9
600 63 - 9 75 - 9 94 - 9
675 66 - 9 82 - 9 100 - 9
750 69 - 9 88 - 9 107 - 9
825 72 150 9 94 - 9 113 - 9
900 75 150 9 100 - 9 119 - 9
1050 88 150 13.5 113 - 9 132 - 9
1200 100 150 18 125 - 9 144 - 9
1350 113 150 18 138 - 9 157 - 9
1500 125 150 18 150 150 9 169 - 9
(Note: 1 kg/m = 1.68 lb/yd ; 1 mm = 0.0394 in).
3 3
148

Table 27. Design requirements for class II synthetic fiber reinforced concrete pipes.

Wall A Wall B Wall C


D (mm) h steel area Wf h steel area Wf h steel area Wf
mm mm2/m (kg/m3) mm mm2/m (kg/m3) mm mm2/m (kg/m3)
300 44 - 9 50 - 9 69 - 9
375 47 - 9 57 - 9 75 - 9
450 50 - 9 63 - 9 82 - 9
525 57 - 9 69 - 9 88 - 9
600 63 - 9 75 - 9 94 - 9
675 66 150 9 82 - 9 100 - 9
750 69 150 9 88 - 9 107 - 9
825 72 150 13.5 94 - 9 113 - 9
900 75 150 18 100 - 9 119 - 9
1050 88 190 18 113 150 9 132 - 9
1200 100 210 18 125 150 13.5 144 - 9
1350 113 250 18 138 150 13.5 157 - 9
1500 125 320 18 150 150 13.5 169 - 9
(Note: 1 kg/m = 1.68 lb/yd ; 1 mm = 0.0394 in)
3 3
149

Table 28. Design requirements for class II synthetic fiber reinforced concrete pipes.

D (mm) Wall A Wall B Wall C


h steel area Wf h steel area Wf h steel area Wf
mm mm2/m kg/m3 mm mm2/m kg/m3 mm mm2/m kg/m3
300 44 - 9 50 - 9 69 - 9
375 47 - 9 57 - 9 75 - 9
450 50 - 9 63 - 9 82 - 9
525 57 150 9 69 - 9 88 - 9
600 63 150 9 75 - 9 94 - 9
675 66 150 13.5 82 - 9 100 - 9
750 69 150 18 88 - 9 107 - 9
825 72 230 18 94 150 14 113 - 9
900 75 280 18 100 150 14 119 - 9
1050 88 360 18 113 150 18 132 - 9
1200 100 420 18 125 170 18 144 150 13.5
1350 113 470 18 138 250 18 157 150 13.5
1500 125 570 18 150 310 18 169 150 18
(Note: 1 kg/m = 1.68 lb/yd ; 1 mm = 0.0394 in)
3 3
150

Table 29. Design requirements for Class IV synthetic fiber reinforced concrete pipes.

D (mm) Wall A Wall B Wall C


h steel area Wf h steel area Wf h steel area Wf
mm mm2/m (kg/m3) (mm) mm2/m (kg/m3) (mm) mm2/m (kg/m3)
300 44 - 9 50 - 9 69 - 9
375 47 150 9 57 - 9 75 - 9
450 50 150 9 63 - 9 82 - 9
525 57 150 13.5 69 150 9 88 - 9
600 63 210 13.5 75 150 9 94 - 9
675 66 300 18 82 150 18 100 - 9
750 69 380 18 88 150 18 107 150 9
825 72 420 18 94 230 18 113 150 13.5
900 75 570 18 100 280 18 119 150 18
1050 88 720 18 113 380 18 132 170 18
1200 100 850 18 125 460 18 144 280 18
1350 113 960 18 138 550 18 157 340 18
1500 125 1060 18 150 660 18 169 470 18
(Note: 1 kg/m = 1.68 lb/yd ; 1 mm = 0.0394 in)
3 3
151

Table 30. Design requirements for Class V synthetic fiber reinforced concrete pipes.

D (mm) Wall A Wall B Wall C


h steel area Wf h steel area Wf h steel area Wf
(mm) mm2/m (kg/m3) (mm) mm2/m (kg/m3) (mm) mm2/m (kg/m3)
300 44 - 9 50 - 9 69 - 9
375 47 150 9 57 150 9 75 - 9
450 50 210 9 63 150 9 82 - 9
525 57 210 18 69 150 9 88 - 9
600 63 300 18 75 150 9 94 150 9
675 66 490 18 82 300 18 100 150 13.5
750 69 570 18 88 250 18 107 150 13.5
825 72 550 18 94 380 18 113 190 18
900 75 820 18 100 490 18 119 270 18
1050 88 980 18 113 570 18 132 300 18
1200 100 1160 18 125 700 18 144 470 18
1350 113 1220 18 138 810 18 157 530 18
1500 125 1530 18 150 950 18 169 700 18
(Note: 1 kg/m = 1.68 lb/yd ; 1 mm = 0.0394 in)
3 3

To compare the FE analysis results of one of the concrete pipe configurations with

the strength requirement specified for each ASTM class, the load-deflection responses of

1350 mm diameter pipe were compared to 0.3 mm cracking load and ultimate load of each

class, as shown in Fig. 16 (a) and (b), respectively. It can be deduced that increasing fiber

dosage only is not always enough to meet the strength requirement for the targeted class.

For example, increasing fiber dosage from 13.5 to 18 kg/m3 along with the same steel area

was not enough to fulfill the strength requirement of Class III. Instead, the only factor

changed was the steel area to fulfill the strength requirements for Classes III, IV, and V.
152

200
190
180
170
160
150 Class V
140
D-Load (kN/m/m)

130
120
110 Class IV
100
90
80
70 Class III 1350-9-0
60 Class II
50 1350-9-1.5
40 Class I 1350-13.5-1.5
30 1350-18-1.5
20 1350-18-2.5
10 1350-18-5.5
0
0 5 10 15 20 25 30 35 40 45 50
Deflection (mm)

(a)

200
190
180 Class V
170
160 Class IV
150
140
D-Load (kN/m/m)

130
120
110 Class III
100
90
80 Class II
70 Class I
60 1350-9-0
50 1350-9-1.5
40 1350-13.5-1.5
30 1350-18-1.5
20 1350-18-2.5
10 1350-18-5.5
0
0 10 20Deflection (mm)30 40 50

(b)

Figure 49. Load-deflection responses of 1350 mm (54 in) (B wall) pipe diameter with

different fiber dosage and steel cage areas imposed on (a) 0.3 mm (0.01 in.) and (b)

ultimate loads specified in ASTM C76 for pipe strength classes. (Note: 1 kN = 0.2248

kip; 1 mm = 0.0394 in.) .


153

The minimum fiber dosage used in the design tables as sole reinforcement was 9

kg/m3. As strength requirements increase for higher ASTM classes, the reinforcement

was augmented with one layer of a steel cage. The maximum fiber dosage utilized in this

study was 18 kg/m3, to avoid problems with the workability of the concrete mix at high

fiber dosages. It may be preferred to use a pipe that includes minimum steel cage

reinforcement, 150 mm2/m as specified by ASTM C76, along with fiber dosage of

9kg/m3

4.6 Experimental Investigation of Thin-Wall Synthetic Fiber Reinforced Concrete

Pipes

4.6.1 Load-Deflection Response

The deflection responses of 1200 and 1500 mm (48 and 60 in.) concrete pipes

under load are depicted in Figure 50and Figure 51, respectively. The vertical (horizontal)

deflection is the average of readings from the two string potentiometers installed in the

vertical (horizontal) direction. Note that the vertical deflections (solid lines) represent a

decrease in vertical pipe diameter, while the horizontal deflections (dashed lines)

represent an increase in the horizontal pipe diameter. The vertical lines in each graph

mark 5% deflection, which represents the maximum allowed deflection in flexible pipe

installations (e.g. polyethylene pipes). For the 1200 mm (48 in.) diameter concrete pipe,

Figure 50 (a) shows the deflection behavior of pipe reinforced with 9 kg/m 3(15 lb/yd3) PP

fiber and a steel cage of area 5.7 cm2/m (0.27 in2/ft) under load. The deflection increased

approximately linearly with load up to about 10 kN (2.248 kip), where the line changes

slope, indicating the first crack in the pipe. Rather than rapidly failing, the deflection
154

increased with applied load (as can be seen by the lower slope in the graph), reflecting

the bridging of cracks in the pipe by the PP fibers and steel cage. The ultimate load

appears at the peak of the curve, which for vertical deflection was about 50 kN (11.25

kip) at a deflection of about 55 mm (2.16 in.), just shy of the 5% deflection line, while the

horizontal deflection at ultimate load was approximately 75 mm (2.95 in.), above the 5%

line. When the steel area was increased to 10.2 cm2/m (0.48 in2/ft), the slope of the load-

deflection curve increased, as shown in Figure 50 (b). The increase in the slope of load-

deflection curve associated with steel area increasing indicated a gain in pipe stiffness.

Using a steel reinforcement area of 10.2 cm2/m (0.48 in2/ft) governed the load-defection

behavior after first crack (at about 20 kN [4.5 kip]) as the load-deflection response

continued to be linear. Increasing the steel area from 5.7 cm2/m to 10.2 cm2/m (0.27 to

0.48 in2/ft) increased the cracking load (which produced 0.3 mm crack width) and

ultimate load by 88% and 120%, respectively, and the ultimate loads for the latter were

well past the 5% deflection line.

For concrete pipes with a diameter 1500 mm (60 in.), Figure 51 (a) and (b) shows

the load deflection behavior of pipe reinforced with a fiber content of 9 kg/m3(15 lb/yd3)

and steel cage of area 5.7 cm2/m (0.27 in2/ft). The deflection behavior under load was

linear until the appearance of the first crack at just under 20 kN. Under greater loads, the

deflection increased following a non-linear curve to about 85 mm, above the 5% line.

Increasing steel area to 8.9 cm2/m (0.42 in2/ft) increased the cracking and ultimate loads

by 36% and 21%, respectively, as shown in Figure 51(b).


155

Incorporating PP fiber in the concrete material enhances the non-linear structural

behavior of the pipe and increases the post-cracking residual strength. The bridging

effects of fiber play a major role in absorbing energy, resisting the propagation of cracks,

and limiting their width (Banthia and Dubey, 2000; Li and Maalej, 1996; Mobasher and

Li, 1996; Zollo, 1997). This is mainly influenced by the number of fibers crossing the

cracks and bond strength between the concrete matrix and the fibers. As the surface area

of the fiber and number of fibers bridging the cracks increase, the bond strength between

fiber and concrete components increases, resulting in more stress transmitted by the fiber

when cracks are initiated. Comparing the load deflection responses of the tested pipes

showed that concrete pipes with a diameter of 1200 mm (48 in.) were stiffer than those

with a diameter of 1500 mm (60 in.). This may be attributed to the number of fibers

engaged in resisting the initiated cracks, which is a function of pipe wall thickness. As

pipe wall thickness increases, the total number of fibers across the thickness of the wall

increases. As a result, the number of fibers expected to bridge the propagated cracks

increases. Also, the fibers contribute significantly to the response of the pipes with a

larger diameter as failure modes such as radial and shear become more critical. This can

be attributed to fiber dispersion in the concrete pipe wall that bridged cracks propagated

due to these failure modes as well as flexural cracks. On the other hand, steel

reinforcement cage contributes only to the flexural strength of the pipe because it is

parallel to the pipe wall. Thus, concrete pipes with a diameter of 1500 mm (60 in.)

showed a better performance in terms of load-deflection behavior than 1200 mm (48 in.)

pipe diameter. Also, it was observed that pipes reinforced with greater steel area
156

exhibited stiffer behavior after the first crack as steel reinforcement started to govern the

load-deflection response. Thus, there is a need to balance between steel reinforcement

area and flexibility from synthetic fibers for each pipe diameter.

All tested pipes exhibited deflection, which exceeded 5% of their inside diameter

with high load carrying capacity and greatly exceeded the 2% deflection criterion for

flexible pipes, which may indicate that fiber reinforced concrete pipes may take

advantage of soil-structure interaction to carry more load. Therefore, buried pipes with

wall thickness less than that specified by ASTM C76 (2015) reinforced with synthetic

fiber along with minimum steel cage area can resist greater loads.

Using fiber dosage of 9 kg/m3(15 lb/yd3) was necessary to ensure the maximum

benefit of fiber contribution to the ultimate strength and avoid workability problems

associated with concrete using higher fiber dosage. Comparing the ultimate strength

results of 1500 mm (60 in.) pipe diameter with past research (Park et al, 2016) showed

that the ultimate strength increased by 11.5% as fiber content was increased from 4.75 to

9 kg/m3 (8 to 15 lb/yd3). The findings of this research concur with past research results

that including synthetic fiber enhances the overall performance of concrete pipes

(Peyvandi et al., 2013 and 2014; Fuente et al., 2014; Abolmaali et al., 2014; Wilson et

al.,2014; Park et al, 2016, Al Rikabi et al., 2017).


