Sei sulla pagina 1di 355

Experimental and Analytical Studies of Geo-Composite

Applications in Soil Reinforcement

Item type text; Electronic Dissertation

Authors Toufigh, Vahab

Publisher The University of Arizona.

Rights Copyright is held by the author. Digital access to this


material is made possible by the University Libraries,
University of Arizona. Further transmission, reproduction
or presentation (such as public display or performance) of
protected items is prohibited except with permission of the
author.

Downloaded 30-Nov-2017 17:57:01

Link to item http://hdl.handle.net/10150/255167


EXPERIMENTAL AND ANALYTICAL STUDIES OF GEO-COMPOSITE

APPLICATIONS IN SOIL REINFORCEMENT

by

Vahab Toufigh

A Dissertation Submitted to the Faculty of the


DEPARTMENT OF CIVIL ENGINEERING AND ENGINEERING MECHANICS

In Partial Fullfillment of the Requirements


For the Degree of

DOCTOR OF PHILOSOPHY
WITH A MAJOR IN CIVIL ENGINEERING

In the Graduate College

THE UNIVERSITY OF ARIZONA

2012
2

THE UNIVERSITY OF ARIZONA


GRADUATE COLLEGE

As members of the Dissertation Committee, we certify that we have read the dissertation

prepared by Vahab Toufigh

entitled Experimental and Analytical Studies of Geo-Composite Applications in Soil


Reinforcement

and recommend that it be accepted as fulfilling the dissertation requirement for the

Degree of Doctor of Philosophy

_______________________________________________________________________ Date:
Hamid Saadatmanesh

_______________________________________________________________________ Date:
Tribikram Kundu

_______________________________________________________________________ Date:
Lianyang Zhang

_______________________________________________________________________ Date:
John M. Kemeny

_______________________________________________________________________ Date:

Final approval and acceptance of this dissertation is contingent upon the candidates
submission of the final copies of the dissertation to the Graduate College.

I hereby certify that I have read this dissertation prepared under my direction and
recommend that it be accepted as fulfilling the dissertation requirement.

________________________________________________ Date:
Dissertation Director: Hamid Saadatmanesh
3

STATEMENT BY AUTHOR

This dissertation has been submitted in partial fulfillment of requirements for an


advanced degree at the University of Arizona and is deposited in the University Library
to be made available to borrowers under rules of the Library.

Brief quotations from this dissertation are allowable without special permission, provided
that accurate acknowledgment of source is made. Requests for permission for extended
quotation from or reproduction of this manuscript in whole or in part may be granted by
the head of the major department or the Dean of the Graduate College when in his or her
judgment the proposed use of the material is in the interests of scholarship. In all other
instances, however, permission must be obtained from the author.

SIGNED: Vahab Toufigh


4

ACKNOWLEDGEMENTS

I would like to thank all those who stood behind me and supported me throughout my life.

First of all, I would like to thank God. I would like to thank my parents, Saleheh Jafari

and Mohammad M. Toufigh, for all their love and support through my entire life. This

dissertation would not have been possible without Professor Desai who not only served

as my co-advisor but also encouraged and challenged me throughout my academic

program. I would like to thank Professor Desai for the use of the constitutive modeling

laboratory, helping me with the disturbed state concept (DSC) model and helping me

with my PhD dissertation and journal manuscripts. I would like to thank Professor

Budhu for letting me use Plaxis software. I also would like to thank my friends Ahad

Ouria, Ehsan Kabiri, Ed A. Farnaghi, Saeed Ahmari and Wafa Alfatesh for their help. I

would like to thank my advisor, Professor Saadatmanesh for all his support. I would like

to thank all my friends and my family for their support.


5

DEDICATION

To my parents

for their love, endless support

and encouragement.
6

TABLE OF CONTENTS

LIST OF FIGURES .......................................................................................................... 12

LIST OF TABLES ............................................................................................................ 28

ABSTRACT...................................................................................................................... 30

CHAPTER 1. INTRODUCTION ..................................................................................... 32

CHAPTER 2. LITERATURE REVIEW .......................................................................... 36

2.1. Soils Type ............................................................................................................. 36

2.1.1. Saline Soil ....................................................................................................... 36

2.1.2. Soft Soil .......................................................................................................... 38

2.1.3. Collapsing Soils .............................................................................................. 38

2.1.4. Modification of Soil Weakness....................................................................... 40

2.2. Soil Stabilization .................................................................................................... 41

2.2.1. Reinforcement Geometry ................................................................................ 44

2.2.2. Reinforcement Material .................................................................................. 44

2.2.3. Reinforcement Extensibility ........................................................................... 45

2.2.4. Facing Systems ............................................................................................... 45

2.2.5. Reinforcement Types ...................................................................................... 49

2.2.6. Application Area ............................................................................................. 51

2.2.7. Theory ............................................................................................................. 60


7

TABLE OF CONTENTS-Continued

2.2.8. Failure Modes of Soil ..................................................................................... 65

2.3. Factors Effecting Soil Reinforcements .................................................................. 68

2.3.1. Reinforcement ................................................................................................. 69

2.3.2. Reinforcement Distribution ............................................................................ 76

2.3.3. Soil .................................................................................................................. 77

2.3.4. Soil State ......................................................................................................... 78

2.4. Durability ............................................................................................................... 83

2.4.1. Introduction ..................................................................................................... 83

2.4.2. Corrosion ........................................................................................................ 84

2.4.3. Degradation ..................................................................................................... 85

2.5. Fiber reinforcement polymer (FRP)....................................................................... 88

2.5.1. Introduction ..................................................................................................... 88

2.5.2. Fiber ................................................................................................................ 90

2.5.3. Glass Fiber ...................................................................................................... 92

2.5.4. Carbon Fiber ................................................................................................... 92

2.5.5. Matrix Materials ............................................................................................. 93

2.5.6. Advantages...................................................................................................... 94

2.6. Interface Behavior ................................................................................................ 100

2.6.1. Introduction ................................................................................................... 100

2.6.2. Pullout Test ................................................................................................... 104


8

TABLE OF CONTENTS-Continued

2.6.3. Direct shear test ............................................................................................ 107

2.7. Different Models .................................................................................................. 108

CHAPTER 3. LABORATORY RESULTS.................................................................... 112

3.1. Direct Shear Test.................................................................................................. 113

3.1.1. Direct shear machine .................................................................................... 113

3.1.2. Calibration direct shear test .......................................................................... 114

3.1.3. Soil ................................................................................................................ 117

3.1.4. Fiber Reinforcement Polymer ....................................................................... 119

3.1.5. Preparation .................................................................................................... 121

3.1.6. Effects of Initial normal force ....................................................................... 123

3.1.7. Friction angle between soil and CFRP ......................................................... 132

3.1.8. Precast preparation and results ..................................................................... 136

3.1.9. Discussion ..................................................................................................... 140

3.2. Pull out test .......................................................................................................... 140

3.2.1. Test materials: ............................................................................................... 141

3.2.1.1. Sand........................................................................................................ 141

3.2.1.2. Epoxy ..................................................................................................... 141

3.2.1.3. CFRP ...................................................................................................... 141

3.2.2. Test equipment .............................................................................................. 141

3.2.3. Test procedure............................................................................................... 143


9

TABLE OF CONTENTS-Continued

3.2.3.1. Precast tests ............................................................................................ 143

3.2.3.2. Cast-in-place tests .................................................................................. 144

3.2.4. Test results and discussion ............................................................................ 144

3.2.4.1. Precast mode .......................................................................................... 144

3.2.4.2. Cast-in-place mode ................................................................................ 148

3.2.5. Discussion ..................................................................................................... 149

3.3. CYMDOF Test..................................................................................................... 152

3.3.1. CYMDOF machine ....................................................................................... 152

3.3.2. Soil ................................................................................................................ 155

3.3.3. Samples Preparation ..................................................................................... 159

3.3.4. Installing CFRP for Direct Shear Test Using CYMDOF Machine .............. 163

3.3.5. Test Result .................................................................................................... 174

3.3.6. Discussion ..................................................................................................... 206

CHAPTER 4. CONSTITUTIVE MODELS FOR SOIL AND INTERFACE................ 208

4.1. Introduction .......................................................................................................... 208

4.2. Different Types of Constitutive Models .............................................................. 210

4.2.1. Elasticity Theory ........................................................................................... 210

4.2.2. Plasticity Theory ........................................................................................... 211

4.2.3. Hierarchical Single Surface (HISS) Model .................................................. 216


10

TABLE OF CONTENTS-Continued

CHAPTER 5. DETERMINATION OF MATERIALS PARAMETERS AND MODEL

VERIFICATION ............................................................................................................ 223

5.1. Introduction .......................................................................................................... 223

5.2. Soil ....................................................................................................................... 224

5.3. Interface ............................................................................................................... 226

5.3.1. Elastic Parameters for HISS Model .............................................................. 229

5.3.2. Elastoplastic Parameters for HISS Model .................................................... 232

5.3.2.1. A Comparison between q=1.54 and q=2 ............................................... 251

5.3.3. Computer Programming for Back Prediction ............................................... 253

CHAPTER 6. NUMERICAL MODELING OF MSE WALL USING PLAXIS ........... 271

6.1. Introduction .......................................................................................................... 271