157

120
Vertical deflection
Horizontal deflection
100

80

Load (kN)
60

40

20 5% deflection ratio

0
0 20 40 60 80 100
Deflection (mm)

(a)

120
Vertical deflection
Horizontal deflection
100

80
Load (kN)

60

5% deflection ratio
40

20

0
0 20 40 60 80 100
Deflection (mm)

(b)

Figure 50. Load deflection curves of 1200 mm (48 in.) diameter pipes reinforced with PP

fiber dosage of 9 kg/m3 along with reinforcing steel area of: (a) 5.7 cm2/m (0.27 in2/ft);

(b) 10.2 cm2/m (0.48 in2/ft). (Note: 1 kN = 0.2248 kip; 1 mm = 0.0394 in.).
158

120
Vertical Deflection
Horizontal deflection
100

80

Load (kN)
60

5% deflection ratio
40

20

0
0 20 40 60 80 100 120
Deflection (mm)

(a)

120
Verical deflection
Horizontal deflection
100

80
Load (kN)

60

40
5% deflection ratio
20

0
0 20 40 60 80 100 120
Deflection (mm)

(b)

Figure 51. Load deflection curves of 1500 mm (60 in.) diameter pipes reinforced with PP

fiber dosage of 9 kg/m3 along with reinforcing steel area of: (a) 5.7 cm2/m (0.27 in2/ft);

(b) 8.9 cm2/m (0.42 in2/ft). (Note: 1 kN = 0.2248 kip; 1 mm = 0.0394 in.).
159

4.6.2 Load-Strain Response

The strain was measured at the inner crown, inner invert, inner springline, and

outer springline of concrete pipe with a diameter of 1200 and 1500 mm (48 and 60 in.)

using strain gauges labeled SG#1, SG#4, SG#2, and SG#3, respectively. The load versus

strain measurements are depicted in Figure 52. Due to the flexural deformation of tested

pipes, the measured strain at the inner crown, inner invert, and outer of springline were

positive (tensile strain). The strain measured on the inner surface of the springline was

negative (compressive strains). This strain distribution was consistent with the oval-

shaped response reported in relation to pipe deflections in the previous section. All tested

pipes exhibited linear load-strain relationships before cracking began. As cracks started to

propagate, the load-strain relationship became non-linear, as did the load-deflection

response. For 1200 mm (48 in.) concrete pipe diameter reinforced with 5.7 cm2/m (0.27

in2/ft) steel area, the maximum elastic tensile strain within the elastic range was 300 με.

The ultimate tensile strains recorded at the inner crown, inner invert, and outer springline

were 3500, 2830, and 2105 με, respectively. SG#1 mounted on the inner crown failed

after reaching its maximum strain capacity when one of the propagated cracks passed

under the strain gauge at a load of 20 kN (4.496 kip), as shown in Figure 52 (a). Thus,

this strain gauge did not measure the strain corresponding to the ultimate load. The

maximum compressive strain recorded on the inner surface of springline was -1950 με.

For 1200 mm (48 in.) pipe diameter reinforced with 10.2 cm2/m (0.48 in2/ft) area

steel, the maximum tensile elastic strain was 400 με at the outer face of the springline, as

shown in Figure 52(b). The maximum tensile strains recorded on the surface of the inner
160

crown, inner invert, and outer springline were 2006, 3410, and 2875 με, respectively. A

sudden decrease in the tensile strains at the crown and invert was observed when load of

13.5 kN (3.035 kip) and 20.3 kN (4.56 kip) was reached, respectively, as shown in Figure

52 (a) and (c). This decrease comes from the stress distribution associated with crack

propagation. Then, each of these strain gauges experienced failure as a crack propagated

underneath, causing the gauge to fail. The strain gauges affixed on the outer face at the

springline continued to record the strain as the cracks propagated away from the strain

gauge positions. The maximum compressive strain was -2665 με. Comparing the strain

results of both pipes with a diameter of 1200 mm (48 in.) showed that pipe reinforced

with 10.2 cm2/m (0.48 in2/ft) steel area experienced more strain at the inner crown and

the springline than pipe reinforced with 5.7 cm2/m (0.27 in2/ft) steel area. This high strain

capacity may be attributed to the increase in the steel area, which increases the

confinement of the concrete. Comparing the strain results at the crown showed that strain

experienced by pipe reinforced with steel area of 5.7 cm2/m (0.27 in2/ft) was higher than

that for the pipe reinforced with steel area 10.2 cm2/m (0.48 in2/ft). This can be attributed

to the crack propagation pattern, which started away from the strain gauge location,

causing an abrupt decrease in strain reading with a load of 13.5 kN (3.1 kip), as shown in

Figure 52(a).
161

120
5.7cm2/m
100 10.2 cm2/m
80

Load (kN)
60

40

20

0
0 1000 2000 3000 4000
Strain (με)

(a)

120
5.7 cm2/m
10.2 cm2/m
100

80

Load (kN)
60

40

20

0
-4000 -3000 -2000 -1000 0
Strain (με)

(b)

120
5.7 cm2/m
100 10.2 cm2/m
80
Load (kN)

60

40

20

0
0 1000 2000 3000 4000
Strain (με)

(c)
162

120

100

80

Load (kN)
60

40

20 5.7 cm2/m
10.2 cm2/m
0
0 1000 2000 3000 4000
Strain (με)

(d)

Figure 52. Strain measurements of pipe with diameter of 1200 mm (48 in) reinforced

with PP fiber dosage of 9 kg/m3(15 lb/yd3) along with steel mesh at amounts indicated in

the legend: (a) inner crown (b) inner springline (c) inner invert (d) outer springline.

(Note: 1 kN = 0.2248 kip).

For 1500 mm (60 in.) pipe diameter reinforced with steel area of 5.7 cm2/m (0.27

in2/ft) (see Figure 53), the maximum linear tensile strain is 155 με. Also, the maximum

compressive strain within the linear part of the load -strain response is -85 με (Figure

53(b)). Then, non-linear behavior was observed as the cracks started to appear on the

concrete pipe wall. The maximum strains recorded on the inner crown, inner invert, inner

springline, and outer springline were 1000, 2586, -1543, and 256 με, respectively. As the

steel area increased to 8.9 cm2/m (0.42 in2/ft), the strain capacity of the pipe increased.

The ultimate strains recorded on the surface of the inner crown, inner invert, inner

springline, outer springline were 3500, 2443, -1820, and 2412 με, respectively. Also, all

strain-load curves showed the same trend after the first crack at a load of 16 kN (3.6-kip).
163

As the applied load increased, an abrupt decrease appears in measurements from the

strain gauge located on the inner surface of the pipe invert (Figure 53(c)). This was likely

because a longitudinal crack passed through the gauge. Concrete pipes reinforced with

synthetic fiber can endure more strain (tensile and compressive) when the steel area was

increased. Using a steel reinforced cage along synthetic fiber increased the restrained

concrete material and, as a result, enhanced strain capacity. Also, strain readings at the

compression vicinity were more stable than those at the tension region as no cracks were

initiated at the compression zone.

120
5.7 cm2/m
100 8.9 cm2/m

80
Load (kN)

60

40

20

0
0 1000 2000 3000 4000
Strain (με)

(a)

120
5.7 cm2/m
10.2 cm2/m 100

80
Load (kN)

60

40

20

0
-4000 -3000 -2000 -1000 0
Strain (με)
164

(b)

80
5.7 cm2/m
70 8.9 cm2/m
60
50

Load (kN)
40
30
20
10
0
0 1000 2000 3000
Strain (με)

(c)

80
5.7 cm2/m
70 8.9 cm2/m
60
50
Load (kN)

40
30
20
10
0
0 1000 2000 3000
Strain (με)

(d)

Figure 53. Strain measurements of pipe with diameter of 1500 (60 in.) mm reinforced

with PP fiber dosage of 9 kg/m3 (15 lb/yd3) along with steel mesh at amounts indicated in

the legend: (a) inner crown (b) inner springline (c) inner invert (d) outer springline.

(Note: 1 kN = 0.2248 kip).


165

4.6.3 Failure Mechanism

Crack growth and propagation were monitored during the test to determine the

failure mechanism associated with each pipe configuration. The main observed failure

mode was flexural, where cracks initiated at the inside face of the crown and invert as

well as the outer face of the springlines. Other failure modes such as radial and shear

failures were also observed, where the cracks propagated radially and diagonally at the

invert and crown of some tested pipes, respectively, as shown in Figures 54, 55, 56 and

57.

For concrete pipe with a diameter of 1200 mm (48 in.) reinforced with steel area

of 5.7 cm2/m (0.27 in2/ft) (Figure. 54), cracks first appeared on the outside face at the

springline with a load of 6.6 kN (1.5 kip) (Figure. 54c), and subsequently initiated at the

inner face at the crown (Fig. 9a) and invert (Figure. 54b). The number of cracks increased

as the applied load increased. On the other hand, for 1200 mm (48 in.) diameter pipe

reinforced with 10.2 cm2/m (0.48 in2/ft) (see Figure. 55), multiple hairline cracks initially

appeared at the crown (Figure. 55a) and invert (Figure. 55b) and then at the springlines

(Figure. 55c). The number of propagated cracks for 1200 mm (48 in.) diameter pipe

reinforced with steel area of 5.7 cm2/m (0.27 in2/ft) was higher than that reinforced with

steel area of 10.2 cm2/m (0.48 in2/ft). Also, the latter pipe experienced larger cracks than

the former. Thus, as the steel area increased, pipes become stiffer and the failure

mechanism tended to be governed by the steel cage rather than the synthetic fiber.

Therefore, it is possible to adjust the steel reinforcement content to achieve a desired

balance between flexibility and serviceability.


166

For 1500 mm (60 in.) diameter pipe reinforced with steel area of 5.7 cm2/m (0.27

in2/ft) (see Figure. 56), the first crack initiated at the inner face of the invert (Figure. 56b).

As the applied load increased, more hairline cracks initiated at the inner face of the crown

(Figure. 56a) and the outer face at the springlines (Figure. 56c). On other hand, the same

diameter pipe reinforced with a steel of area 8.9 cm2/m (0.42 in2/ft) experienced a

different crack pattern (see Error! Reference source not found.) as the first crack

initiated at the outer face of the springline. This also can be observed from Figure 53(d)

and Figure 50(b) as the pipe showed flexible load-strain/deflection responses in the

springline compared to other pipes. Moreover, the observed cracks were smaller and

greater in number than those observed in the pipe reinforced with a steel area of 5.7

cm2/m (0.27 in2/ft). After the flexural cracks first appeared, other crack types such as

radial and shear cracks were observed at a load of 60 kN (13.5 kip). These cracks were

formed mainly at the invert and crown of tested pipes. It should be highlighted that pipes

of 1500 (60 in.) diameter reinforced with 8.9 cm2/m (0.42 in2/ft) experienced shear and

radial cracks larger than those observed in pipes reinforced with steel area of 5.7 cm2/m

(0.27 in2/ft).

(a) (b) (c)


167

Figure. 54. Crack development in 1200 mm (48 in.) concrete pipe diameter reinforced

with 9 kg/m3 (15 lb/yd3) PP fiber and 5.7 cm2/m (0.27 in2/ft) steel cage: (a) crown; (b)

invert; (c) springline.

(a) (b) (c)

Figure. 55. Crack development in 1200 mm (48 in.) concrete pipe diameter reinforced

with 9 kg/m3 (15 lb/yd3) PP fiber and 10.2 cm2/m (0.48 in2/ft) steel cage: (a) crown; (b)

invert; (c) springline.

(a) (b) (c)

Figure. 56. Crack development in 1500 mm (60 in.) concrete pipe diameter reinforced

with 9 kg/m3 (15 lb/yd3) PP fiber and 5.7 cm2/m (0.27 in2/ft) steel cage: (a) crown; (b)

invert; (c) springline.


168

(a) (b) (c)

Figure 57. Crack development in 1500 mm (60 in.) concrete pipe diameter reinforced

with 9 kg/m3 (15 lb/yd3) PP fiber and 8.9 cm2/m (0.42 in2/ft) steel cage: (a) crown; (b)

invert; (c) springline.

4.6.4 ASTM Strength Requirements

According to ASTM C76 (2015), concrete pipes are classified into five Classes,

numbered I through V, based on the strength required for each class, as listed in Table 4.

The cracking load (load that produces 0.3 mm [0.01 in] crack width) and ultimate load

measurements were compared with the requirements of each class in ASTM C76 (2015).

The applied load was divided by the diameter and length of test pipes, and the results are

summarized in Table 5. Based on these results, concrete pipe with a diameter of 1200

mm (48 in.) reinforced with a fiber content of 9 kg/m3(15 lb/yd3) and steel cage area of

5.7 cm2/m (0.27 in2/ft) did not meet the strength requirement of any of the ASTM classes.

As the steel area was increased to 10.2 cm2/m (0.48 in2/ft), the ultimate strength of the

tested pipe fulfilled the ultimate strength requirements of Classes I, II, and III. However,

the 0.3 mm (0.01 in.) cracking load did not comply with the cracking load specified with

any of these classes. The 1500 mm (60 in.) diameter pipe reinforced with a fiber content
169

of 9 kg/m3(15 lb/yd3) and steel area of 5.7 cm2/m (0.27 in2/ft) satisfied the ultimate

strength criterion of Class I. As the steel area increased to 10.2 cm2/m (0.48 in2/ft), the

tested pipe met the ultimate strength requirement for Classes I and II. However, the

cracking test results of neither 1500 mm (60 in.) diameter pipe met ASTM Class I or II

requirements. Due to the high flexibility of tested pipes as mentioned in the previous

section, more carrying capacity can be expected when the pipe is buried in the soil, due to

soil-structure interaction. The wall thickness of 1200 mm (48 in.) diameter fiber

reinforced concrete pipe represents a reduction of 50%, 60%, and 65%, compared to wall

thicknesses specified by ASTM C76 (2015) standard, Wall A, Wall B, and Wall C,

respectively. For 1500 mm (60 in.) pipe diameter, this corresponding wall thickness

reduction represents 50%, 58%, and 63%. This reduction in the concrete pipe thickness

can reduce material and transportation costs. Furthermore, incorporating synthetic

reduced the steel area by more than 50% compared to those specified by ASTM for Class

I, Class II, and III.

Table 31. ASTM C76 (2015) Class Criteria for Cracking Load and Ultimate Load

ASTM C76 class Cracking load, kN/m/m (kip/ft/ft) Ultimate load, kN/m/m (kip/ft/ft)

I 40 (0.8) 60 (1.2)

II 50 (1) 70 (1.5)

III 65 (1.35) 100 (2)

IV 100 (2) 150 (3)

V 140 (3) 175 (3.75)


170

Table 32. Summary of the Crack and Ultimate Loads of Tested Pipe

Pipe diameter, Steel area, cm2/m Fiber dosage, Cracking load, Ultimate load,
mm (in.) (in /ft)
2
kg/m (lb/yd )
3 3
kN/m/m (kip/ft/ft) kN/m/m (kip/ft/ft)

1200 (48) 5.7 (0.27) 9 (15) 22.3 (0.465) 49.3 (1.03)

1200 (48) 10.2 (0.48) 9 (15) 25.15 (0.525) 108 (2.25)

1500 (60) 5.7 (0.27) 9 (15) 22.35 (0.4667) 62.5 (1.30)

1500 (60) 8.9 (0.42) 9 (15) 30.25 (0.631) 75.32 (1.57)

4.6.5 Stiffness

Most standards consider pipe as a flexible member when it experiences a

deflection more than 2% of its inside diameter. All tested pipes experienced deflection of

5% of their inside diameter with greater stiffness than high-density polyethylene pipes as

specified by ASTM F2648 / F2648M (2013) and AASHTO M294 (2011). The stiffness

of the fiber reinforced concrete pipes was calculated by the applied load by the

corresponding deflection and the length of the pipe. The stiffness at 5% deflection for

each of the concrete pipes is summarized and compared to the corresponding ASMT

F2648 / F2648M (2013) specified stiffness in Table 6. The stiffness of 1200 mm (48 in.)

diameter pipe with 9 kg/m3(15 lb/yd3) PP fiber content and steel area of 5.7 cm2/m (0.27

in2/ft) was 4.8 (10) times that specified by ASTM standard for HDPE pipes. Note that as

the steel area in the concrete pipes increased from 5.7 to 10.2 cm2/m (0.27 to 0.48 in2/ft),

the stiffness at 5% deflection increased by 110%. For the 1500 mm (60 in.) diameter pipe

with 9 kg/m3(15 lb/yd3) PP fiber and steel area of 5.7 cm2/m (8.9 cm2/m), the stiffness at

5% deflection was 6.5(7.8) times that specified for HDPE pipe. As the steel area in the
171

1500 mm (60 in.) diameter pipes increased from 5.7 to 8.9 cm2/m (0.27 to 0.42 in2/ft), the

stiffness at 5% deflection increased by 19.5%.

The ratio of the deflection corresponding to the ultimate load to the 5% deflection

of the pipe diameter (ΔUltimate/Δ5%) experienced by concrete pipe under three-edge bearing

test is also listed in Table 6. All tested pipes experienced ΔUltimate/Δ5% ratio > 1, which

indicates that these pipes can each sustain a load greater than that which creates a 5%

deflection ratio, even without the benefit of soil-structure interaction when placed in the

earth. Incorporating synthetic fiber had a significant impact on the retained bending

strength of all tested pipes as all tested pipe experienced (ΔUltimate/Δ5%) ratio over 1.

Furthermore, the effect of fiber addition on the pipe stiffness was studied by

comparing the stiffness results of 1500 mm (60 in.) diameter pipe reinforced with fiber

content of 9 kg/m3 (15 lb/yd3) and steel area of 5.7 cm2/m (0.27 in2/ft) with past research

(Parke et al 2016). The stiffness results corresponding to a 3% deflection ratio were

employed for the comparison since no data are available from the literature regarding 5%

deflection ratio. The data in Table 7 show the fiber dosage of 9kg/m3 significantly

enhances the stiffness of the pipe. The stiffness of pipes reinforced with fiber increased

by 146.5%, 140.4%, and 95% compared to pipes reinforced with fiber dosage of 0, 3.56,

and 4.75 kg/m3, respectively. The results highlight the significance of using fiber content

of 9 kg/m3 (15 lb/yd3) as overall pipe performance-enhanced substantially. Therefore, the

inclusion of fiber along with one layer steel cage improved the stiffness of concrete pipes

significantly due to the fibers’ tendency to bridge cracks, and as a result, increased load

capacity.
172

Table 33. Stiffness of Tested Pipes at deflection ratio of 5%.

Steel area, Fiber stiffness@5% Specification of pipe Δutimate/Δ5%


Pipe stiffness- HDPE
cm2/m dosage, deflection, kN/m/m pipe @5%
diameter, deflection (ASTM
(in2/ft) kg/m3 (kip/ft/ft) F2648 / F2648M -
mm (in.) 13), kN/m/m
(lb/yd3) (kip/ft/ft)

1200 (48) 5.7 (0.27) 9 (15) 602 (12.6) 125 (2.6) 1.18

1200 (48) 10.2 (0.48) 9 (15) 1269 (26.5) 125 (2.6) 1.31

1500 (60) 5.7 (0.27) 9 (15) 691 (14.4) 105 (2.2) 1.04

1500 (60) 8.9 (0.42) 9 (15) 826 (17.3) 105 (2.2) 1.03

Table 34. Stiffness of Tested Pipes at Deflection Ratio of 3%.

Pipe Steel area, cm2/m Wall thickness, mm Fiber dosage, kg/m3 Stiffness @3%
diameter (in2/ft) (in.) (lb/yd3) deflection, kN/m/m
(kip/ft/ft)
1500 (60 in.)a 5.7 (0.27) 63 (2.5) 0 393 (8.2)
1500 (60 in.)a 5.7 (0.27) 63 (2.5) 3.56 (6) 403 (8.4)

1500 (60 in.)a 5.7 (0.27) 63 (2.5) 4.75 (8) 496 (10.3)

1500 5.7 (0.27) 63 (2.5) 9 (15) 969 (20.2)


a
past research results obtained by (Park et al, 2016).
173

4.7 Performance of Thin-wall Synthetic Fiber Reinforced Concrete Pipes Under Short

and Long-term Loading

4.7.1 Short-term Performance

The load-displacement curves of the 1200 and 1500 mm (48 and 60 in.) diameter

concrete pipes are shown in Figure 58 and Figure 59, respectively. For 1200 (48 in)mm

pipe reinforced with fiber dosage of 9 kg/m3(15 lb/yd3) and steel area of 5.7 cm2/m (0.27

in2/ft), a linear load-deflection response was observed prior to the first crack, which

occurred at a load of 10 kN, as depicted in Figure 58(a). At higher loads, the pipe

exhibited deflection hardening as the reinforcement resisted the propagation of cracks.

With the steel area was increased to 10.2 cm2/m(0.27 in2/ft), stiffer behavior was

observed in non-linear zone of the load-deflection response, as shown in Figure 58(b).

Increasing the steel area increased the cracking and ultimate load by 88% and 120%,

respectively. The 1200 mm (48 in.) diameter pipe reinforced with 9 kg/m3(15 lb/yd3) PP

fiber and steel area of 5.7 cm2/m (0.27 in2/ft) did not fulfill the strength requirements for

any class specified in ASTM C76. The 1200 mm (48 in.) diameter pipe reinforced with

steel area of 10.2 cm2/m (0.48 in2/ft) met the ultimate strength requirements for Classes I,

II, and III.

For 1500 mm (60 in.) diameter pipe reinforced with PP fiber content of 9

kg/m3(15 lb/yd3) and steel area of 5.7 cm2/m (0.27 in2/ft), the load-deflection response

was approximately linear until the first crack occurred at a load of 20 kN. As the load

increased, the pipe showed a non-linear response, with an ultimate load of 62 kN, as

depicted in Figure 59 (a). The cracking and ultimate loads increased by 36% and 21%,
174

respectively when the steel area was increased to 8.9 cm2/m (0.42 in2/ft). The first pipe

met the ultimate strength requirement of Class I, while the second pipe with steel

reinforcement area increased to 8.9 cm2/m (0.42 in2/ft)fulfilled the strength requirements

for Class I and II. The PP fibers bridged the cracks and did not pull out from the

concrete.

All tested pipes qualified as flexible, sustaining deflections of at least 2% of the

pipe diameter. Indeed, the pipes experienced deflection up to 5% of their inside diameter

with high load capacity and residual strength. This indicates that these pipes can be

expected to carry more loads when buried in the soil as a portion of the load can be

carried through the soil-pipe interaction.


175

120
Vertical deflection
100 Horizontal deflection

80

Load (kN)
60

40
5% deflection ratio
20

0
0 20 40 60 80 100
Deflection (mm)

(a)

120
Vertical deflection
100 Horizontal deflection

80
Load (kN)

60

40
5% deflection ratio
20

0
0 20 40 60 80 100
Deflection (mm)

(b)

Figure 58. Load deflection curves of 1200 mm diameter pipes reinforced with PP fiber

dosage of 9 kg/m3 and reinforcing steel area of: (a) 5.7 cm2/m(0.27 in2/ft); (b) 10.2

cm2/m(0.48 in2/ft). (Note: 1 kN = 0.2248 kip; 1 mm = 0.0394 in.)


176

120
Vertical Deflection
100 Horizontal deflection

80

Load (kN)
60

40
5% deflection ratio
20

0
0 20 40 60 80 100 120
Deflection (mm)

(a)

120
Verical deflection
Horizontal deflection
100

80
Load (kN)

60

40
5% deflection ratio
20

0
0 20 40 60 80 100 120
Deflection (mm)

(b)

Figure 59. Load deflection curves of 1500 mm diameter pipes reinforced with PP fiber

dosage of 9 kg/m3 and reinforcing steel area of: (a) 5.7 cm2/m (0.27 in2/ft); (b) 8.9 cm2/m

(0.42 in2/ft). (Note: 1 kN = 0.2248 kip; 1 mm = 0.0394 in.)


177

4.7.2 Long-term Performance for 30 days

Deflection Response

The load-displacement responses of three load stages of the long-term experiment

on the concrete pipe with diameter of 1200 mm (48 in.) reinforced with steel area of 5.7

cm2/m (0.27 in2/ft) and PP fiber dosage of 9 kg/m3(15 lb/yd3) are shown in Figure 60.

For Stage 1 (see Figure 60(a)), the behavior was linear for load up to 10 kN (2.25 kip);

then non-linear behavior was observed as the load increased. During Load Stage 1, there

are two unstable regions in the non-linear a portion of the graph, indicating the formation

of cracks that initiated at the inner crown, inner invert, and outer surface at the springline.

The initiation of each crack caused an abrupt increase in deflection accompanied by a

decrease in load. As the load was increased for Stage 2(see Figure 60(b)), linear behavior

was observed until new cracks were initiated at a load of 20.5 kN. During Stage 3 (see

Figure 60(c)), new cracks initiated at a load of 30 kN (6.74 kip) with better performance

in terms of the load deflection response as fewer cracks initiated than in the previous two

stages. During Load Stages 2 and 3, fewer cracks were initiated at the concrete pipe wall.

These cracks were hairline cracks connecting the main cracks formed during Load Stage

1 at the inner invert, inner crown, and springline. The hairline cracks did not influence

the load deflection responses significantly due to the fiber bridging effect. This

enhancement in the pipe deflection behavior after each load stage is because the fibers

resisting crack propagation are fully stressed, creating high tensile strength and enhancing

ductility. During Stages 1, 2, and 3, the pipe exhibited a deflection ratio of 1.75%, 2%,

and 3%, respectively. This confirms the flexible behavior observed in the short-term test.
178

Thus, in the field, a portion of the applied load will be carried through soil-structure

interaction.

(a)

(b)
179

(c)

Figure 60. Load vs. displacement plot of 1200 mm diameter concrete pipe reinforced

with steel area of 5.7 cm2/m (0.27 in2/ft) and PP fiber dosage of 9 kg/m3, (a) Stage 1

(40% ultimate load); (b) Stage 2 (50% ultimate load); (c) Stage 3 (70% ultimate load).

(Note: 1 kN = 0.2248 kip; 1 mm = 0.0394 in.).

The deflection responses as a function of time during Load Stages 1, 2, and 3 for

30 days each are shown in Figure 61, where the increase in the deflection is synchronized

with the change in the sustained load. The time-dependent deformation (subtracted from

the instantaneous deflection due to the application of the load) of the pipe during Stage 1

appeared to stabilize after 4 days as after which the deflection was observed to approach

an asymptote and become almost constant. The time-dependent vertical (horizontal)

deflection after 4 days was 4.3 mm (4.3 mm), and the total time-dependent deflection

after 30 days was 6.1 mm (5.6 mm). For Stage 2, linear time-dependent deflection was
180

observed during the first day and from 9.6 to 12.7 days after applying the load. Over

these ranges, almost constant time-dependent deflection response can be observed. The

time-dependent vertical (horizontal) deflections after 1, 9.6 12.7, and 30 days were 4.2

mm (4.0 mm), 5.3 mm (5.0 mm), 6.0 mm (5.9 mm), 6.8 mm (6.7 mm), respectively.

Under Stage 3, the deflection response over time was linear for 5 days after applying the

load, followed by a stable response. There is a slight jump in the deflection value on Day

17, which is attributed to a slight change in the load value. The time-dependent vertical

(horizontal) deflections after 5 and 30 days are 5.4 mm (4.7 mm) and 6.9 mm (6.0 mm),

respectively. After the deflection response stabilized, only a very slight increase in the

total deflection was noted. This increase in the time-dependent deflection could be

attributed to the pullout or stretch of the synthetic fibers under the sustained load. Also,

as the sustained load increased, the total time-dependent deflection was increased. This

behavior is ascribed mainly to the increase in the number of initiated cracks as the load

was increased, which resulted in fiber deformation contributing more to the total

deflection. These results concur with those from Mackay and Trottier (2004), who argued

synthetic fiber creep showed no significant impact on the deflection response and did not

adversely affect the fiber as a replacement for the steel reinforcing mesh.
181

(a)

(b)
182

(c)

Figure 61. Displacement vs. time plots of 1200 mm (48 in.) diameter concrete pipe

reinforced with steel area of 5.7 cm2/m (0.27 in2/ft) and PP fiber dosage of 9 kg/m3. (a)

Load Stage 1 (b) Load Stage 2 (c) Load Stage 3. (Note: 1 mm = 0.0394 in.).