6.2. Mechanical Stabilized Earth Walls (MSE) .......................................................... 272

6.3. Numerical Modeling ............................................................................................ 273

6.4. MSE Wall in Tucson Arizona .............................................................................. 281

6.5. Comparing the Plaxis (HS) Results with Field Results ....................................... 283

6.6. Design guideline for MSE wall............................................................................ 290

6.7. Plaxis (HS) analyses on MSE wall using CFRP .................................................. 297

6.8. Analyzing number of reinforcing layers and vertical spacing ............................. 302

6.8.1. Number of reinforcing layers ........................................................................ 302

6.8.2. Vertical spacing at the center of the wall (CW)............................................ 303


11

TABLE OF CONTENTS-Continued

6.8.3. Vertical spacing at the centroid of the lateral forces (CF) ............................ 309

6.8.4. Finial analysis ............................................................................................... 314

6.9. Construction sequences and cost estimation ........................................................ 324

CHAPTER 7. CONCLUSIONS ..................................................................................... 335

REFERENCES ............................................................................................................... 340


12

LIST OF FIGURES

Figure 1-1 Flow chart for the research program. .............................................................. 35

Figure 2-1 Bridge abutment (Abu-Hejleh et al., 2002). .................................................... 52

Figure 2-2 Construction of the Valence dAlbi Dam (Girard 1990). ............................... 53

Figure 2-3 Reinforced embankment (Bergado and Teerawattanasuk 2008). ................... 54

Figure 2-4 Geosynthetic membrane for foundation (Chen et al. 2007). ........................... 55

Figure 2-5 Land fill (Palmeira 2002). ............................................................................... 56

Figure 2-6 Reinforced pack for roof supporting underground mining (Mining, 2010). ... 57

Figure 2-7 Pipe structure (Chen, Y et al 2008). ................................................................ 58

Figure 2-8 Railway support for embankment( Liu et al., 2008). ...................................... 59

Figure 2-9 Sea wall (Yasuhara, 2007). ............................................................................. 60

Figure 2-10 Coyne-LadderWall. ....................................................................................... 60

Figure 2-11 Howe beam.................................................................................................... 61

Figure 2-12 Howe beam analogy of ladder wall. .............................................................. 62

Figure 2-13 Stress state in soil. ......................................................................................... 63

Figure 2-14 Soil & reinforcement. .................................................................................... 64

Figure 2-15 Failure modes for reinforced soil slopes (Allen and Bathurst, 2001). .......... 65

Figure 2-16 Reinforced soil wall. ..................................................................................... 66

Figure 2-17 External failure modes for reinforced soil slopes (Allen and Bathurst, 2001).

................................................................................................................................... 68

Figure 2-18 Load-displacement result for shear test (Noorzad & Mirmoradi, 2010). ...... 69
13

LIST OF FIGURES-Continued

Figure 2-19 Results of photo-elastic studies in the direct shear test. (a) Unreinforced; (b)

Vertical reinforcement; (c) Inclined reinforcement(Dyer, 1985). ............................ 70

Figure 2-20 Common forms of reinforcement. ................................................................. 71

Figure 2-21 Rough and smooth surface (Palmeria, 2009). ............................................... 72

Figure 2-22 Different size of reinforcement (Khedkar and Mandal, 2009). ..................... 73

Figure 2-23 Pullout load versus displacement curve comparison for various reinforcement

(Khedkar and Mandal, 2009). ................................................................................... 73

Figure 2-24 Pressure vs. strain data for various geotextiles (Andrejack & Wartman, 2010).

................................................................................................................................... 74

Figure 2-25 Rupture stress-time-temperature relationship of PE geomembrane under

axisymmetric loading (Duvall, 1993). ...................................................................... 75

Figure 2-26 Effect of number of geogrids, n, on the load required for a settlement of 0.5

mm (Phanikumar, Prasad, and Singh, 2009)............................................................. 79

Figure 2-27 Influence of overburden stress on index friction angle (Narejo, 2003). ....... 81

Figure 2-28 Active earth pressure distribution vs. height of wall (Ahmadabadi, and

Ghanbari, 2009). ....................................................................................................... 82

Figure 2-29 Coefficient of permeability vs. normal soil (Raisinghani and Viswanadham,

2010). ........................................................................................................................ 83

Figure 2-30 Relationship among OIT; temperature and pressure (Hsuan and Li, 2005) . 86

Figure 2-31 Chemical reaction.......................................................................................... 88


14

LIST OF FIGURES-Continued

Figure 2-32 FRP to the various aggressive environments (Tavkolizadeh and

Saadatmanesh, 2005) ................................................................................................ 99

Figure 2-33 Modes of deformations at interface (Desai et al., 1984). ............................ 103

Figure 2-34 Interaction mechanisms of reinforced soil wall (modified from Palmeira and

Milligan, 1989a). ..................................................................................................... 105

Figure 2-35 Boundary of pullout tests. ........................................................................... 106

Figure 3-1 Direct shear machine. .................................................................................... 114

Figure 3-2 Using steel balls on direct shear boxes. ........................................................ 115

Figure 3-3 Results on direct shear test on old-shear box (soil-soil). .............................. 116

Figure 3-4 Results on direct shear test on new shear box (soil-soil). ............................. 116

Figure 3-5 Friction angle for two different shear boxes. ................................................ 117

Figure 3-6 Particle size analysis. .................................................................................... 118

Figure 3-7 Compaction curve for sand soil..................................................................... 118

Figure 3-8 Shear Stress vs. Horizontal Displacement. ................................................... 119

Figure 3-9 One layer of carbon fiber. ............................................................................. 120

Figure 3-10 Failure of CFRP. ......................................................................................... 121

Figure 3-11 Saturated carbon fiber. ................................................................................ 122

Figure 3-12 Block and CFRP.......................................................................................... 123

Figure 3-13 Schematic diagram of modified shear box, used in experiments. ............... 123

Figure 3-14 Direct shear test with 0 m compaction depth. ............................................. 124
15

LIST OF FIGURES-Continued

Figure 3-15 Variation of maximum shear stress vs. normal stress for 0 m compaction

depth. ....................................................................................................................... 125

Figure 3-16 Direct shear test with 0.1 m compaction depth. .......................................... 125

Figure 3-17 Variation of maximum shear stress vs. normal stress for 0.1 m compaction

depth. ....................................................................................................................... 126

Figure 3-18 Direct shear test with 0.2 m compaction depth. .......................................... 126

Figure 3-19 Variation of maximum shear stress vs. normal stress for 0.2 m compaction

depth. ....................................................................................................................... 127

Figure 3-20 Direct shear test with 0.3 m compaction depth. .......................................... 127

Figure 3-21 Variation of maximum shear stress vs. normal stress for 0.3 m compaction

depth. ....................................................................................................................... 128

Figure 3-22 Direct shear test with 0.4 m compaction depth. .......................................... 128

Figure 3-23 Variation of maximum shear stress vs. normal stress for 0.4 m compaction

depth. ....................................................................................................................... 129

Figure 3-24 Direct shear test with 0.5 m compaction depth. .......................................... 129

Figure 3-25 Variation of maximum shear stress vs. normal stress for 0.5 m compaction

depth. ....................................................................................................................... 130

Figure 3-26 Direct shear test with 1 m compaction depth. ............................................. 130

Figure 3-27 Variation of maximum shear stress vs. normal stress for 1 m compaction

depth. ....................................................................................................................... 131


16

LIST OF FIGURES-Continued

Figure 3-28 Friction angle vs. compaction depth. .......................................................... 132

Figure 3-29 Direct shear test between CFRP and Soil. .................................................. 133

Figure 3-30 Variation of maximum shear stress vs. normal stress for CFRP and sand. 133

Figure 3-31 CFRP w/ soil & CFRP w/o soil with normal stress of 49 kPa. ................... 134

Figure 3-32 CFRP w/ soil & CFRP w/o soil with normal stress of 98 kPa .................... 135

Figure 3-33 CFRP w/ soil & CFRP w/o soil with normal stress of 196 kPa. ................. 135

Figure 3-34 Soil & cast in place CFRP w/ 100 cm soil in top i=16.1 kPa.................... 137

Figure 3-35 Variation of maximum shear stress vs. normal stress on Sand & cast in place

CFRP w/ 100 cm soil in top i=16.1 kPa. ............................................................... 137

Figure 3-36 Precast CFRP & cast in placed CFRP with normal stress of 49 kPa. ......... 138

Figure 3-37 Precast CFRP & cast in placed CFRP with normal stress of 98 kPa. ......... 138

Figure 3-38 Precast CFRP & cast in placed CFRP with normal stress of 196 kPa. ....... 139

Figure 3-39 Comparing ratio the of max shear strength to normal strength b/w cast in

place & precast........................................................................................................ 139

Figure 3-40 Pull out box. ................................................................................................ 142