Crack Width

Figure 62 shows the crack width versus time measured during Load Stage 2 and

Load Stage 3 at the inside invert, outer springline, and inside crown using LVDTs (see

Fig. 11) labeled as LVDT-1, LVDT-2, and LVDT-3, respectively. For Load Stage 1, a

feeler gauge was used to measure the crack width. The maximum recorded crack width

recorded initially was 0.25 mm during this stage and did not exceed the 0.3 mm (0.01 in.)

upper limit specified by ASTM Standard C76. However, after 30 days, the crack widths

at invert, springline, and crown increased to 0.56, 0.45, and 0.56 mm (0.022, 0.017, and

0.022 in.), respectively. Upon increasing the load to 50% of ultimate value for Load
183

Stage 2, the crack widths at the inside crown, inside invert, and outer springline increased

to 1.49, 0.78, and 1.83 mm (0.058, 0.03, and 0.072 in.), respectively. The crack widths at

the crown and invert increased almost linearly during the first day, and between 9.6 days

and 12.7 days, following the same pattern observed for the deflection response. The

time-dependent crack width (subtracted from the instantaneous crack width during the

initial response to the applied load in this stage) at the invert (crown) was 0.25 mm ( 0.01

in.)(0.13 mm [0.005 in.]), 0.32 mm ( 0.0126 in) (0.19 mm [0.0074 in.]), 0.35 mm (0.0137

in.) (0.22 mm [0.00866]), and 0.40 mm (0.0157 in.) (0.27 mm [0.0106]) mm after 1 day,

9.6 days, 12.7 days, and 30 days, respectively. The crack width time dependence at the

springline after the first day was almost linear before essentially stopping. The time-

dependent after crack width was 0.05 mm (0.001968 in.) at 1 day and 0.11(0.0043 in.)

mm at 30 days. During Load Stage 3, the total maximum crack width at the invert,

springline, and crown increased to 2.18, 1.01, and 2.53 mm (0.083, 0.04, and 0.099 in.),

respectively. The crack width response was linear crack over time within 5 days of

applying the load. At the springline, the crack width stabilized after a few hours applying

load with only 0.05 mm increase over the next 30 days. The time-dependent crack width

at the invert (crown) was 0.47 mm (0.018 in.) (0.37 mm [0.0145 in.]) and 0.54 mm (0.021

in.) (0.43 mm [0.017 in.]) after 5 days and 30 days, respectively. The time-dependent

crack width at the springline showed almost no change after 30 days. This increase in the

crack width can be attributed mainly to fiber creep or pull out, with other possible

contributing factors, including fiber aspect ratio, fiber volume, matrix composition, and

fiber geometry. Based on the physical properties of the PP fiber used in this study, high
184

bond strength is expected. However, when the stresses exerted on a fiber exceeds the

bond strength, the fiber will slip inside the matrix, and the crack width increase. As this

experiment represents a worst-case scenario, a study of a properly installed pipe buried in

soil will provide a more realistic evaluation of thin-walled fiber-reinforced concrete pipe,

including soil-structure interaction effects.


185

(a)

(b)

Figure 62. Crack width versus time plots of 1200 mm diameter concrete pipe reinforced

with steel area of 5.7 cm2/m and PP fiber dosage of 9 kg/m3(15 lb/yd3) during (a) Load

Stage 2 (b) Load Stage 3. (Note: 1 mm = 0.0394 in.)


186

LVDT-1

LVDT-3

LVDT-2

Figure. 63. LVDTs used to measure the crack width.

Strain Response

The strain readings were collected from six strain gauges installed on the surface

of the inner invert (labeled SG#1 in Figure 25 (b)), inner springline (SG#2), outer

springline (SG#3), inner midway crown-haunch (SG#4), outer midway crown-haunch

(SG#5), and inner crown (SG#6). The load versus strain measurements are depicted in

Figure 64. Due to flexural deformation, the strains measured at the inner crown, inner

invert, inner midway crown-haunch, and outer springline were positive (tensile), and

those measured on the inner surface of the springline and outer midway crown-haunch

were negative (compressive). This strain distribution was consistent with the oval-shaped

pipe deflection pattern reported in the previous section. All tested pipes exhibited linear

load-strain relationships prior to cracking. As cracks began to propagate, the load-strain

relationship became non-linear, as did the load-deflection response. During Load Stage
187

1, Strain gauges SG#1 and SG#6 at the inner crown and invert surfaces failed when their

maximum capacity was exceeded after cracks started underneath; the maximum strains

recorded were 12,236 με at the inner invert (SG#1) and 14,860 με at the inner crown

(SG#6). The other strain gauges, SG#2 through SG#5, continued recording as the cracks

propagated away from their locations. The maximum tensile strain within the elastic

range was 1000 με at the invert. The ultimate tensile (compressive) strains recorded at

outer springline and inner midway crown-haunch were 629 με (-1180 με) and 307 με (-

432) με, respectively. For Load Stage 2, the strain recorded at the outer springline and

inner midway crown-haunch increased to 825 με (-1320 με) and 376 με (-639 με),

respectively. During Load Stage 3 the corresponding ultimate tensile (compressive)

strains increased to 3108 με (-2887 με) and 516 με (-849 με), respectively. The fibers

bridging the cracks adequately transferred stresses generated in the concrete pipe wall

even as the number of cracks increased at each stage of loading. The ultimate tensile

strain value at the springline surpassed the compression strain value in Load Stage 3. This

behavior can be mainly attributed to that the neutral axis became closer to the

compressive face. On the other hand, ultimate tensile strain at the midway crown-haunch

was larger than compressive strain for all load stages; the stress distribution at the pipe

haunches are expected to have less stress intensity than the crown, invert, and springline.
188

(a)

(b)
189

(c)

Figure 64. Load-strain responses (a) invert and crown (b) springline (c) midway crown-

haunch.

During Load Stages 1 and 2, strains induced at the springline (see Figure 65(a))

increased linearly as the load was applied during the first day, then the strain leveled off

once the load was held constant at its final value, increasing slightly over the next 30

days. The tensile (compressive) strain change at the end of the first day of constant load

during Load Stage 1 was 79 με (-155 με) and at 30 days was 276 με (-451 με). Even

though the total strain increased in Load stage 2, the change in strain while the load was

held constant was less after 30 days than that recorded at the end of Load Stage 1. The

time-dependent tensile (compressive) strains in Load Stage 2 at the end of 1 day and 30

days were respectively 98 με (-77 με) and 142 με (-326 με). The lower 30-day strain

values could be attributed to stress redistribution associated with crack propagation.


190

During Load Stage 3, strains induced at the springline exhibited linear response during

the 5 days after the final load increment was applied. A sudden increase in strain was

observed on Day 17, perhaps in response to a slight increase in the load. The tensile

(compressive) strain change at the end of 5 days and 30 days were 1020 με (-661 με) and

1237 με (-902 με), respectively.

Figure 65 (b) shows the strain response in the midway crown-haunch as a

function of time. During Load Stage 1, the tensile (compressive) strain under constant

load after 1 day and 30 days were 14 με (-39 με) and 103 με (-168 με), respectively. As

the load was increased for Stages 2 and 3, the compressive strain-time dependent

increased, and the tensile strain time-dependent decreased. The tensile (compressive)

strains-time dependent at the end of Load Stages 2 and 3 were 9 με (-224 με) and 43 με (-

286 με), respectively. The tensile and compressive strain responses exhibited different

trends in terms of the increase, and the decrease in the strain values, which is attributed

primarily to the cracks, propagated the inner crown causing stress redistribution.
191

(a)

(b)

Figure 65. Strain responses as a function of time at (a) springline (b) midway crown-

haunch.
192

Crack Patterns

The growth and propagation of cracks were monitored to determine the effect of

the sustained load change on their number and the pattern. Flexural cracks were the only

type created under all loads. The flexural cracks propagated longitudinally where there

was high flexural tensile stress. Other crack types associated with larger diameter pipes,

such as shear and radial cracks, were not observed. During Load Stage 1, a single crack

first appeared on the inside surface at the crown, then one on the inside at the invert and

two on the outside at the springline, as depicted in Figure 66. Additional fine cracks were

also observed, especially at the invert and crown. The primary cracks propagated

longitudinally at the invert and crown across where a strain gauge was installed, causing

failure in the strain gauge. On the other hand, the primary cracks at the springline

occurred away from the centerline and avoided the strain gauge installed there, as shown

in Figure 66(b). Additional cracks appeared during Load Stage 2 at the crown and

springlines, as illustrated in Figure 67, and new cracks formed at the haunches. This

indicates, under certain conditions of pipe geometry and support, the haunches absorb the

excess bending moment from the springlines. The crack patterns resemble those reported

previously by Mont et al (2016), who also observed haunch cracked after crown and

springline. During Load Stage 3, newly formed fine cracks connected the existing cracks

from previous stages; these are presented in Figure 68.


193

3 (a) (b)

5 (c)

Figure 66. Cracks formed during Load Stage 1 at (a) crown (b) springline (c) invert.
194

Figure 67. Crack propagated during Load Stage 2 at (a) crown (b) springline
195

New cracks
New crack

2
3 (a) (b)

New crack

5
6 (c)

Figure 68. Cracks propagated during Load Stage 3 (a) crown (b) invert (c) springline.
196

4.7.3 Long-term Performance for 120 days

Deflection Response

The deflection response as a function of time under load stage-1 and 3 for 120

days is shown in Figure 69. It can be seen the increase in the incremental deflection is

synchronized with the change in the sustained load. The most important factor in this

research is the actual change in deflection after stabilization (Park et al, 2013). The time-

dependent deflection (subtracted from the instantaneous deflection due to applying load)

of the pipe under load stage-1 appeared to stabilize after 5 days as slight changes in the

time-dependent relationship was observed. The time-dependent vertical (horizontal)

deflection after 5 days was 4.3 mm (0.17 in.) (4.3 mm [0.017 in.]), and the total time-

dependent deflection after 120 days was 7.8 mm (0.3 in.) (6.9 mm [0.27]). Under load

stage-3, the rate of increase in the deflection was rapid within 5 days of applying the load,

but then it fell and exhibited a very slow rate of increase. The time-dependent vertical

(horizontal) deflections after 5 and 120 days were 5.4mm (0.21 in.) (4.7mm [0.185 in.])

and 7.5 mm (0.295 in.) (6.4mm [0.297 in.]), respectively. After the pipe deflection

response stabilized, a very slight increase in the total deflection was observed. This

increase in the time-dependent deflection could be attributed to the pullout or stretch of

the fibers under sustained load. Also, as the sustained load was increased, the total time-

dependent deflection was increased. This behavior can be mainly ascribed to the increase

in the number of initiated cracks as the load was increased, and as a result, more fiber

deformation contributing to the total deflection was expected. The results obtained from

the present study concur with previous research by Mackay and Trottier (2004), showing
197

that synthetic fiber can be used as a replacement for the steel reinforcement cage without

adverse effect due to fiber creep, as synthetic fiber showed no significant impact on the

deflection response. Also, the results herein in congruent with previous findings reported

by Park et al, (2013), showing that synthetic fiber the time-dependent material properties

of synthetic fiber concrete will not adversely affect the performance of the concrete pipes.

Furthermore, as shown in Figure 69, the total vertical deflections of the crown

and the horizontal deflection of springline during load stage 1 were 21.1 and 22.5 mm (

0.83 and 0.88 in.), respectively, while the corresponding deflection under load stage 2

was 34.3 and 30.7 mm (1.35 and 1.2 in.), respectively. This increase in the deflection

could be attributed to fiber pull-out or stretch of the fibers and the initiation of new cracks

(Park et al., 2014; Pujadas et al., 2017; Babafemi and Boshoff, 2013, 2015, 2016;

Bernard, 2010; MacKay and Trottier, 2004; Boshoff et al., 2009).On the other hand,

including synthetic fiber along with a steel reinforcement cage enhanced the ductility and

energy absorption capacity for concrete pipes. Also, the deflection ratio (deflection to the

pipe diameter) is an important index to classify pipes as flexible, semi-rigid pipes, or

rigid (Plastics Pipe Institute, 2011; AS4, 2003). Maintaining a deflection ratio of 2% or

more in concrete pipe ensures that it could perform as a non-rigid behavior when buried.

Thin-walled fiber reinforced concrete pipe exhibited a deflection ratio of 1.75 % and

2.5% under load stages 1 and 2, respectively. The results indicate that the pipe can endure

more load when buried in soil as the new concrete system has sufficient flexibility to

obtain passive support from the surrounding soil.