Figure 3-41 Test apparatus for pullout test. .................................................................. 142

Figure 3-42 Twenty-four CFRP-Sand specimens prepared for precast pull-out test...... 144

Figure 3-43 Pull out test on Precast CFRP-Sand with normal stress 25kPa................... 145

Figure 3-44 Pull out test on Precast CFRP-Sand with normal stress 50kPa................... 146

Figure 3-45 Pull out test on Precast CFRP-Sand with normal stress 100kPa................. 147
17

LIST OF FIGURES-Continued

Figure 3-46 Friction angle of CFRP-Sand with soil. ...................................................... 148

Figure 3-47 Cast-in-place CFRP-Sand specimens pull-out test results. ........................ 149

Figure 3-48 Comparison results of pull-out test for cast-in-place and pre-casted for height

of 25 cm. ................................................................................................................. 150

Figure 3-49 Comparison results of pull-out test for cast-in-place and pre-casted for height

of 50cm. .................................................................................................................. 150

Figure 3-50 Comparison results of pull-out test for cast-in-place and pre-casted for height

of 100cm. ................................................................................................................ 151

Figure 3-51 Cyclic multi degree of freedom machine. ................................................... 154

Figure 3-52 Hydraulic pump........................................................................................... 154

Figure 3-53 Tanque Verde Wash. ................................................................................... 155

Figure 3-54 Tanque Verdes soil. ................................................................................... 156

Figure 3-55 Particle size distribution curve for the soil.................................................. 157

Figure 3-56 Standard compaction test. ........................................................................... 158

Figure 3-57 Compaction curve. ...................................................................................... 158

Figure 3-58 Cutting CFRP. ............................................................................................. 160

Figure 3-59 Sample A. .................................................................................................... 161

Figure 3-60 Sample B. .................................................................................................... 162

Figure 3-61 Sample C. .................................................................................................... 163

Figure 3-62 CYMDOF valve. ......................................................................................... 164


18

LIST OF FIGURES-Continued

Figure 3-63 Placing Teflon (a) and sample holding container(b). .................................. 165

Figure 3-64 Soil placed in lower sample holder. ............................................................ 165

Figure 3-65 Using bolts to attach CFRP to bottom of upper guide box. ........................ 166

Figure 3-66 Aluminum block is placed on top of the upper guide box. ......................... 167

Figure 3-67 Interface result for backfill soil and sample A with n=17.5 kPa. .............. 174

Figure 3-68 Tangential displacement vs. normal displacement for backfill soil & sample

A with n=17.5 kPa. ................................................................................................ 174

Figure 3-69 Interface result for backfill soil and sample A with n=35 kPa. ................. 175

Figure 3-70 Tangential displacement vs. normal displacement for backfill soil & sample

A with n=35 kPa. ................................................................................................... 175

Figure 3-71 Interface result for backfill soil and sample A with n=70 kPa. ................. 176

Figure 3-72 Tangential displacement vs. normal displacement for backfill soil & sample

A with n=70 kPa .................................................................................................... 176

Figure 3-73 Interface result for backfill soil and sample A with n=140 kPa. ............... 177

Figure 3-74 Tangential displacement vs. normal displacement for backfill soil & sample

A with n=140 kPa. ................................................................................................. 177

Figure 3-75 Interface result for backfill soil and sample A with n=210 kPa ................ 178

Figure 3-76 Tangential displacement vs. normal displacement for backfill soil & sample

A withn=210 kPa ................................................................................................... 178

Figure 3-77 Interface result for backfill soil and sample A with n=350 kPa. ............... 179
19

LIST OF FIGURES-Continued

Figure 3-78 Tangential displacement vs. normal displacement for backfill soil & sample

A with n=350 kPa .................................................................................................. 179

Figure 3-79 Interface result for backfill soil and sample A with n=525 kPa ................ 180

Figure 3-80 Tangential displacement vs. normal displacement for backfill soil & sample

A with n=525 kPa .................................................................................................. 180

Figure 3-81 Interface result for backfill soil and sample A with n=875 kPa. ............... 181

Figure 3-82 Tangential displacement vs. normal displacement for backfill soil & sample

A with n=875 kPa. ................................................................................................. 181

Figure 3-83 Interface result for backfill soil and sample A with n=1050 kPa. ............. 182

Figure 3-84 Tangential displacement vs. normal displacement for backfill soil & sample

A with n=1050 kPa. ............................................................................................... 182

Figure 3-85 Normal stress vs. max shear stress for sample A and backfill soil. ............ 184

Figure 3-86 Interface result for backfill soil and sample B with n=17.5 kPa. .............. 185

Figure 3-87 Tangential displacement vs. normal displacement for backfill soil & sample

B with n=17.5 kPa. ................................................................................................ 185

Figure 3-88 Interface result for backfill soil and sample B with n=35 kPa. ................. 186

Figure 3-89 Tangential displacement vs. normal displacement for backfill soil & sample

B with n=35 kPa .................................................................................................... 186

Figure 3-90 Interface result for backfill soil and sample B with n=70 kPa .................. 187
20

LIST OF FIGURES-Continued

Figure 3-91 Tangential displacement vs. normal displacement for backfill soil & sample

B with n=70 kPa .................................................................................................... 187

Figure 3-92 Interface result for backfill soil and sample B with n=140 kPa. ............... 188

Figure 3-93 Tangential displacement vs. normal displacement for backfill soil & sample

B with n=140 kPa. ................................................................................................. 188

Figure 3-94 Interface result for backfill soil and sample B with n=210 kPa. ............... 189

Figure 3-95 Tangential displacement vs. normal displacement for backfill soil & sample

B with n=210 kPa. ................................................................................................. 189

Figure 3-96 Interface result for backfill soil and sample B with n=350 kPa. ............... 190

Figure 3-97 Tangential displacement vs. normal displacement for backfill soil & sample

B with n=350 kPa. ................................................................................................. 190

Figure 3-98 Interface result for backfill soil and sample B with n=525 kPa. ............... 191

Figure 3-99 Tangential displacement vs. normal displacement for backfill soil & sample

B with n=525 kPa. ................................................................................................. 191

Figure 3-100 Interface result for backfill soil and sample B with n=875 kPa. ............. 192

Figure 3-101 Tangential displacement vs. normal displacement for backfill soil & sample

B with n=875 kPa. ................................................................................................. 192

Figure 3-102 Interface result for backfill soil and sample B with n=1050 kPa. ........... 193

Figure 3-103 Tangential displacement vs. normal displacement for backfill soil & sample

B with n=1050 kPa. ............................................................................................... 193


21

LIST OF FIGURES-Continued

Figure 3-104 Normal stress vs. max shear stress for sample B and backfill soil. .......... 195

Figure 3-105 Interface result backfill for soil and Sample C with n=17.5 kPa. ............ 196

Figure 3-106 Tangential displacement vs. normal displacement for backfill soil & Sample

C with n=17.5 kPa. ................................................................................................ 196

Figure 3-107 Interface result for backfill soil and Sample C with n=35 kPa. ............... 197

Figure 3-108 Tangential displacement vs. normal displacement for backfill soil & Sample

C with n=35 kPa. ................................................................................................... 197

Figure 3-109 Tangential displacement vs. normal displacement for backfill soil & Sample

C with n=70 kPa. ................................................................................................... 198

Figure 3-110 Interface result for backfill soil and Sample C with n=140 kPa. ............. 199

Figure 3-111 Tangential displacement vs. normal displacement for backfill soil & Sample

C with n=140 kPa. ................................................................................................. 199

Figure 3-112 Interface result for backfill soil and Sample C with n=210 kPa. ............. 200

Figure 3-113 Tangential displacement vs. normal displacement for backfill soil & Sample

C with n=210 kPa. ................................................................................................. 200

Figure 3-114 Interface result for backfill soil and Sample C with n=350 kPa .............. 201

Figure 3-115 Tangential displacement vs. normal displacement for backfill soil & Sample

C with n=350 kPa. ................................................................................................. 201

Figure 3-116 Interface result for backfill soil and Sample C with n=525 kPa. ............. 202
22

LIST OF FIGURES-Continued

Figure 3-117 Tangential displacement vs. normal displacement for backfill soil & Sample

C with n=525 kPa. ................................................................................................. 202

Figure 3-118 Interface result for backfill soil and Sample C with n=875 kPa. ............. 203

Figure 3-119 Tangential displacement vs. normal displacement for backfill soil & Sample

C with n=875 kPa. ................................................................................................. 203

Figure 3-120 Interface result for backfill soil and Sample C with n=1050 kPa ............ 204

Figure 3-121 Tangential displacement vs. normal displacement for backfill soil & Sample

C with n=1050 kPa. ............................................................................................... 204

Figure 3-122 Normal stress vs. max shear stress for sample C and backfill soil. .......... 206

Figure 4-1 Schematic of Stress-Strain Response of linear and nonlinear Elastic Body

(Desai, 2001). .......................................................................................................... 211

Figure 4-2 Schematic of Stress-Strain Response in Elastoplastic Deformaion (Desai,

2001). ...................................................................................................................... 212

Figure 4-3 Representation of yield surface for von Mises and Mohr-Coulomb criteria in

the octahedral (Desai, 2001). .................................................................................. 214

Figure 4-4 Loose sand and dense sand. .......................................................................... 216

Figure 4-5 J1-J2D (Desai, 2001). .................................................................................... 219

Figure 4-6 Octahedral plane (Desai, 2001). .................................................................... 220

Figure 4-7 1-2-3 space (Desai, 2001). ........................................................................ 220


23

LIST OF FIGURES-Continued

Figure 4-8 Decomposition of behavior for observed stress-strain as normally consolidated

soil and of part causing overconsolidaton (Desai, 2001). ....................................... 221

Figure 5-1 Typical comparisons between predicted and observed responses for soil and

interface (Desai and El-Hoseiny, 2005). ................................................................. 226

Figure 5-2 Interface between CFRP and backfill soil..................................................... 227

Figure 5-3 Normal Displacement vs. Normal Stress. ..................................................... 229

Figure 5-4 Variation of Normal Stiffness with Normal Stress for Interface. ................. 230

Figure 5-5 Determination of interface shear stiffness, ks. .............................................. 231

Figure 5-6 Variation of shear stiffness vs. normal stress ................................................ 232

Figure 5-7 Determination of hardening parameters (Desai, 2001). ................................ 234

Figure 5-8 Yield and ultimate surface (Desai, 2001). ..................................................... 235

Figure 5-9 State of stress at transition point. .................................................................. 236

Figure 5-10 Ultimate shear stress vs. Normal stress. ..................................................... 237

Figure 5-11 1.1*2 vs. n2. ............................................................................................. 238

Figure 5-12 Transition shear stress vs. normal stress ..................................................... 239

Figure 5-13 Determination of hardening parameters for n=70 kPa. ............................ 240

Figure 5-14 Determination of hardening parameters for n=140 kPa. ........................... 241

Figure 5-15 Determination of hardening parameters for n=210 kPa. ........................... 241

Figure 5-16 Determination of hardening parameters for n=350 kPa. ........................... 242

Figure 5-17 Determination of hardening parameters for n=525 kPa. ........................... 242
24

LIST OF FIGURES-Continued

Figure 5-18 Determination of hardening parameters for n=875 kPa. ........................... 243

Figure 5-19 Determination of hardening parameters for n=1050 kPa. ......................... 243

Figure 5-20 Plot for determination of and q................................................................. 245

Figure 5-21 Shear stress vs. normal stress at transition. ................................................. 246

Figure 5-22 Determination of hardening parameters for n=70 kPa (q=1.54). .............. 247

Figure 5-23 Determination of hardening parameters for n=140 kPa (q=1.54). ............ 247

Figure 5-24 Determination of hardening parameters for n=210 kPa (q=1.54). ............ 248

Figure 5-25 Determination of hardening parameters for n=350 kPa (q=1.54). ............ 248

Figure 5-26 Determination of hardening parameters for n=525 kPa (q=1.54). ............ 249

Figure 5-27 Determination of hardening parameters for n=875 kPa (q=1.54). ........... 249

Figure 5-28 Determination of hardening parameters for n=1050 kPa (q=1.54). ......... 250

Figure 5-29 Comparison between natural log of u vs. n of two different qs and lab

results. ..................................................................................................................... 252

Figure 5-30 Comparison between u vs. n of two different qs and lab results. ............ 252

Figure 5-31 Lab data and prediction of interface test (n=17.5 kPa).............................. 258

Figure 5-32 Lab data and prediction of interface test (n=35 kPa)................................. 258

Figure 5-33 Lab data and prediction of interface test (n=70 kPa)................................. 259

Figure 5-34 Lab data and prediction of interface test (n=140 kPa)............................... 259

Figure 5-35 Lab data and prediction of interface test (n=210 kPa)............................... 260

Figure 5-36 Lab data and prediction of interface test (n=350 kPa)............................... 260
25

LIST OF FIGURES-Continued

Figure 5-37 Lab data and prediction of interface test (n=525 kPa)............................... 261

Figure 5-38 Lab data and prediction of interface test (n=875 kPa)............................... 261

Figure 5-39 Lab data and prediction of interface test (n=1050 kPa)............................. 262

Figure 5-40 Lab data and prediction of interface test (n=17.5 kPa).............................. 264

Figure 5-41 Lab data and prediction of interface test (n=35 kPa)................................. 264

Figure 5-42 Lab data and prediction of interface test (n=70 kPa)................................. 265

Figure 5-43 Lab data and prediction of interface test (n=140 kPa)............................... 265

Figure 5-44 Lab data and prediction of interface test (n=210 kPa)............................... 266

Figure 5-45 Lab data and prediction of interface test (n=350 kPa)............................... 266

Figure 5-46 Lab data and prediction of interface test (n=525 kPa)............................... 267

Figure 5-47 Lab data and prediction of interface test (n=875 kPa)............................... 267

Figure 5-48 Lab data and prediction of interface test (n=1050 kPa)............................. 268

Figure 5-49 Independent laboratory tests and predictions of the HISS model with n=280

kPa........................................................................................................................... 269

Figure 5-50 Independent laboratory tests and predictions of the HISS model with n=700

kPa........................................................................................................................... 269

Figure 6-1 Hyperbolic stress-strain relation in primary loading for a standard drained

triaxial test (Plaxis manual, 2006). ......................................................................... 276

Figure 6-2 Definition of Eoedreff in oedometer test results (Plaxis manual, 2006)........... 277
26

LIST OF FIGURES-Continued

Figure 6-3 Representation of total yield contour of the Hardening soil model in principal

stress space for cohesionless soil (Plaxis manual, 2006). ....................................... 279

Figure 6-4 Location of instruments for wall panel 26-32 (Berg et al., 1986). ............... 282

Figure 6-5 FE Mesh for Tucson MSE wall panel 26-32 in Plaxis. ................................. 286

Figure 6-6 Vertical stress vs. distance from wall face for panel 26-32 face at elevation

5.0 ft. ....................................................................................................................... 286

Figure 6-7 Vertical soil strain vs. distance from wall for panel 26-32 face at elevation 3.5

ft. ............................................................................................................................. 287

Figure 6-8 Geogrid strain vs. distance from wall for panel 26-32 face at elevation 0.5 ft.

................................................................................................................................. 287

Figure 6-9 Geogrid strain vs. distance from wall for panel 26-32 face at elevation 1.5 ft.

................................................................................................................................. 288

Figure 6-10 Geogrid strain vs. distance from wall for panel 26-32 face at elevation 4.5 ft.

................................................................................................................................. 288

Figure 6-11 Geogrid strain vs. distance from wall for panel 26-32 face at elevation 11.5 ft.

................................................................................................................................. 289

Figure 6-12 Soil strain vs. distance from wall for panel 26-32 face at elevation 8ft. ..... 289

Figure 6-13 MSE wall using FRP. .................................................................................. 294

Figure 6-14 Lateral forces on MSE wall......................................................................... 295


27

LIST OF FIGURES-Continued

Figure 6-15 Vertical stress vs. distance from wall face for CFRP and Tensar SR2 at

elevation 5.0 ft. ....................................................................................................... 298

Figure 6-16 Vertical soil strain vs. distance from wall for CFRP and Tensar SR2 at

elevation 3.5 ft. ....................................................................................................... 298

Figure 6-17 Horizontal stress vs. elevation of the wall for CFRP and Tensar SR2. ...... 299

Figure 6-18 Geogrid strain vs. distance from wall for CFRP and Tensar SR2. ............. 299

Figure 6-19 Soil strain vs. distance from wall for CFRP and Tensar SR2 at elevation 8ft.

................................................................................................................................. 300

Figure 6-20 Movement of the wall face for CFRP and Tensar SR2. .............................. 300

Figure 6-21 Axial force in CFRP and Geogrids. ............................................................ 301

Figure 6-22 Effect of reinforcement spacing and wall deformation ............................... 303

Figure 6-23 Plastic points for different spacing between reinforcements (CW) ............ 307

Figure 6-24 Effect of vertical spacing on Msf and wall deformation for CW MSE-wall

................................................................................................................................. 309

Figure 6-25 Plastic points for different spacing between reinforcements (CF) .............. 312

Figure 6-26 Effect of spacing on Msf and wall deformation (CF) ................................. 314

Figure 6-27 Comparison results for two cases ................................................................ 322


28

LISTS OF TABLES

Table 2-1 Factors Influence the Soil Reinforcement. ....................................................... 69

Table 2-2 Vertical compressive loads (N) for different cases (d=0.5 mm) (Phanikumar,

Prasad, and Singh, 2009). ......................................................................................... 80

Table 2-3 Durability (Agaiby and Colin, 1996)................................................................ 84

Table 2-4 Relationship among OIT; temperature and pressure (Hsuan and Li, 2005) ..... 87

Table 3-1 Index properties of investigated soil............................................................... 119

Table 3-2 One layer of carbon fiber. ............................................................................... 121

Table 3-3 Particle size distribution curve for the soil. .................................................... 156