198

40
Vertical displacement
35 Horizontal displacement
Displacement (mm) 30
25
20
15
10
5
0
0 10 20 30 40 50 60 70 80 90 100 110 120 130
Time (days)

(a)

40
35
30
Displacement (mm)

25
20
15
10
5 Vertical displacement.
0 Horizontal displacement
0 10 20 30 40 50 60 70 80 90 100 110 120
Time (days)

(b)

Figure 69. Displacement vs. time plots of 1200 mm (48 in) diameter concrete pipe

reinforced with steel area of 5.7 cm2/m (0.27 in2/ft) and PP fiber dosage of 9

kg/m3(15lb/yd3). (a) Load Stage 1 (b) Load Stage 3. (Note:1 mm = 0.0394 in.).
199

Crack Width

Figure 70 shows the crack width versus time measured under load stage 1 and 3 at

the crown, invert, and outer face of the springline. For load stage 1, the crack width was

measured using Feeler gauge. On the other hand, during load stage 3, LVDTs labeled as

LVDT-1, LVDT-2, and LVDT-3 were used to monitor the crack width (see Figure 71),

respectively. The crack width response under all load stages coincides with the deflection

response due to the ovaling response of the pipe. The maximum recorded crack width

was 0.25 mm when the pipe was loaded under load stage 1. The crack width did not

exceed the limit specified by the ASTM C76 standard, 0.3 mm (0.01 in.). However, after

120 days, the crack widths at the invert, springline, and crown increased to 0.56, 0.53,

and 0.76 mm ( 0.022, 0.020, and 0.03 in.), respectively. As the load was increased to 70%

of the ultimate capacity during load stage 3, the crack widths at the crown, invert, and

outer face of the springline increased to 2.55, 2.22, and 1 mm (0.1, 0.087 and 0.039 in.),

respectively. The increase rate of crack width at the crown and invert was almost linear

within 5 days. The time-dependent crack width (subtracted from the instantaneous crack

width due to the applied load) at the invert (crown) was 0.29 mm (0.011 in.) (0.38 mm

[0.015]), 0.37 mm (0.0145 in.) (0.47 mm [0.0185]), and 0.45 mm (0.0177) (0.57 mm

[0.022]) mm after 1 day, 5 days, and 120 days, respectively. Also, the time-dependent

crack width at the springline after the first day was linear before essentially stopping. It

can be seen that the fiber has a slight impact on the crack width time-dependent increase

and this increase can be mainly attributed to the fiber creep or pull out. Moreover, the

increase in crack width could be also attributed to other factors associated with fiber-
200

matrix bond strength, including fiber aspect ratio, fiber volume, matrix composition, and

fiber geometry. Thus, based on the physical properties of the PP fiber used in this study,

high bond strength could be expected. However, when the stresses exerted on each fiber

exceeds the bond strength, a slip could occur between fibers and matrix and as a result, an

increase in the crack width is expected. Therefore, it is recommended that the synthetic

fiber-reinforced thin-wall concrete pipe be investigated further while buried in the soil as

the test setup employed in this study represents the most severe load condition.
201

0.7

0.6

0.5

Crack width (mm)


0.4

0.3

0.2

0.1 Crown
Invert
Springline
0
0 10 20 30 40 50 60 70 80 90 100 110 120 130
Time (days)

(a)

2.5
Displacement (mm)

1.5

1
LVDT-1
0.5
LVDT-2
LVDT-3
0
0 10 20 30 40 50 60 70 80 90 100 110 120
Time (days)

(b)

Figure 70. Crack width versus time plots of 1200 mm diameter concrete pipe reinforced

with steel area of 5.7 cm2/m (0.27 in2/ft) and PP fiber dosage of 9 kg/m3(15lb/yd3). (a)

Load Stage 1 (b) Load Stage 3. (Note:1 mm = 0.0394 in.).


202

LVDT-3

LVDT-2

LVDT-1

Figure 71. LVDTs used to measure the crack width.

Creep Coefficient

The creep coefficient was proposed by several authors to get a better

understanding of the creep behavior and feasibly compare the results of cracked fiber

reinforced concrete members under different load levels. The creep coefficient is the ratio

of creep deformation to the elastic deformation. Nevertheless, since the creep

deformation was not directly measured, the creep coefficient can be calculated using

crack width of deflection data as reported in the previous research (Babafemi and

Boshoff, 2013 ; Abrishambaf et al., 2015; Buratti, 2015; Buratti and Mazzotti, 2015;

Arango et al., 2011). Thus, the creep coefficient φ(t) was calculated as the ratio of crack

width due to the creep at a time (t) (𝑤𝑐𝑡 ) to the crack width measured immediately after

applying the load (𝑤𝑐𝑜 ) as presented in Eq. (1).


203

𝑤𝑐𝑡 −𝑤𝑐0 𝛥𝑤𝑐𝑡


𝜑(𝑡) = = (1)
𝑤𝑐0 𝑤𝑐0

The development of the creep coefficient after 120 days φ(t=120) under load

stages 1 and 3 is presented in

(a)

(b)

Figure 72. Following the trend indicated in previous sections for the time-

dependent deformation, the creep coefficient stabilized within several days in specimens

subjected to load stages 1 and 3. Also, as can be seen, the creep coefficient value ranged

between 0.05 to 2.2 for cracks propagated at the crown, invert, and springline under both

load stages. These results are significantly lower than those obtained by previous

researchers who reported that PP fiber reinforced concrete specimens experienced φ(t)

ranged from 1.5 to 3.5.

According to previous research by MacKay and Trottier (2004), the synthetic

fiber-reinforced specimens exhibit greater creep initially after loading, but the tensile
204

fiber creep becomes less significant as time elapses, and the remaining creep deformation

is due to the fiber pull out. However, the effect of fiber pull-out on the total creep

deformation can be minimized when increasing the bond strength of the fiber-matrix

interface, which is a function of fiber geometry and cement paste surrounding the fiber.

Thus, PP with embossed surface used in this study might have increased the bond

strength, and as a result, the total creep deformation was reduced. Also, this behavior

could be attributed to the fact that using a steel reinforcement cage along with synthetic

fiber aided to reduce the creep deformation as steel reinforcement is much lower than that

for synthetic fiber. The results herein indicate that synthetic fiber can be used along with

steel as reinforcement without concerns about the visco-elastic behavior of synthetic

fiber. This outcome may have a positive repercussion in the philosophy adopted to

account for the creep of FRC in the design.

(a)
205

(b)

Figure 72. Creep coefficient (a) Load stage 1 (b) load stage 2

Strain Response

The strain readings were collected from six strain gauges: SG#1, SG#2, SG#3,

SG#4, SG#5, and SG#6 installed on the surface of the inner invert, inner springline, outer

springline, inner midway crown-haunch, outer midway crown-haunch, and inner crown,

respectively. The instrumentation profile utilized in this study was very important to

study the stress distribution associated with strain distribution. Due to the flexural

deformation of tested pipes, the measured strain at the inner crown, inner invert, inner

midway crown-haunch, and outer of springline was positive (tensile strain). The strains

measured on the inner surface of the springline and outer midway crown-haunch were

negative (compressive strains). The strain distribution was consistent with the oval-

shaped behavior observed in response to pipe deflections reported in the previous section.
206

Under all load stages, the pipe exhibited linear load-strain relationships prior to the first

crack, followed by a non-linear response. During load stage-1, Strain gauges SG#1 and

SG#6 experienced high strains due to the cracks initiated at the inner crown and invert

surfaces. As a result, these strain gauges failed as their maximum capacity was reached.

The maximum recorded strains at the inner invert and crown were 12,236 με and 14,860

με, respectively. On the other hand, strain gauges, SG#2, SG#3, SG#4, and SG#5,

continued recording as the cracks propagated away from the strain gauges.

Figure 73. shows strain responses versus time from SG#2, SG#3, SG#4, and

SG#5. Under load stages 1 and 2, strains generated at the springline (see Figure 73 (a))

showed a linear response within 5 days of applying the load, followed by a stable

response. The tensile (compressive) time-dependent strains under load stage 1 after 5 and

120 days were 102 με (-327με) and 356 με (-653 με), respectively. As the load was

increased under load stage 2, the total and time-dependent strain increased. The time-

dependent tensile (compressive) strains under load stage 2 after 5 days and 120 days were

1019 με (-661 με) and 1596 με (-1062 με), respectively. This increase in the time-

dependent and total strains could be attributed primarily to stress redistribution associated

with the cracks propagation.

Figure 73(b) shows the strain response in the midway crown- hunch as a function

of time. The tensile (compressive) strain-time dependent under load stage 1 after 5 and 30

days were -33 με (-127 με) and 220 (-460 με), respectively. As the load was increased

under load case 2 and 3, it was observed that there was an increase in the tensile

compressive strain-time dependent. The tensile (compressive) strains-time dependent


207

under load case 2 were 37 με (-168 με) and 234 με (-346 με), respectively. Also, it was

noted that tensile and compressive strains responses exhibited different trends in terms of

the increase and the decrease in the strain values. This behavior is mainly attributed to the

cracks propagated at the inner crown, causing stress redistribution.

Furthermore, concrete pipes reinforced with synthetic fiber experienced an

increase in the strain as the load was increased beyond that for plain concrete. Also, the

ultimate tensile and compressive strain values at the springline are 3468 με and -3046 με,

respectively. The strain results indicate that concrete pipes reinforced with synthetic can

endure more strain beyond that for plain concrete. Also, it can be concluded that fiber

bridging the propagated cracks adequately transfer stresses generated in the concrete pipe

wall even though the number of cracks increased after each load stage. This behavior

highlights the fiber contribution to the pipe strength and ductility, indicating that the

fracture energy absorption improved significantly. As a result, thin-wall synthetic fiber

reinforced concrete pipes are expected to behave like a flexible pipe when buried in the

soil. This subject is the current focus of further research.


208

(a)

(b)

Figure 73. Strain responses as a function of time at (a) springline (b) midway crown-

haunch.
209

Crack Patterns

Crack growth and propagation were monitored during the test to determine the

effect of the sustained load change on the number and pattern of cracks propagation. It

was observed that the flexural cracks were the only type of cracks initiated under all load

stages. Flexural cracks propagated longitudinally at the vicinities that experienced high

flexural tensile stress. Under load stage 1, cracks first appeared on the inner face at the

crown and subsequently initiated on the inner face at the invert and the outer side of the

springline. Under load stage 1, These cracks were characterized with one main crack at

the crown and invert, and two main cracks on the outer surface at the springlines, as

depicted in Figure 74. Additional hairline cracks were also observed, especially, at the

invert and crown. The main cracks that propagated longitudinally at the invert and crown

passed through strain gauges, causing failure in the strain gauge. On the other hand, the

main cracks at the springline propagated away from the strain gauges, as shown in Figure

74(a). The number of cracks increased as the load was increased during load stage-2.

More cracks were observed at the crown and springlines, as illustrated inFigure 74. Also,

new cracks formed at the haunches, indicating that haunches at a certain grade of stress

redistribution absorb the excess of bending moment from the springlines. The

redistribution grade is a function of the geometry and support condition of the pipe. The

observed cracks’ patterns are similar to those reported in previous research conducted by

Mont et al., (2016), who reported that haunches experienced cracks after the cracks

occurred at the crown and springline. The presence of fibers bridging the cracks at the

crown, invert, and springline, allows section rotation while resisting the crack, resulting
210

in stress redistribution. Under load stage 3, it was also observed forming new fine cracks,

connecting cracks occurred in the previous load stage, as presented in Figure 75. During

all load stages, it was noted that the sustained load formed new cracks. The existing

cracks experienced increases in the width and depth, making these cracks largely visible

to the naked eye over time.

(a)

(b)
211

(c)

Figure 74. Cracks formed during Load Stage 1 at (a) crown (b) springline (c) invert.

(a)
212

(b)

(c)

Figure 75. Cracks propagated during Load Stage 3 (a) crown (b) invert (c) springline.
213

CHAPTER 5 : CONCLUSIONS, SUMMARY, AND RECOMMENDATION

5.1 Conclusions

The aim of this study was to evaluate the synthetic fiber reinforced concrete pipes

performance in terms of ASTM requirements for strength, stiffness, and ductility, and

develop design tables for synthetic fiber reinforced concrete pipe similar to those

proposed in ASTM C76 standard using the numerical analysis. The performance of the

synthetic fiber reinforced concrete pipes was evaluated under short- and long-term

loading in accordance with ASTM protocols using different pipe diameters. High fidelity

3-dimensional FEMs were developed to capture the complex behavior for the fiber-

reinforced concrete pipes. The FEMs were calibrated and validated with the field and

laboratory results. The validated FEM models were used to conduct A parametric study

on four parameters: pipe diameter, pipe wall thickness, fiber dosage, and steel cage area.

Based on this study, the following conclusions can be drawn:

5.1.1 Material Properties of Synthetic Fiber Reinforced Concrete under Freeze-

Material Properties

The effect of using synthetic polypropylene (PP) and polyvinyl alcohol (PVA)

reinforcing fibers on mechanical properties of concrete before and after applying freeze-

thaw cycles was presented in this study. The coefficient of thermal expansion (CTE) of

concrete reinforced with PVA and PP fibers increased over the value of plain concrete.

The dynamic modulus of elasticity of specimens reinforced with fiber dosage of 9 kg/m 3

PVA and PP fibers was the lowest compared to other specimens reinforced with low fiber

content. Adding PVA and PP fiber has a significant impact on the flexural strength of
214

concrete. The flexural strength of the concrete specimens increased substantially with

using fiber dosage of 9 kg/m3 (15 lb/yd3) for both types of fiber compared to the plain

concrete. Repeated freezing and thawing increased the CTEs of fiber reinforced

specimens, except for that with 6 kg/m3 PVA fiber. After 300 freeze-thaw cycles, plain

concrete specimens lost over 70% of their mass by the end of 300 freeze-thaw cycles.

The corresponding mass loss for fiber reinforced specimens was under 5%. For dosages

of 6 and 7 kg/m3 (10 and 12 lb/yd3), the concrete material loss was less than 1% for both

types of fiber. PVA fiber reinforced specimens showed a slight reduction in the dynamic

modulus of elasticity compared with PP fiber after 300 freeze-thaw cycles. The flexural

strength of fiber reinforced specimens decreased due to freeze-thaw cycles.