Table 3-4 Compaction curve data. .................................................................................. 158

Table 3-5 Properties of the soil. ...................................................................................... 159

Table 3-6 TestStar programming for 17.5 kPa and 35 kPa normal stresses ................... 169

Table 3-7 TestStar programming for 70 kPa and 140 kPa normal stresses .................... 170

Table 3-8 TestStar programming for 210 kPa and 350 kPa normal stresses .................. 171

Table 3-9 TestStar programming for 525 kPa and 875 kPa normal stresses. ................. 172

Table 3-10 TestStar programming for 1050 kPa normal stress. ..................................... 173

Table 3-11 Normal stress vs. max shear stress for sample A and backfill soil............... 183

Table 3-12 Normal stress vs. max shear stress for sample B and backfill soil. .............. 194

Table 3-13 Normal stress vs. max shear stress for sample C and backfill soil. .............. 205

Table 5-1 Material Constants for Tanque Verde Pantano Wash Sand. .......................... 225

Table 5-2 HISS-0 constant parameters. ......................................................................... 228


29

LISTS OF TABLES-Continued

Table 5-3 Material constants for soil-FRP interface. ..................................................... 244

Table 5-4 Material constants for soil-FRP interface ....................................................... 251

Table 6-1 Parameters for Plaxis analysis. ....................................................................... 285

Table 6-2 Location of reinforcement in the MSE wall ................................................... 302

Table 6-3 Maximum wall deformation and Msf value for CW MSE wall ..................... 307

Table 6-4 Summary of FE analysis for different spacing (CF) ...................................... 313

Table 6-5 Evaluation of Msf with verification of the length of geogrid (CFU10) ......... 316

Table 6-6 Evaluation of Msf with verification of the length of geogrid (CFU20) ......... 317

Table 6-7 Elaluation of Msf with verification of the length of geogrid (2 layers CUF20)

................................................................................................................................. 317

Table 6-8 Axial force (lb/ft) fot CFRP ........................................................................... 318

Table 6-9 Factor of safety for CFRP............................................................................... 319

Table 6-10 Evaluation of Msf with verification of the length of geogrid (2 layers CUF20)

................................................................................................................................. 320
30

ABSTRACT

The main weakness of soil is its inability to resist tensile stresses. Civil engineers have

been trying to address this problem for decades. To increase the tensile and shear

strengths of soil, different methods of reinforcing such as using geosynthetics have been

used in different types of earth structures such as retaining walls, earth dams, slopes, etc.

Due to the excellent corrosion resistance of polymers, the use of geosynthetics has

increased dramatically in recent years. However, there are some significant problems

associated with geosynthetics, such as creep and low modulus of elasticity. In this

research, a new Geo-Composite which is made of Carbon Fiber Reinforced Polymer

(CFRP) is used to overcome some of the short comings of the existing geosynthetics.

The new Geo-Composite has all the benefits of the geotextiles plus higher strength,

higher modulus and no creep.

In first part of the investigation, over eighty experiments were carried out using direct

shear test. The interface properties of the Geo-Composite (CFRP) and fine sand were

investigated. Tests showed that the interface shear behavior between Geo-Composite and

fine sand depended on the normal forces during the curing of epoxy and curing age of

epoxy. The two methods used to prepare the specimen are pre-casting and casting in

place, and the results of these two methods are compared.

In the second part of the investigation, the pull-out test device was designed and

assembled using a triaxial loading device and a direct shear device. In the pull-out test,

the normal force applied by the triaxial loading and pull out force is applied by a direct
31

shear device. CFRP samples were prepared in the lab, and pre-cast and cast-in-place

samples were tested using fine sand. The pull-out force and corresponding displacements

of each of the materials were recorded and compared.

In the third part of the investigation, the behavior of the interface between coarse sand

and modified CFRP has been studied in larger scale using a device known as Cyclic

Multi Degree of Freedom (CYMDOF) device. A constitutive Model, Hierachical Single

Surface (HISS) model, is used to characterize the behavior of the interfaces. The

constitutive model is verified by predicting the laboratory behavior of interface.

In the forth part of the investigation, using the laboratory test data results, a finite element

procedure with the hardening model is used to simulate field behavior of a CFRP

reinforced earth retaining wall, and compare the results with a geotextile reinforced earth

retaining wall. This section shows the advantages and disadvantages of using CFRP in

MSE walls.
32

CHAPTER 1. INTRODUCTION

The core weakness of soil is its incapability to resist tensile stresses. Civil engineers have

attempted to address this problem for decades. To increase the tensile and shear strengths

of soil, differing methods of reinforcing have been used in various types of earth

structures such as retaining walls, earth dams, slopes, etc. Numerous reinforced earth

practices have been used throughout the world. One particularly common application of

reinforced earth is found in mechanically stabilized earth (MSE) walls. Currently there is

large a large range of reinforcement that is used for reinforced earth. Mechanically

stabilized earth consists of soil constructed with artificial reinforcing. MSE is commonly

used for retaining walls, bridge abutments, dams, and seawalls. The basic principles of

MSE have been seen in historical construction; however, the current form of MSE was

developed in the 1960s. The elements used for reinforcing can differ but they commonly

include steel and geosynthetics. The use of straw, sticks, and branches to reinforce adobe

bricks and mud dwellings has been traced to the beginning of human history, and in the

16th and 17th centuries, French engineers used sticks to reinforce dikes. Reinforcing

levees with branches is a long used technique in China. Other elements for reinforcement

have been used in by many individuals to prevent soil erosion. The modern use of soil

reinforcing for retaining wall construction was founded by French architect and engineer

Henri Vidal in the 1960s. The first MSE wall in the United States was built in 1971 on

State Route 39 near Los Angeles. It is estimated that since 1997, about 23,000 MSE walls
33

have been built throughout the world. The highest MSE wall built in the United States is

30 m (98 ft) high (Retaining wall, 2012).

The reinforcement materials of modern MSE can vary. Originally, long steel strips 50 to

120 mm (2 to 5 in) wide were used as reinforcement. These strips are often ribbed,

although not always, to increase friction. Often steel grids or meshes are also used as

reinforcement. Several kinds of geosynthetics can be used including geogrids and

geotextiles. Reinforcing geosynthetics can be composed of high density polyethylene,

polyester, and polypropylene. These materials may also be ribbed and are available in

various sizes and strengths (Retaining wall, 2012).

However, there are some significant problems associated with geosynthetics, such as

creep and low modulus of elasticity. In this research, a new Geo-Composite which is

made of Carbon Fiber Reinforced Polymer cured CFRP is used to overcome some of the

short comings of existing geosynthetics. The new Geo-Composite has all the benefits of

the current geotextiles plus higher strength, higher modulus of elasticity and no creep.

The general objective of this study is to establish some practical analytical models for

predicting the strength and behavior of soil material reinforced with CFRP. This general

objective was achieved through the following steps.

Investigate soils properties which are fine sand and coarse sand.

Perform a normal compression test on sand / CFRP interfaces which covers the

range of normal stresses expected in this research and includes the unloading and

reloading measurements;
34

Perform sets of laboratory tests using direct shear test, pull out test, and

CYMDOF test on different sands / CFRP interfaces.

Study the effect of different types of the CFRP samples (cast in place, precast

place), and different types of epoxy used for preparation CFRP.

Developing a model based on HISS to determine the constitutive model

parameters using the data from laboratory test results for back prediction.

A finite element procedure with the hardening model is used to simulate an earth

retaining wall reinforced by CFRP, and compare the results with a geotextile

reinforced earth retaining wall, which shows the benefits of using CFRP.

There are eight chapters in this thesis as follow:

Chapter 1: Presents the object and scope of this study.

Chapter 2: This chapter detailed literature review concerning the present study is

described. Different areas covered in this review include different techniques of

stabilization and numerical analysis of reinforced soil and interfaces.

Chapter 3: Describe the devices used in this investigation; direct shear test, pull out test,

and CYMDOF test. It presents the results of the various laboratory test performed and

discussion of these results.

Chapter 4: Explanation the development of various constitutive models and the proposed

model, which is based on the disturbed state concept.

Chapter 5: Describes the determination of model parameters and its verification and its

application in simulating direct shear interface tests. It reports the predicted results.
35

Chapter 6: Model a MSE wall using CFRP and comparing the results with geotextiles.

Chapter 7: Includes a summary and conclusions from this study and provides

recommendations for future research.

A flow chart showing the research program of this study is shown in Figure 1-1.

Figure 1-1 Flow chart for the research program.


36

CHAPTER 2. LITERATURE REVIEW

2.1. Soils Type

2.1.1. Saline Soil

Saline soils are separated into different collections, and they are well distributed all over

the earth including places like Australia (Akpokodje, 1985), California (Kinsman, 1969),

Middle East (Johnson et al., 1978; Tomlinson, 1978; Al-Amoudi et al., 1991), and Utah

(Lund et al., 1990). Salines are generally located in low land areas previously flooded

with saline water in hot arid climates where evaporation rates surpass precipitation rates.

In areas of agricultural and certain soil chemistry, these types of soils are named as Salt-

Affected soils (Bohn et al. 1979). The concentration of high salts varies depending on

where the soils are located. Saline soils are also identified as playa, salt playa, salina, and

sabkha (Fooks, 1976). Coastal Sabkhas are defined as broad flat desert areas of silty

sands including varying small portions of gypsum, anhydrite and other salts and a high

percentage of calcium carbonate developed by evaporation of sea water or groundwater

from high water tables (Ellis, 1973).

Mechanisms such as wind, temperature and biological factors work together to form

sabkhas. Sabkhas contain salt rich sandy layers that are typically within sandy,

calcareous mud. These soils commonly have high void ratios, they often creep under

large loads, and are loose. These soils also have high permeability, unpredictable

heterogeneity, and sandy to gritty textures. These soils have are known to have low
37

bearing resistance, and are characterized as a collapsing soil. There is a high probabiliy

that these kinds of soils will cause damage to structural elements.

The geotechnical properties of saline soils vary in the vertical direction and in the

horizontal direction. The variation of geotechnical properties in the horizontal direction

are related to the shoreline (Akili and Thorrance, 1981), while the vertical variations

signify successive stages in the development of the sebkha cycle. These types of

characteristics of Saline Soils create issues for some types of geotechnical structures

(Shehata et al., 1990; Al-Amoudi et al., 1991 ).

Generally, a Saline layer in its natural state has low strength (Abduljauvad et al., 1993),

and the average unconfined compressive strength of such a layer is approximately 18.6

kPa. Saline soils may lose a large amount of their shear strength because of wetting due

to leaching of salts that bond the soil grains (Al-Amoudi et al. 1991). The existing

sulfates can breakdown rocks that have high porosities and affect the natural and physical

behavior of foundations. High water tables and high salt contents can severely damage

surface roads in salina terrains; especially when aggregates are friable and porous. Salts

easily adhere to sub-structure materials such as concrete and its reinforcement and

intensify the process of deterioration. Soil replacements, vibrofloating, preloading heavy

tamping, and stone columns are some of the actions that can solve or improve the issues.

For example, using geotextile in sabka soil is one method to stabilize the soil, and it can

solve road construction problems with sabkha sub-grade (Abdulijauwad et. al. 1993).
38

2.1.2. Soft Soil

Soft soils are a type of soil that is commonly known for its low shear strength and high

compressibility. Soft soils have a high shrinkswell potential and low permeability.

They have the least desirable engineering properties for construction purposes (Abdullah

and Al-Abadi 2005; Anagnostopoulos and Stavridakis 2003; Lo et al. 2000). When soft

soils are present, the goal of improvement methods is to increase the soils strength and

decrease the compressibility which improves a soils performance when loaded.

The moisture content, which is the ratio of the mass of water to the mass of solids in the

soil, is an important parameter in soft soils. Cohesive force and internal friction angle are

shear strength parameters of soil. Past research has shown from test data of numerous

soil samples that the value of soil water content may be linked with its shear strengths.

As an example, smaller water contents gives larger values of cohesive force and internal

friction angle (Yang, 2007).

2.1.3. Collapsing Soils

Unsaturated soil that undergoes a thorough repositioning of particles and changes in

volume due to additional water by way of or devoid of additional loading is called

collapsing soil (Bara, 1977). The quantity and rate of collapse is a function of grain

shape, grain size, existing moisture content, types and amounts of clay present, void ratio,

adsorbed ions, etc. Maximum collapse ensues at optimum clay content (10% in a single

case) and optimum moisture content (between 13% and 39%). Most of the collapse may
39

occur sometime before the soil reaches 100% saturation (Dudley, 1970). Collapsing

soils typically define a large reduction in bulk volume when the saturation degree rises.

From laboratory testing, large decreases on bearing resistance and shear strength can

occur in the presence of wetting or deformations (Youssef, 1984), and it is problematic

to build any type of structure with this kind of soil. These soils are found throughout the

world, typically in areas that have inadequate economic resources.

Aeolian, water deposited, and residuals soils have collapsible grain arrangements

(Dudley 1970). The composition of these types of deposits is usually loose with bulky

shaped grains, and the array of the sizes is between silt and fine sand. Collapsing soil is

often found in arid and semiarid locations and it is usually deposited by wind or water.

Sands (aeolian) and silts (loess) are deposited by wind in most arid and semiarid locations.

Collapsible soils may be produced by mudflows, colluvial deposits, alluvial flood,

residual soils, and volcanic tuffs. The deposits that are carried by wind and the size of

the particles that are close to silt are called Loess soil. Loess soil it found in many areas

throughout the earth (Koerner, 1985) and they tend to deform under foundation or

building loads on top of the soil as the degree of saturation rises.

The bonding elements between some types of collapsing soil can be salt, and these soils

can be dispersive and metastable, if the sodium content is high in the salt. When the

shear stress between the bonds of the particle exceeds the shear strength of the bonding

mechanisms, the soil will collapse.


40

Most soils collapse due to the interaction of clay particles between the bonds of sand

grains. Unsaturated sands and soils can collapse under load because of wetting (Knight

1963). Barden (1973) stated that there are three situations in which the structure will

collapse, and they are as follows: 1. structures that are potentially unstable, partially

saturated; 2. a large enough value of an applied existing stress component to cause a

metastable condition; 3. A strong bonding or cementing agent to stabilize intergranular

contacts, with a reduction caused by wetting, will produce collapse.

From the above statements, a collapsing soil is defined as a soil which rapidly

experiences a considerable loss of volume due to a major modification of the soil fabric

upon wetting or upon the application of a boundary force. A boundary force may or may

not be required to induce collapse.

2.1.4. Modification of Soil Weakness

It has been well documented that soil is a material that is weak in tension, and in certain

cases such as in sand, it does not have any cohesion. As mentioned earlier, most soils are

weak and have high compressibility. There are many different techniques available to

improve the properties of soil in order to reduce cost. Replacing available soil with a

higher quality soil is one type of technique used to improve a soils properties. Some

other procedures used include using geotextile to increase the tensile strength of soil. This

researchs emphasis is on the study of a new form of soil reinforcement utilizing carbon

fiber fabrics.
41

2.2. Soil Stabilization

There are different methods used to stabilize soil in order to reduce the deformation of

the soil under loading, to increase the shear strength, to increase the friction angle, and to

improve the volume stability, permeability and durability. These methods frequently

increase the strength of the soil by suitable drainage or enhancing the bond between

particles. Some of these techniques are described below:

1. Soil Stabilization with Cement Grouts

Soil stabilization with cement grouts injected under pressure is widely used in

construction. The core idea of grouting is that a cement grout, which has a required

strength, is injected into the pores of a sandy silt, or into the cracks of rocks under

pressure by bore holes (Ibragimov, 2005). Currently, the method of grouting is very

customary in a number of areas, and the method is quite economical. It does not require

intricate equipment, and it is also safe for the environment.

2. Soil Stabilization with Chemicals

Chemical soil stabilization was first used in the 1930s and it has been widely used during

the past 20 years. Over this recent period it has been widely applied in foundation

engineering, mining, and hydraulic construction. The experience of using chemical soil

stabilization by injecting chemical soil stabilization shows that this method is very

reliable and in many cases it is the only possible method for strengthening weak soils at

the base of existing buildings and proposed buildings. Chemical soil stabilization has

been used to preserve important structures and unique historic monuments, to construct
42

many retaining walls and antifiltration curtains, and to drive mine excavations under

complex mining--geologic conditions. Soviet hydraulic engineers constructed one of the

world's largest antifiltration curtain at the base of the Aswan High Dam. The stabilized

soil volume was about 2 million m (Sokolovich, 1987).

3. Mechanical Stabilization

Different particle sizes can be added to existing soil to change the grading size and the

degree of uniformity. This addition of particle sizes increases the magnitude of the

friction angle and cohesion (Ingles, 1973). This method is used before beginning

construction because it is problematic to perform this procedure during construction.

Chemical stabilization or grouting are preferred once construction has started.

4. Thermal Stabilization

The method of thermal stabilization of soils is has become more frequently used in

construction to increase the bearing capacity of bases and foundations of a building and

structures. Russian scientists and engineers have developed and presented several

methods of thermal stabilization which permit roasting soil masses through boreholes

with a diameter of 0.1-0.3 m and depth to 20 m (Sychev, 1954). As a result of heat

treatment of soils their strength increases significantly; this increase depends on the

monolithic stage of the roasted soil.

5. Soil Improvement Using Stone Columns

There are various techniques for improving in situ ground conditions, and reinforcing the

soil with stone columns or granular piles is one of the most versatile and cost effective
43

methods. Stone columns provide the primary functions of reinforcement and drainage by

improving the strength and deformation properties of soft soil. They increase the unit

weight of soil, quickly drain the excess pore pressures produced and act as strong and

stiff elements and carry higher shear stresses. Stone columns can be installed in an

extensive variety of soils, ranging from loose sands to soft compressible clays. Stone

columns have been and are being used in many problematic foundation sites throughout

the world to increase the bearing capacity, reduce settlements, increase the rate of

consolidation, improve slope stability of embankments and also improve liquefaction

resistance (Shivashankar, 2010).