5.1.2 Evaluation of Synthetic Fiber Reinforced Concrete Pipe Performance Using

Three-Edge Bearing Test

The impact of incorporating synthetic fiber in concrete pipes on minimizing the

use of conventional steel reinforcing, improving the mechanical properties, and on the

overall performance was investigated. A total of 36 pipes of different diameters with

varying fiber content and presence or absence of a steel reinforcement cage were utilized

in conducting this study. Pipes were tested in accordance with ASTM C497. The

dominant failure mode in pipes with diameters of 600, 1200, and 1500 mm (24, 48, and

60 in.) was a flexural failure. Other failure modes associated with large pipes

diameter,1200 and 1500 mm (48 and 60 in.), was observed at high load rates. Inclusion of

synthetic fiber in the concrete pipes as a new reinforcement enhanced the load carrying

capacity, post-cracking behavior, ductility, and the stiffness of tested pipes. Each pipe
215

diameter has an optimum fiber dosage that meets strength requirements specified by

ASTM standards. Using synthetic fiber dosage of 18 kg/m3 (30 lb/yd3) reduced the steel

cage area as high as 100%, 78.76%, and 85.84% for pipes with the diameter of 600, 1200,

and 1500 mm (24, 48, and 60 in.), respectively, allowing for a significant reduction in

cost over conventional steel reinforced concrete pipes. Increasing fiber dosage improved

the stiffness of tested pipes. The size of this improvement was influenced by pipe size,

fiber dosage, and amount of steel reinforcement. For large size pipes, increasing fiber

dosage had more impact on stiffness enhancement than small size pipes. Also, using a

steel cage along with fiber as reinforcement showed less gain in the stiffness compared to

those reinforced with fiber only. The (ΔUltimate/Δ2%) ratio of 600 and 1200 mm (24 and 48

in.) pipe diameters reinforced with synthetic fiber along with steel cage was always larger

than 1, which indicates there is residual bending strength even after 2% deflection of the

inside diameter. Incorporating synthetic fiber in the concrete pipes with a large diameter

is effective in reducing some of the cracks that occur due to other failure modes

associated with a large diameter.

5.1.3 Design Proposal for Synthetic Fiber Reinforced Concrete Pipes Using Finite

Element Analysis

A three-dimensional finite element model was developed to simulate the three-

edge bearing test of synthetic fiber reinforced concrete pipes. Concrete damage plasticity

model was used to model polypropylene fiber reinforced concrete material. In order to

provide the finite element model with concrete material properties, compression strength,

tensile strength, modulus of elasticity, and Poisson’s ratio were measured. The finite
216

element model was calibrated and validated using the experimental results from full-scale

testing of PP fiber reinforced concrete pipes. A new compressive stress-strain model was

developed for PP fiber reinforced concrete material as the compressive strength of PP

fiber reinforced concrete reduced with increasing fiber dosage. The ultimate load results

of validated models were more conservative than the experimental results. The ultimate

and cracking loads of the finite element models were compared to the strength

requirements of each class specified by ASTM C76, and the results summarized in tables

similar to those presented in the ASTM C76 standard. A fiber dosage of 9 kg/m3 (15

lb/yd3) was sufficient to fulfill Class I strength requirement for Type A wall pipe with a

diameter of 300 to 750 mm (12 to 30 in.). For Type B wall pipe, using the same fiber

dosage fulfilled the strength requirements of Class I, except for pipe with diameter of

1200 mm (48 in.) where one layer of a steel cage was needed to achieve strength

requirements. Pipes with Type C walls fulfilled the strength requirements of Classes I and

II when fiber dosage of 9 kg/m3(15 lb/yd3) was used. As pipe diameter and certain class

strength increase, fiber dosage, steel cage area, or both were increased. The tables present

in the study can be used as a design reference for PP fiber reinforced concrete pipes.

5.1.4 Experimental Investigation of Thin-wall Synthetic Fiber Reinforced Concrete

Pipes

The effect of incorporating synthetic fiber with content of 9 kg/m 3(15 lb/yd3)

along with minimum steel reinforcement area on the performance of concrete pipe was

investigated. In an attempt to increase the concrete pipe flexibility and reduce the

production and transportation cost, the wall thickness of concrete pipe was reduced to
217

50% of that specified by ASTM C76 (2015) for Wall Type A. The pipes were subjected

to the three-edge bearing load test in accordance ASTM C496 (2015). The results were

evaluated in terms of load-deflection response, achieving ASTM strength requirements,

failure mode, stiffness, and strain capacity. The inclusion of PP fiber along with

minimum steel area and reduced wall thickness increased the flexibility of tested pipes as

all tested pipes experienced a deflection ratio of 5% of their inside diameter before

reaching their ultimate load. Fiber contribution to pipe strength and post cracking

behavior was more for pipes with diameter of 1500 mm (60 in.) than those with diameter

of 1200 mm (48 in.) due to wall thickness and failure modes associated with each pipe

configuration. The dominant failure mode was a flexural failure with cracks initiated

longitudinally at the inner invert, inner crown, and outer springline. Other failure modes

such as radial and shear failures were observed at high loading levels due to fiber

contribution. The synthetic fiber dosage of 9 kg/m3(15 lb/yd3) enhanced the overall pipe

performance compared to pipes studied in past research that used lower fiber dosages.

Fiber reinforced concrete pipe with a diameter of 1200 mm (48 in.) augmented with steel

reinforcement mesh at 10.2 cm2/m (0.48 in2/ft) fulfilled the ultimate load requirements of

ASTM Classes I, II and III. Pipe with diameter of 1500 mm (60 in.) met the ultimate load

requirements for Classes I and II. In addition to the reduction in steel reinforcement

relative to conventional concrete pipes, the wall thickness was reduced as much as 65%

and 63% for pipe with diameter of 1200mm (48 in.), and 1500 mm (60in.) compared to

conventional pipe specifications for Wall Type C, respectively. This significantly reduces

the cost of fiber reinforced concrete pipes compared to conventional steel reinforced
218

concrete pipes. The tested pipes exhibited a stiffness at 5% deflection ratio as much as

7.8 and10.2 times that of HDPE pipe of diameter of 1200mm (48 in.) and 1500 mm

(60in.), respectively. Using the fiber content of 9 kg/m3 (15 lb/yd3) has a significant

impact on the pipe stiffness compared to past research. The stiffness increased by 95%

when fiber dosage increased from 4.75 to 9 kg/m3 (8 to15 lb/yd3). The (ΔUltimate/Δ5%) ratio

of the fiber reinforced pipe was always larger than 1, which means there is residual

bending strength even when the pipe reaches a 5% deflection of their inside diameter.

Increasing the steel area along with 9 kg/m3(15 lb/yd3) synthetic fiber increased the strain

capacity of tested pipes as the concrete material was restrained by fiber and steel

reinforcement.

5.2 Performance of Thin-wall Synthetic Fiber Reinforced Concrete Pipes Under Short

and Long-term Loading

The performance of concrete pipes reinforced with synthetic fiber dosage of 9

kg/m3 (15 lb/yd3) and minimum steel reinforcement area was investigated. To reduce the

cost of production and transportation and increase the concrete pipe flexibility, these

concrete pipes had a wall thickness reduced to 50% of that specified by ASTM C76

(2015) for Wall A. The pipes were subjected to short and long term three-edge bearing

load tests. The results were evaluated in terms of deflection, strain, crack width, and

crack patterns. Incorporating synthetic fiber with dosage of 9 kg/m3 (15 lb/yd3) along

with minimum steel reinforcement areas enhanced the flexibility of the thin-walled

concrete pipes as the pipes sustained deflection ratios of 5% before reaching ultimate

load capacity. Under constant load, deflection, strain, and crack width readings increased
219

linearly within a specific period of time, and then leveled off. Only a slight increase in the

time-dependent deflection was observed after 30 days during all load stages, indicating

the fiber creep has no significant impact on the long-term deflection. The time-dependent

crack width increased slightly even when the load was increased to 70% of the ultimate

load capacity. Synthetic fiber reinforced concrete pipe with steel reinforcement can

endure greater strains than those reported by past research for the conventional concrete

material without fiber reinforcement. As the increased load was applied at different

stages, new cracks were observed at invert, crown, and springline. Also, during Load

Stages 2 and 3 cracks were observed in the pipe haunches, indicating the haunches

absorbed the excess bending moment from the crown, invert, and springlines.

5.3 Recommendation

The behavior of joints between segments of synthetic fiber reinforced concrete

pipes joint should be investigated since joints are typically the weakest points in a

concrete pipe system. Also, it should be noted that pipes are often more likely to be

damaged during installation than in service. Therefore, additional studies using dynamic

and concentrated load conditions are necessary to fully understand the behavior of the

synthetic fiber reinforced concrete pipes.


220

REFERENCES

AASHTO (American Association of State Highway and Transportation Officials).

(1989). “AASHTO guide specifications thermal effect in concrete bridge

superstructures.” Washington, D.C.

AASHTO, AASHTO LRFD bridge design specifications, SI Units, 3rd Ed., Washington,

DC, 2012.

AASHTO. (2011). “Standard specification for corrugated polyethylene pipe, 300- to

1500-mm (12- to 60-in.) Diameter.” M 294-11, Washington, DC.

Abaqus 6.12-3 [Computer software]. Dassault Systèmes, Providence, RI.

Abolmaali, A. and Kararam, A. (2013). “Nonlinear Finite-Element Modeling Analysis of

Soil-Pipe Interaction.” International Journal of Geomechanics, 46(June), 197–

204.

Abrishambaf, A., Barros, J.A. and Cunha, V.M., (2015) Time-dependent flexural

behaviour of cracked steel fibre reinforced self-compacting concrete

panels. Cement and Concrete Research, 72: 21-36.

ACI (American Concrete Institute). (2010). “Report on the Physical properties and

Durability of Fiber-Reinforced Concrete.” ACI 544.5R-10, ACI Committee 544,

Farmington Hills, Michigan.

Ahn, T., Thermal and mechanical studies of Thin Spray-on Liner (TSL) for concrete

tunnel linings (Doctoral dissertation, The University of Western Ontario), 2011.


221

Akentuna, M., Kim, S. S., Nazzal, M., and Abbas, A. R. (2017). “Asphalt Mixture CTE

Measurement and the Determination of Factors Affecting CTE.” Journal of

Materials in Civil Engineering, 04017010

Al Rikabi, F. T., Sargand, S. M., and Hussein, H. H. (2020) Design Proposal for

Synthetic Fiber-Reinforced Concrete Pipes Using Finite Element

Analysis. Journal of Testing and Evaluation, 48(2),

https://doi.org/10.1520/JTE20180097.

Al Rikabi, F. T., Sargand, S. M., Hussein, H. H., and Khoury, I. “Material Properties of

Synthetic Fiber Reinforced Concrete under Freeze-Thaw Conditions.” Journal of

Materials in Civil Engineering,10.1061/(ASCE)MT.1943-5533.0002297.

Al Rikabi, F. T., Sargand, S. M., Kurdziel, J.M, and Hussein, H. H., “Experimental

Investigation of Thin-wall Synthetic Fiber Reinforced Concrete Pipes”, ACI

Structural journal, 2018. Vol. 115, 6, pp.1671-1681.

https://doi:10.14359/51702413.

Al Rikabi, F. T., Sargand, S.M., Kurdziel, J.M., “Evaluation of Synthetic Fiber

Reinforced Concrete Pipe Performance Using Three-Edge Bearing Test,” J. of

Testing and Evaluation. (In press).

Al Rikabi, F., Sargand, S., and Kurdziel, J., "Evaluation of Synthetic Fiber Reinforced

Concrete Pipe Performance Using Three-Edge Bearing Test," Journal of Testing and

Evaluation, Vol. 7, Issue 2, 2019, https://doi.org/10.1520/JTE20170369.

Al Rikabi, F.T., Sargand, S.M. and Kurdziel, J., (2018a) Evaluation of Synthetic Fiber

Reinforced Concrete Pipe Performance Using Three-Edge Bearing Test. Journal of

Testing and Evaluation, 47(2): 942-958.


222

Al Rikabi, F.T., Sargand, S.M., Kurdziel, J. and Hussein, H.H. (2018b) Experimental

Investigation of Thin-Wall Synthetic Fiber-Reinforced Concrete Pipes. ACI

Structural Journal, 115(6): 1671-1681.

Allan, M. L. and Kukacka, L. E. (1995). “Strength and durability of polypropylene fiber

reinforced grouts.” Cement and concrete research, 25(3), 511-521.

Al-Ostaz, A. (2007). “Study Effect of moisture content on the coefficient of thermal

expansion of concrete.” Technical Report, Department of Civil Engineering,

University of Mississippi

Altoubat, S., Yazdanbakhsh, A. and Rieder, K.A. (2009) Shear behavior of macro-

synthetic fiber-reinforced concrete beams without stirrups. ACI Materials

Journal, 106(4), 381-389

American Association of State Highway and Transportation (AASHTO). (2007).

“Standard test method for the coefficient of thermal expansion of hydraulic

cement concrete.” Standard Specifications for Transportation Materials and

Methods of Sampling and Testing, TP60-00, Washington, DC.

American Association of State Highway and Transportation Officials (AASHTO).

(1989). AASHTO guide specifications thermal effect in concrete bridge

superstructures, Washington, D.C.

American Concrete Institute (ACI). (2008). “Prediction of creep, shrinkage and

temperature effects in concrete structures.” ACI-209R-92, ACI Committee 209,

Farmington Hills, Michigan.


223

Arango, S.E., Serna, P., Martí-Vargas, J.R. and García-Taengua, E. (2012) A test method

to characterize flexural creep behavior of pre-cracked FRC

specimens. Experimental mechanics, 52(8), 1067-1078.

AS4139-2003, (2003) Fiber-Reinforced Concrete Pipe and Fitting, Standards Australia

International, Sydney.

ASCE. (2001). “Standard practice for direct design of precast concrete pipe using

standard installations (SIDD).” ASCE 15-93, Reston, VA.

ASTM C1818-15, Standard Specification for Synthetic Fiber Reinforced Concrete

Culvert, Storm Drain, and Sewer Pipe, ASTM International, West Conshohocken,

PA, 2015, www.astm.org

ASTM C39/C39M-15, Standard test method for compressive strength of cylindrical

concrete specimens, ASTM International, West Conshohocken, PA, 2015,

www.astm.org.