6. Biotechnical Stabilization

This type of stabilization is characterized by the combined use of vegetation with

structural-mechanical elements in a ground cover system. Extreme rainfall events

produce shallow slope failures by elevating pore pressures and decreasing effective stress

which influences the potential for slope instability. One Biotechnical method uses tree

roots which have the ability to resist tension. This method increases the shear strength of

shallow soils through mechanical reinforcement (Roering, 2003).

7. Soil Reinforcement

Soil Reinforcement is another technique that is commonly used to stabilize soil. Soil

reinforcement is made up of two main parts, the tension element and the soil. Soil is the

most plentiful and least expensive construction material which has proper strength to

carry compressive stresses, but has almost no tensile strength. Soil dilatancy reports the
44

volumetric change which associates dense soil during shear processes. By adding linear

or planer reinforcements that are strong in tension to the soil, a composite material can be

created. This material is similar to reinforced concrete and has enhanced strength

characteristics than when unreinforced. The soil can be constructed at angles much

steeper than the internal angle of friction of the soil.

2.2.1. Reinforcement Geometry

Three types of reinforcement geometry can be considered:

Linear unidirectional

Strips, including smooth or ribbed steel strips, or coated geosynthetic strips over a load-

carrying fiber

Composite unidirectional

Grids or bar mats characterized by grid spacing greater than 150 mm (6 inches)

Planar bidirectional. Continuous sheets of geosynthetics

Welded wire mesh, and woven wire mesh. The mesh is characterized by element spacing

of less than 150 mm (6 inches)

2.2.2. Reinforcement Material

Distinctions can be made between the characteristics of metallic and nonmetallic

reinforcements:

Metallic reinforcements
45

Typically of mild steel and the steel is usually galvanized or may be epoxy coated

Nonmetallic reinforcements

Generally polymeric materials consisting of polypropylene, polyethylene, or polyester

The performance and durability considerations for these two classes of reinforcement

vary considerably and are detailed in the companion Corrosion/Degradation document

2.2.3. Reinforcement Extensibility

There are two classes of extensibility:

Inextensible

The deformation of the reinforcement at failure is much less than the deformability of the

soil

Extensible

The deformation of the reinforcement at failure is comparable to or even greater than the

deformability of the soil

2.2.4. Facing Systems

Special facing elements are used in MSE wall construction because they are the only

visible parts of the completed structure. A large range of finishes and colors can be

provided in the facing. Facings provide protection against backfill sloughing and erosion,
46

and facings may also provide drainage paths. The type of facing influences settlement

tolerances. Some major facing types are:

Segmental precast concrete panels

Precast concrete panels typically have a minimum thickness of 140 mm (5- inches) and

are of a cruciform, square, rectangular, diamond, or hexagonal geometry. Temperature

and tensile reinforcement are necessary and will vary with the size of the panel.

Vertically adjacent panels are usually connected with shear pins.

Dry cast modular block wall units

These units are usually relatively small, square concrete units that are designed and

manufactured specifically for retaining wall uses. The mass of these units generally

ranges from 15 to 50 kg (30 to 110 lbs), with units of 35 to 50 kg (75 to 110 lbs) routinely

used for highway projects.

Unit heights often range from 100 to 200 mm (4 to 8 inches). The exposed face length

usually varies from 200 to 450 mm (8 to 18 inches). The nominal width (dimension

perpendicular to the wall face) of units typically ranges from 200 to 600 mm (8 and 24

inches). Units may be manufactured as a solid or with a core. Full height cores are filled

with aggregate during construction. Units are normally dry-stacked (i.e. without mortar)

and in a running bond formation. Vertically adjacent units may be connected with shear
47

pins, lips, or keys. They are referred to by trademarked names such as Keystone, Versa-

Lok, Allan etc.

Metallic Facings

Original reinforced earth systems had facing elements composed of galvanized steel

sheets formed into half cylinders. Precast concrete panels are now commonly used in

reinforced earth walls, however, metallic facings may be appropriate in structures where

difficult access or handling necessitates lighter facing elements.

Welded Wire Grids

Wire grid can be bent up at the front of the wall to form the wall face. This type of facing

is used in the Hilfiker, Tensar, and reinforced earth wire retaining wall systems.

Gabion Facing

Gabions (rock-filled wire baskets) can be used as facing with reinforcing elements

consisting of welded wire mesh, welded bar-mats, geogrids, geotextiles or the double-

twisted woven mesh placed between or connected to the gabion baskets.

Geosynthetic Facing

Numerous types of geotextile reinforcement are looped around at the facing to form the

exposed face of the retaining wall. These faces are vulnerable to ultraviolet light
48

degradation, vandalism (e.g. target practice) and damage due to fire. Alternately, a

geosynthetic grid used for soil reinforcement can be looped around to form the face of the

completed retaining structure in a similar manner to welded wire mesh and fabric facing.

Vegetation can grow through the grid structure and can provide both ultraviolet light

protection for the geogrid and an aesthetically pleasing appearance.

Postconstruction Facing

For wrapped faced walls, the facing whether geotextile, geogrid, or wire mesh can be

attached after construction of the wall by shotcreting, guniting, cast-in-place concrete or

attaching prefabricated facing panels made of concrete, wood, or other materials. This

multi-staging facing approach adds cost but is advantageous where substantial settlement

is expected.

Precast elements can be cast in several shapes and can have facing textures that match

environmental requirements and blend aesthetically into the environment. Retaining

structures using precast concrete elements as the facings can have surface finishes similar

to any reinforced concrete structure.

Retaining structures with metal facings have the disadvantage of shorter life because of

corrosion, unless provisions are made to compensate for it. Facings using welded wire or

gabions have the disadvantages of an uneven surface, exposed backfill materials, more

tendency for erosion of the retained soil, possible shorter life from corrosion of the wires,

and more susceptibility to vandalism. These disadvantages can, of course, be countered


49

by providing shotcrete or by hanging facing panels on the exposed face and

compensating for possible corrosion. The greatest advantages of such facings are low

cost, ease of installation, design flexibility, good drainage (depending on the type of

backfill) that provides increased stability, and possible treatment of the face for

vegetative and other architectural effects. The facing can easily be modified and well-

blended with a natural environment. These facings, as well as geosynthetic wrapped

facings, are especially advantageous for construction of temporary or other structures

with a short-term design life.

Dry cast segmental block facings may raise some concerns as to durability in aggressive

freeze-thaw environments when produced with a water absorption capacity significantly

higher than that of wet-cast concrete. Historical data provides little understanding as their

usage history is less than two decades. Since the cement is not completely hydrated

during the dry cast process, (as is often evidenced by efflorescence on the surface of

units), a highly alkaline regime may establish itself at or near the face area, and may limit

the use of some geosynthetic products as reinforcements. Freeze-thaw durability is

enhanced for products produced at higher compressive strengths and low water

absorption ratios.

2.2.5. Reinforcement Types

Most, although not all systems using precast concrete panels use steel reinforcements

which are typically galvanized but may also be epoxy coated. Two types of steel

reinforcements are in current use:


50

Steel strips

The currently commercially available strips are ribbed top and bottom, 50 mm (2 inches)

wide and 4 mm (5/32-inch) thick. Smooth strips 60 to 120 mm (2-d to 4-inch) wide, 3

to 4 mm (c to 5/32-inch) thick have been used.

Steel grids

Welded wire grid using 2 to 6 W7.5 to W24 longitudinal wire spaced at either 150 or 200

mm (6 or 8 inches). The transverse wire may vary from W11 to W20 and are spaced

based on design requirements from 230 to 600 mm (9 to 24 inches). Welded steel wire

mesh spaced at 50 by 50 mm (2 by 2-inch) of thinner wire has been used in combination

with a welded wire facing.

The following geogrids are widely used and available:

High Density Polyethylene (HDPE) geogrid

These are of uniaxial manufacture and are available in up to 6 styles differing in strength.

PVC coated polyester (PET) geogrid

They are available from a number of manufacturers. They are categorized by bundled

high tenacity PET fibers in the longitudinal load carrying direction.

For longevity the PET is supplied as a high molecular weight fiber and is further

characterized by a low carboxyl end group number.


51

Geotextiles

High strength geotextiles can be used mainly in connection with reinforced soil slope

(RSS) construction. Both polyester (PET) and polypropylene (PP) geotextiles have been

used.

2.2.6. Application Area

This section focuses on some of the application areas for the use of earth reinforcement

and shows where soil structures of various forms have been found to provide economic

and technical benefits. Each case is an illustration of the concept of earth reinforcement

but should not be taken as being the only solution to any problem. In practice each

application should be considered separately. Applications and techniques may often be

combined and the introduction of new construction materials enables other applications

to be considered. The variety and range of the areas of application for these techniques is

unlimited.