ASTM C469/C469M-14, Standard Test Method for Static Modulus of Elasticity and

Poisson’s Ratio of Concrete in Compression, ASTM International, West

Conshohocken, PA, 2014, www.astm.org.

ASTM C496 / C496M-15, Standard Test Method for Splitting Tensile Strength of

Cylindrical Concrete Specimens, ASTM International, West Conshohocken, PA,

2017, www.astm.org

ASTM C497-15a, (2015b) Standard Test Methods for Concrete Pipe, Concrete Box

Sections, Manhole Sections, or Tile, ASTM International, West Conshohocken,

PA, www.astm.org.
224

ASTM C76-15, Standard Specification for Reinforced Concrete Culvert, Storm Drain,

and Sewer Pipe,” West Conshohocken, PA, 2015, www.astm.org.

ASTM, Standard test methods for concrete pipe, manhole sections, or tile.” C497-15,

(2015b), West Conshohocken, PA, 2015, DOI: 10.1520/C0412-15.

www.astm.org.

ASTM. (2012). “Standard Test Method for Linear Shrinkage and Coefficient of Thermal

Expansion of Chemical-Resistant Mortars, Grouts, Monolithic Surfacings, and

Polymer Concretes.” C531-00(2012), West Conshohocken, PA.

ASTM. (2013). Standard Specification for 2 to 60 inch [50 to 1500 mm] Annular

Corrugated Profile Wall Polyethylene (PE) Pipe and Fittings for Land Drainage

Applications, F2648/F2648M-13, West Conshohocken, PA.

ASTM. (2014) “Standard test methods for concrete pipe, manhole sections, or tile.”

C497-14, West Conshohocken, PA.

ASTM. (2014). “Standard Test Method for Fundamental Transverse, Longitudinal,

and Torsional Resonant Frequencies of Concrete Specimens.” ASTM C215-14

(2014), West Conshohocken, PA.

ASTM. (2015) “Standard Specification for Reinforced Concrete Culvert, Storm Drain,

and Sewer Pipe.” C76M-15, West Conshohocken, PA.

ASTM. (2015a). “Standard Test Method for Resistance of Concrete to a Rapid Freezing

and Thawing.” C666 / C666M-15, West Conshohocken, PA.


225

ASTM. (2015b). “Standard Test Method for Flexural Performance of Fiber-Reinforced

Concrete (Using Beam with Third-Point Loading).” C1609 / C1609M-15, West

Conshohocken, PA.

ASTM. (2015c). “Standard test method for flexural strength of concrete (using simple

beam with third-point loading).” C78/C78M-15, West Conshohocken, PA.

Babafemi, A. J., and Boshoff, W. P., “Tensile creep of macro-synthetic fibre reinforced

concrete (MSFRC) under uni-axial tensile loading,” Cement and Concrete

Composites,” Vol. 55, (2015), pp. 62-69.

https://doi.org/10.1016/j.cemconcomp.2014.08.002.

Babafemi, A. J., and Boshoff, W. P., “Testing and modelling the creep of cracked macro-

synthetic fibre reinforced concrete (MSFRC) under flexural loading,” Materials

and Structures, 49(10), (2016), pp. 4389-4400. https://doi.org/10.1617/s11527-

016-0795-7.

Babafemi, A.J. and Boshoff, W.P. (2013) Time-dependent behaviour of pre-cracked

polypropylene fibre reinforced concrete (PFRC) under sustained loading. Res.

Appl. Struct. Eng. Mech. Comput:1593-1598.

Babafemi, A.J. and Boshoff, W.P. (2016) Testing and modelling the creep of cracked

macro-synthetic fibre reinforced concrete (MSFRC) under flexural

loading. Materials and Structures, 49(10), 4389-4400.

Babafemi, A.J. and Boshoff, W.P., (2015) Tensile creep of macro-synthetic fibre

reinforced concrete (MSFRC) under uni-axial tensile loading. Cement and

Concrete Composites, 55: 62-69.


226

Bagherzadeh, R. Sadeghi, A. H. and Latifi, M. (2011). “Utilizing polypropylene fibers to

improve physical and mechanical properties of concrete.” Textile Research

Journal, 0040517511420767.

Bagherzadeh, R., Pakravan, H. R., Sadeghi, A. H., Latifi, M., and Merati, A. A. (2012).

“An Investigation on Adding Polypropylene Fibers to Reinforce Lightweight

Cement Composites (LWC).” Journal of Engineered Fabrics and Fibers

(JEFF), 7(4).

Bagherzadeh, Roohollah, Abdol-Hossein Sadeghi, and Masoud Latifi. "Utilizing

polypropylene fibers to improve physical and mechanical properties of

concrete." Textile Research Journal, Vol. 82, No.1, 2012, pp. 88-96.

https://doi.org/10.1177/0040517511420767.

Balaguru P, Shah S. (1992) Fiber-reinforced cement composites. New York: McGraw-

Hill Inc.

Balaguru, P. N. and Shah, S. P. (1992). Fiber-reinforced cement composites. McGraw-

Hill, Incorporated, New York.

Balaguru, P. N. and Shah, S. P. Fiber-reinforced cement composites. McGraw-Hill,

Banthia, N. and Sappakittipakorn, M. (2007). “Toughness enhancement in steel fiber

reinforced concrete through fiber hybridization.” Cement and Concrete

Research, 37(9), 1366-1372.

Banthia, N., and Dubey, A. (2000). “Measurement of flexural toughness of fiber-

reinforced concrete using a novel technique part 2: performance of various

composites.” Materials Journal, 97(1), 3-11.


227

Barros, J. A., and Figueiras, J. A., “Flexural behavior of SFRC: testing and

modeling,” Journal of Materials in Civil Engineering, Vol. 11. No. 4, pp. 331-

339, https://doi.org/10.1061/(ASCE)0899-1561(1999)11:4(331).

Bernard, E. S. (2010) Influence of Fiber Type on Creep Deformation of Cracked Fiber-

Reinforced Shotcrete Panels. ACI Materials Journal, 107(5).

Blazejowski, M., Flexural Behaviour of Steel Fibre Reinforced Concrete Tunnel Linings,

Doctoral dissertation, The University of Western Ontario, 2012.

Boshoff, W. P., Mechtcherine, V., and van Zijl, G. P. (2009) Characterising the time-

dependant behaviour on the single fibre level of SHCC: Part 1: Mechanism of

fibre pull-out creep. Cement and Concrete Research, 39(9), 779-786.

British Standards (BS) Institution. (1986). “Testing Concrete-Recommendations for

Measurement of Velocity of Ultrasonic Pulses in Concrete.” BS 1881-203.

Buratti, N. and Mazzotti, C., 2015. Experimental tests on the effect of temperature on the

long-term behaviour of macrosynthetic Fibre Reinforced Concretes. Construction

and Building Materials, 95, pp.133-142.

Buratti, N., and Mazzotti, C. (2015). Experimental tests on the effect of temperature on

the long-term behaviour of macrosynthetic Fibre Reinforced

Concretes. Construction and Building Materials, 95, 133-142.

Canada, 2011.

Çavdar, A. (2013). “The effects of high temperature on mechanical properties of

cementitious composites reinforced with polymeric fibers.” Composites Part B:

Engineering, 45(1), 78-88.


228

Çavdar, A. (2014). “Investigation of freeze–thaw effects on mechanical properties of

fiber reinforced cement mortars.” Composites Part B: Engineering, 58, 463-472.

Chen, L. and Graybeal, B. A. (2012) Modeling Structural Performance of Ultrahigh

Performance Concrete I-Girders, Journal Bridge Eng., 17(5): 754–764,

Choi, Y., and Yuan, R. L., “Experimental relationship between splitting tensile strength

and compressive strength of GFRC and PFRC,” Cem. And Concr. Re., Vol. 35,

No. 8, 2005, pp.1587-1591, https://doi.org/10.1016/j.cemconres.2004.09.010.

Choubane, B. and Tia, M. A. N. G. (1992). “Nonlinear temperature gradient effect on

maximum warping stresses in rigid pavements.” Transportation Research

Record, 1370(1), 11.

concrete I-girders.” Journal of Bridge Engineering, Vol. 17. No. 5, 2011, pp. 754-

764. https://doi.org/10.1061/(ASCE)BE.1943-5592.0000305

de la Fuente, A., Escariz, R. C., de Figueiredo, A. D., and Aguado, A. (2013) Design of

macro-synthetic fibre reinforced concrete pipes, Construction and Building

Material., 43: 523-532.

Ezeldin, A. S., and Balaguru, P. N., “Normal-and high-strength fiber-reinforced concrete

under compression,” J. of materials in civil engineering, Vol. 4. No. 4, 1992, pp.

415-429, https://doi.org/10.1061/(ASCE)0899-1561(1992)4:4(415).

Gergely, P. and Lutz, L. A. (1968). “Maximum crack width in reinforced concrete

flexural members.” ACI special publication, 87-117.

Gruebl, P., (1980). “Rapid ice formation in hardened cement paste, mortar and concrete

due to supercooling.” Cement and Concrete Research, 10, 334-345.


229

Haktanir T, Ari K, Altun F, Karahan O. A comparative experimental investigation of

concrete, reinforced-concrete and steel-fibre concrete pipes under three-edge-

bearing test. Constr Build Mater 2007;21(8):1702–8

Huang, W. H. (1997). “Properties of cement-fly ash grout admixed with bentonite, silica

fume, or organic fiber.” Cement and Concrete Research, 27(3), 395-406.

Huang, W. H. (2001). “Improving the properties of cement–fly ash grout using fiber and

superplasticizer.” Cement and Concrete Research, 31(7), 1033-1041.

Karahan, O. and Atiş, C. D. (2011). “The durability properties of polypropylene fiber

reinforced fly ash concrete.” Materials and Design, 32(2), 1044-1049.

Kim, D. H., Fasulo, P. D., Rodgers, W. R., and Paul, D. R. (2008). Effect of the ratio of

maleated polypropylene to organoclay on the structure and properties of TPO-

based nanocomposites. Part II: Thermal expansion behavior. Polymer, 49(10),

2492-2506.

Kim, S. S. Nazzal, M. Abbas, A. R. Akentuna, M. and Arefin, M. S. (2015). Evaluation

of Low Temperature Cracking Resistance of WMA (No. FHWA/OH-2015/11).

Korhonen, C., (2002). “Effect of high doses of chemical admixtures on the freeze-thaw

durability of portland cement concrete.” US Army Corps of Engineers: Engineer

and Research Development Center, Technical Report 02-5.

Kuder, K. G. and Shah, S. P. (2010). “Processing of high-performance fiber-reinforced

cement-based composites.” Construction and Building Materials,24(2), 181-186.

Kuder, K. G., and Shah, S. P. (2010) Processing of high-performance fiber-reinforced

cement-based composites. Construction and Building Materials, 24(2): 181-186.


230

Li, V. C., and Maalej, M. (1996). “Toughening in cement based composites. Part II: Fiber

reinforced cementitious composites.” Cement and Concrete Composites, 18(4),

239-249.

MacKay, J., and Trottier, J. F., “Post-crack creep behavior of steel and synthetic FRC

under flexural loading.” Shotcrete: Proceedings of the Second International

Conference on Engineering Developments in Shotcrete, October 2004, Cairns,

Queensland, Australia More engineering developments, pp. 183-192.

https://doi.org/10.1201/9780203023389.

Mansur, M. A., Chin, M. S., and Wee, T. H., “Stress-strain relationship of high-strength

fiber concrete in compression.” J. Mater. Civ. Eng., Vol.11. No. 1, 1999, pp. 21–

29, https://doi.org/10.1061/(ASCE)0899-1561(1999)11:1(21).

Mark, J. E. (Ed.). (2009). Polymer data handbook. Oxford university press.

Mehta PK and Monteiro PJM. (2006). Concrete: microstructure, properties, and

materials, 3rd ed. New York.

Mirza, S. A., and MacGregor, J. G., “Strength and Ductility of Concrete Slabs Reinforced

with Welded Wire Fabric,” In Journal Proceedings, Vol. 78, No. 5, 1981, pp.

374-381, 002-8061/81/050374-08.

Mobasher, B., and Li, C. Y. (1996). “Mechanical properties of hybrid cement-based

composites.” ACI Materials Journal, 93, 284-292.

Mobasher, B., Mechanics of fiber and textile reinforced cement composites, CRC press,

2011.
231

Mohamed, Nedal, and Moncef L. Nehdi. "Rational finite element assisted design of

precast steel fiber reinforced concrete pipes." Engineering Structures, Vol. 124,

2016, pp. 196-206

Monte, R., de la Fuente, A., de Figueiredo, A. D., and Aguado, A.,. (2016). “Barcelona

Test as an Alternative Method to Control and Design Fiber-Reinforced Concrete

Pipes.” ACI Structural Journal, Vol.113, No. 60.

https://doi.org/10.14359/51689018.

Munro, S. M. Moore, I. D. and Brachman, R. W. (2009).” Laboratory testing to examine

deformations and moments in fiber-reinforced cement pipe.” Journal of

geotechnical and geoenvironmental engineering, 135(11), 1722-1731.

Nataraja, M. C., Dhang, N., and Gupta, A. P., “Stress–strain curves for steel-fiber

reinforced concrete under compression,” Cem. and Concr. Comp., Vol. 21, No.5,

1999, pp. 383-390.

Neville, A. M. (1995). Properties of concrete. Wiley, New York

Neville, A. M. (2011). Properties of Concrete, 5th. London: Pitman Publishing, 687, 331.

Neville, A. M. and Brooks, J. J. (1993). Concrete technology. England: Longman

Scientific and Technical, Longman Group UK Limited.

Ng, T. S., Htut, T. N. S., and Foster, S. J., “Fracture of steel fibre reinforced concrete—

the Unified Variable Engagement Model,” UNICIV Report R-460, School of Civil

and Environmental Engineering, The University of New South Wales, Sydney,

2012.
232

Park, Y., Abolmaali, A., Attiogbe, E., and Lee, S. H. (2014). “Time-Dependent Behavior

of Synthetic Fiber-Reinforced Concrete Pipes Under Long-Term Sustained

Loading.” Journal of the Transportation Research Board, (2407), 71-79.

Park, Y., Abolmaali, A., Beakley, J. and Attiogbe, E. (2015) Thin-walled flexible

concrete pipes with synthetic fibers and reduced traditional steel

cage. Engineering Structures, 100: 731-741.

Peyvandi, A., Soroushian, P., and Jahangirnejad, S, (2013) Enhancement of the structural

efficiency and performance of concrete pipes through fiber

reinforcement, Construction and Building Materials, 45: 36-44.

Peyvandi, A., Soroushian, P., and Jahangirnejad, S. (2014). “Structural design

methodologies for concrete pipes with steel and synthetic fiber

reinforcement.” ACI Structural Journal, 111(1), 83.

Plastics Pipe Institute, (2011), “Design Methodology,” Corrugated Polyethylene Pipe

Design Manual and Installation Guide, Plastics Pipe Institute,

https://plasticpipe.org/drainage/cppa-design-guide.html (accessed 17 Jan. 2017).).

Puertas, F., Amat, T., Fernández-Jiménez, A., and Vázquez, T., “Mechanical and durable

behavior of alkaline cement mortars reinforced with polypropylene

fibres,” Cement and Concrete Research, Vol. 33, No. 12, 2003, pp. 2031-2036,

https://doi.org/10.1016/S0008-8846(03)00222-9.

Pujadas, P., Blanco, A., Cavalaro, S., de la Fuente, A., and Aguado, A. (2017) The need

to consider flexural post-cracking creep behavior of macro-synthetic fiber

reinforced concrete, Construction and Building Materials, 149: 790-800.


233

Raghavan, D. Huynh, H. and Ferraris, C. F. (1998). “Workability, mechanical properties,

and chemical stability of a recycled tyre rubber-filled cementitious

composite.” Journal of Materials science, 33(7), 1745-1752.

Richardson, A. E. (2003). “Freeze/thaw durability in concrete with fiber

additions.” Structural Survey, 21(5), 225-233.

Richardson, A. E. Coventry, K. A. and Wilkinson, S. (2012). “Freeze/thaw durability of

concrete with synthetic fiber additions.” Cold regions science and technology, 83,

49-56.

Salih, S. A. and Al-Azaawee, M. E. (2008). “Effect of polypropylene fibers on properties

of mortar containing crushed bricks as aggregate.” Eng Technol,26(12), 1508-

1513.

Sellevold, E. J. and Bjøntegaard. (2006). “Coefficient of thermal expansion of cement

paste and concrete: Mechanisms of moisture interaction.” Materials and

Structures, 39(9), 809-815.

Setzer, M. J. and Liebrecht, A., (2002). “Frost dilatation and pore system of hardened

cement paste under different storage conditions.” Frost Resistance of Concrete,

RILEM Proceedings, 24, 169-178

Simpson, A. L., Schmalzer, P. N., and Rada, G. R. (2007). “Long Term Pavement

Performance Project Laboratory Materials Testing and Handling Guide” Report

No. FHWA-HRT-07-052, Federal Highway Administration, Washington, DC.


234

Soroushian, P., and Lee, C. D., “Constitutive modeling of steel fiber reinforced concrete

under direct tension and compression,” Fibre reinforced cements and concretes,

recent developments, R. N. Swamy and B. Barr, eds., 1989, pp. 363–375.

Sullivan, James F. (1988) Technical Physics. USA, Wiley Publishers.

Tripathi, D. (2002). Practical guide to polypropylene. Smithers Rapra Technology.

U.S. Army Corps of Engineers (CRD) (1981), “Test Method for Coefficient of Linear

Thermal Expansion of Concrete.” CRD-C 39-81.

United States Department of Agriculture (USDA), Natural Resources Conservation

Service. (1976). “Air entrainment and concrete.” publication#C760105.

(http://www.nrcs.usda.gov/Internet/FSE_DOCUMENTS/nrcs141p2_023438.pdf)

(1976)

Voo, J. Y. L. and Foster, S. J. (2003) Variable Engagement Model for Fibre Reinforced

Concrete in Tension,” UNICIV Report No. R-420, University of New South

Wales, Sydney, Australia.

Voo, J. Y. L. and Foster, S. J. (2004) Tensile-Fracture of Fibre-Reinforced Concrete:

Variable Engagement Model,” presented at the Sixth International RILEM

Symposium on Fibre Reinforced Concretes (FRC) 535 – BEFIB 2004, Varenna,

Italy, Sep. 20 536 –22, pp.875–884.

Wang, Y., Backer, S., and Li, V. C., “An experimental study of synthetic fibre reinforced

cementitious composites,” Journal of Materials Science, Vol. 22, No. 12, 1987,

pp. 4281-4291, https://doi.org/10.1007/BF01132019.


235

Wilson, A. and Abolmaali, A. (2014). “Performance of Synthetic Fiber-Reinforced

Concrete Pipes.” Journal of Pipeline Systems Engineering and Practice, 5(3),

04014002.

Wilson, A., Abolmaali*, A., Park, Y., and Attiogbe, E., (2017) “Research and Concepts

behind the ASTM C1818 Specification for Synthetic Fiber Concrete Pipes”,

ASTM Special Topic (STP1601). doi:10.1520/STP160120160125.

Yun, H. D. and Rokugo, K. (2012). “Freeze-thaw influence on the flexural properties of

ductile fiber-reinforced cementitious composites (DFRCCs) for durable

infrastructures.” Cold Regions Science and Technology, 78, 82-88.

Zhang, M., Mirza, J., and Malhotra, V., "Mechanical Properties and Freezing and

Thawing Durability of Polypropylene Fiber-Reinforced Shotcrete Incorporating

Silica Fume and High Volumes of Fly Ash," Cement, Concrete and Aggregates,

Vol. 21, No. 2, 1999, pp. 117-125, https://doi.org/10.1520/CCA10425J.

Zhang, P. and Li, Q. F. (2013). “Effect of polypropylene fiber on durability of concrete

composite containing fly ash and silica fume.” Composites Part B:

Engineering, 45(1), 1587-1594.

Zhong, W. (2007). “Effect of polypropylene fibers on the long-term tensile strength of

concrete.” Journal of Wuhan University of Technology-Mater. Sci. Ed. 22(1), 52-

55.

Zollo, R. F.,1997, “Fiber-reinforced concrete: an overview after 30 years of

development,” Cement and Concrete Composites, 19(2), 107-122.


236

APPENDIX I: LIST OF JOURNAL PAPERS

Al Rikabi, F. T., Sargand, S.M., Kurdziel, J.M. (2017). Evaluation of Synthetic

Fiber Reinforced Concrete Pipe Performance Using Three-Edge Bearing Test. Journal of

Testing and Evaluation, 47(2). https://doi.org/10.1520/JTE20170369. ISSN 0090-3973

Al Rikabi, F. T., Sargand, S. M., Khoury, I., and Hussein, H. H. (2018). Material

Properties of Synthetic Fiber–Reinforced Concrete under Freeze-Thaw

Conditions. Journal of Materials in Civil Engineering, 30(6), 04018090.

Al Rikabi, F. T., Sargand, S. M., Kurdziel. J., and Hussein, H. H. (2017).

Experimental Investigation of Thin-Wall Synthetic Fiber-Reinforced Concrete Pipes. ACI

Structural Journal 115, no. 6 (2018): 1671-1681

Al Rikabi, F. T., Sargand, S. M., and Hussein, H. H. (2018). Design Proposal for

Synthetic Fiber-Reinforced Concrete Pipes Using Finite Element Analysis. Journal of

Testing and Evaluation, 48(2). https://doi.org/10.1520/JTE20180097. ISSN 0090-3973.

Al Rikabi, F. T., Sargand, S. M., and Kurdziel, J.M. Experimental study of Thin-

Walled Synthetic Fiber Reinforced Concrete Pipes Under short and long-term loading.

ASTM- Journal of Testing and Evaluation. https://doi.org/10.1520/JTE20180369.

Al Rikabi, F. T., Sargand, S. M., Khoury, I., and Kurdziel, J.M. (2019). A New

Test Method for Evaluating the Long-Term Performance of Fiber Reinforced Concrete

Pipes. Journal of Advances in Structural Engineering. (In press).

Hussein, H. H., Sargand, S. M., Al Rikabi, F. T., and Steinberg, E. P. (2016).

Laboratory Evaluation of Ultrahigh-Performance Concrete Shear Key for Prestressed


237

Adjacent Precast Concrete Box Girder Bridges. Journal of Bridge Engineering, 22(2),

04016113.

Hussein, H. H., Walsh, K. K., Sargand, S. M., Al Rikabi, F. T., and Steinberg, E.

P. (2017). Modeling the shear connection in adjacent box-beam bridges with ultrahigh-

performance concrete joints. I: Model calibration and validation”. Journal of Bridge

Engineering, 22(8), 04017043.

Sargand, S. M., Walsh, K. K., Hussein, H. H., Al Rikabi, F. T., and Steinberg, E.

P. (2017). Modeling the shear connection in adjacent box-beam bridges with ultrahigh-

performance concrete joints. II: Load transfer mechanism. Journal of Bridge

Engineering, 22(8), 04017044.

Hussein, H. H., Sargand, S. M., Al Rikabi, F. T., and Steinberg, E. P. (2018).

Experimental validation of optimized ultra-high-performance concrete shear key shape

for precast pre-stressed adjacent box girder bridges. Construction and Building Materials,

190, 178-190.

Mutashar, R., Sargand, S., Khoury, I., and T. Al Rikabi, F. (2018). Influence of

Nonuniform Box Beam Dimensions and Bridge Transverse Slope on Environmentally

Induced Stresses in Adjacent Box Beam Bridges. Journal of Performance of Constructed

Facilities, 32(6), 04018081.

Mutashar, R., Sargand, S., Al Rikabi, F. T., and Khoury, I. (2019). Response of a

Composite-Adjacent Box Beam Bridge with Skewed Beams under Static and Quasi-

Static Loads. Journal of Performance of Constructed Facilities, 33(3), 04019022.


238

Mutashar, R., Sargand, S., Khoury, I., and T. Al Rikabi, F. (2020).’’

Experimental and Theoretical Investigation Live Load Distribution Factors for Skewed

Composite Adjacent Box Beam Bridge”. Journal of Bridge Engineering. (In press)

Mutashar, R., Sargand, S., Khoury, I., and T. Al Rikabi, F. (2020). “Early-Age

Crack Prediction of Concrete Decks in The Skewed Adjacent Box Beam Bridges”.

Structure and Infrastructure Engineering. (Submitted).


239

APPENDIX II: LIST OF CONFERENCE PAPERS

Al Rikabi, F. T., Sargand, S. M., Hussein, H. H., and Khoury, I. (2017). The

Thermal Expansion of Synthetic Fiber-Reinforced Concrete under Air-Dry and Saturated

Conditions. First Congress on Technical Advancement, American Society of Civil

Engineers, Duluth, MN, 10-21. https://doi.org/10.1061/9780784481035.002

Al Rikabi, F. T., Sargand, S. M., and Kurdziel, J.(2018)‘ Thin-Walled Steel Fiber

Reinforced Concrete Pipes Performance under Three-Edge Bearing Load. Proc.,Annual

UESI Pipelines Conf., American Society of Civil Engineers, Toronto, Ontario, Canada.

https://doi.org/10.1061/9780784481646.037, 355 – 363.

Hussein, H. H., Sargand, S. M., Al Rikabi, F. T., and Steinberg, E. P. (2017).

Evaluation of Ultra-High Performance Concrete Grout Performance under Longitudinal

Shear. First Congress on Technical Advancement, American Society of Civil Engineers,

Duluth, MN, 34-44. https://doi.org/10.1061/9780784481035.004

Al Rikabi, F. T., Hussein, H. H., and Khoury, I. (2019). Experimental Study on

Shear Strength of Synthetic Fiber Reinforced High Strength Concrete Containing Slag

Aggregate. Structures Congress, American Society of Civil Engineers, Orlando, FL.,

April 24-27. https://doi.org/10.1061/9780784482223.018.

Al Rikabi, F. T., Sargand, S. M., and Kurdziel, J.(2019)‘ Thin-Walled Synthetic

Fiber Reinforced Concrete Pipe Performance Under Cyclic Loading.Annual UESI

Pipelines Conf., American Society of Civil Engineers, Nashville, TN, July 21-24.

Hussein, H. H., Al Rikabi, F. T., and Khoury, I. (2019). Effects of Interface

roughness on Shear Key Performance of Ultra-High Performance Concrete in Adjacent


240

Box Girder Bridges. 2nd International Interactive Symposium on UHPC, Albany , NY,

June 2- 5.

Khoury, I., Sargand, S. M., Hussein, H. H., and Al Rikabi, F. T. (2020). Field

investigation of Metal Multi-Pipe Culvert under Shallow Cover . Performance Under

Cyclic Loading. Annual UESI Pipelines Conf., American Society of Civil Engineers, San

Antonio, Texas, August 9-12.

Hussein, H. H., Al Rikabi, F. T., and Khoury, I. (2021). Experimental

investigation on shrinkage behavior of Synthetic Fiber Reinforced High Strength

Concrete Containing Slag Aggregate. Structures Congress, American Society of Civil

Engineers, Seattle, Washington, March 10-13. (submitted).

Al Rikabi, F. T., Hussein, H. H., and Khoury, I. (2021). Flexural Performance of

fiber Reinforced Concrete Containing Slag. Structures Congress, American Society of

Civil Engineers, Seattle, Washington, March 10-13. (submitted)


!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!
!

!
!
Thesis and Dissertation Services

You might also like