Bridgeworks (Figure 2-1)

Bridge abutment

There has recently been an increase of interest in the construction of reinforced soil

retaining walls to support bridge abutments in place of traditional pile foundations. Two

of the main reasons for the increased interest are the reduction in overall cost of the

project and the reduction, or potential elimination, of bridge bumps which arises from
52

differential settlement between a traditional pile supported abutment and the approach

road (Abu-Hejleh et al., 2002; Helwany et al., 2003).

Figure 2-1 Bridge abutment (Abu-Hejleh et al., 2002).

Dam(Figure 2-2)

Two examples are given below of the use of geomembranes to provide watertightness in

earth fill dams. First, a PVC geomembrane is used as an upstream facing on the Aubrac

dam and, secondly, a bitumen membrane used as an internal core in the Valence d'Albi

dam. The height of both these dams is around 15 m. An incident which occurred on the

Aubrac dam--slipping of a small portion of the facing during construction--led to

measurement of the friction at the geotextile-geomembrane interface on a 1-00 m x 1.00

m inclined plane. The advantage of this method is that it recreates test conditions that are

close to real conditions, particularly where normal stresses are very low (Girard H.,

Fischer S., Alonso E., 1990).


53

Figure 2-2 Construction of the Valence dAlbi Dam (Girard 1990).

Embankments (Figure 2-3)

Rebuilding collapsed embankments can be very costly and from an economic standpoint,

it would be more beneficial to reinforce the embankment so that it does not fail rather

than reconstruct. Today, advances in technology in material science have produced

geosynthetic materials for usage in various aspects of civil engineering. Geosynthetic

materials are used extensively in embankments to increase stability (Bergado and

Teerawattanasuk, 2008; Brianon and Villard, 2008; Chen et al., 2008; Li and Rowe,

2008; Rowe and Taechakumthorn, 2008; Sarsby, 2007). Geotextile layers increase the

embankment stability by virtue of two primary functions: tensile reinforcement and as a

drainage element reducing pore pressures. Most analysis of geotextile reinforced

embankments considers the effect of the tensile stiffness of the geotextiles.


54

Figure 2-3 Reinforced embankment (Bergado and Teerawattanasuk 2008).

Foundation (Figure 2-4)

Often, shallow foundations are built on top of existing cohesive soil deposits, resulting in

low bearing capacity and/or excessive settlement problems. This can cause structural

damage, decrease in the durability, and/or deterioration in the performance level.

Conventional treatment methods are either to replace part of the weak cohesive soil by a

sufficiently thick layer of stronger granular fill, or to increase the dimensions of the

footing, or a combination of two. An alternative and more economical solution is the use

of reinforced soil foundation (RSF). This can be done by either reinforcing cohesive soil

directly or replacing the poor soils with stronger granular fill in combination with

geosynthetics reinforcement. The resulting composite zone (reinforced soil mass) will

improve the load carrying capacity of the footing and provide better pressure distribution

on top of the underlying weak soils, therefore reducing the associated settlements.

Several experimental studies were conducted to evaluate the bearing capacity of footings
55

on reinforced sandy soil (Ghazavi et al., 2008), clayey soil (Chen et al., 2007), aggregate

(DeMerchant et al., 2002), and pond ash (e.g. Ghosh et al., 2005).

Figure 2-4 Geosynthetic membrane for foundation (Chen et al. 2007).

Highways

Geosynthetic reinforced unpaved roads are also easier and quicker solutions compared to

traditional alternatives, such as the use of greater fill heights or the replacement of the

poor foundation soil by a more capable one, which are solutions detrimental to the

environment. Goldfingle (2009) reports on the construction of a 5 to 6 m wide and 65 km

long geogrid reinforced access road for the construction of a 60-turbine wind farm in

Scotland which took only four months to construct


56

Land Fill (Figure 2-5)

Geosynthetics have been increasingly used in environmental protection works. In the case

of waste disposal areas their functions range from barriers to fluids or gases to

reinforcements of earth works and slopes. In engineered landfills, geomembranes are

commonly used to isolate waste from the surrounding ground and groundwater in order

to minimize the potential groundwater contamination. On the sides of a landfill,

geomembranes are placed on prepared sloped surfaces and anchored at the crest level

(Palmeira, 2002). Successive construction of a landfill includes the placement of soil

cover and waste layers up to various heights over these liners. This can result in

application of substantial down-slope shear stresses on the liners leading to development

of significant liner tension.

Figure 2-5 Land fill (Palmeira 2002).


57

Mining (Figure 2-6)

The fast growth seen in the past few years in mining exploration and operation has led to

a sharp increase in the use of a extensive range of geosynthetic materials by the mining

industry for all type of applications. Smith (2008) recounted that from 1987 to 2008 more

than 60 square kilometers of geomembrane liners were installed in leach pads alone.

Figure 2-6 Reinforced pack for roof supporting underground mining (Mining, 2010).

Pipeworks (Figure 2-7)

One of the main uses of geotextiles in the 1980s is in drainage applications. In North

America, the Netherlands and France, more than 20 million square meters of synthetic

geotextiles were installed in agricultural drainage systems (Rollinl 1986). An alternative

for drainage applications where a plastic or vitrified clay drainage pipe is enclosed by a

graded filter is to cover the pipe with a geotextile sock. Lennoz-Gratin 2 has shown that

in France, over 6000 km of wrapped drains were installed in 1983, while Stuyt and

Oosten 3 estimated that in the Netherlands new agricultural drains with value of US$25
58

million a year are installed, with about 80% of these drains being wrapped in some form

of covering.

Figure 2-7 Pipe structure (Chen, Y et al 2008).

Railways (Figure 2-7)

Building railway embankments over soft soils is a challenge for geotechnical engineers

because of the low shear strength of subgrade soil, which causes excessive consolidation

settlements and, occasionally, bearing capacity failure (e.g. Liu et al., 2008; Rowe and

Taechakumthorn, 2008). An assortment of ground improvement techniques, including

vertical drains, grouting, complete soil replacement, geosynthetic reinforcement, and

piling, have been developed to solve the problems.


59

Figure 2-8 Railway support for embankment( Liu et al., 2008).

Quays and sea walls and waterway structures (Figure 2-9),

Recently, traditional forms of river and coastal structures have become very costly to

construct and maintain, because of the shortage of natural rock. As a result, the materials

used in hydraulic and coastal structures are changing from traditional rubble and concrete

systems to low-cost materials and systems such as gabion, slags, and geosynthetics.

Shorelines are continually eroded by the wave action of the sea, and river and coastal

structures are often damaged by both anthropogenic and natural causes (Yasuhara, 2007).

Geosynthetics are being increasingly used in civil and environmental applications. One

of these applications is the use of geotextile tube technology. Geotextile tubes,

hydraulically or mechanically filled with dredged materials, have been variously applied

in hydraulic and coastal engineering fields.


60

Figure 2-9 Sea wall (Yasuhara, 2007).

2.2.7. Theory

A ladder wall is presented, in which a series of reinforcing elements typically, but not

necessarily, have anchors that are connected to a facing to form a reinforced soil structure,

by Coyne in 1924 Figure 2-10.

Figure 2-10 Coyne-LadderWall.


61

The analogy of a Howe Beam was used to describe the action of the structures by Coyne

in Figure 2-11. The top flange and the diagonals are in compression, and the bottom

flanges are in tension. The Howe Beam appears to be erected vertically (i.e. like a ladder)

by rotating the beam through 90 as seen in Figure 2-12. In Coyness words, a beam

whose uprights, represented by the anchorages, are in tension and whose compressed

diagonals are formed in the fill itself. The compressed member of the beam is the facing

AC and its stretched member falls about the vertical plane BD, passing through the tail of

the anchorage. The corresponding extensions are neutralized by the weight of the fill.

The whole may be considered forming a single block of earth coherent in the whole zone

ABCD transversed by the tie rods.

Figure 2-11 Howe beam.


62

Figure 2-12 Howe beam analogy of ladder wall.

Later, Westergaard material (a medium made up of soft elastic material reinforced by

closely spaced horizontal flexible but un-stretchable sheets) was presented by

Westergaard (1938). He defines the properties of the material in terms of the theory of

elasticity, but Harrison and Gerrard (1972) illustrated that Westergaard material is

anisotropic material.

The behavior of soil reinforcement is similar to the behavior of reinforce concrete, in

which rebars are bonded to concrete, and reinforcements are bonded to the soil. It is not

fully possible to compare the two cases, because rebars in concrete beams are intended to

carry the tensile force, but reinforcement in soil are not necessary carrying tensile stresses.

The anisotropic reduction or suppression of one normal strain rate, and the soil particles

are tied together producing a form of pseudo-cohesion (Vidal 1969).


63

The vertical stress for cohesion-less soil at depth h for semi-infinite mass is equal to the

density of the soil times h (v = h), and the lateral stress when the soil is at rest is equal

to vertical stress times K0 ( H = vK0 = h K0) where K0 1- sin , = angle of friction

of the soil. As soon as soil expands laterally, the K0 value reduces to a smaller value

which is called Ka; Therefore, Lateral stress (h K0) reduces to (h Ka) where

Ka=tan(45-/2) (Figure 2-13).

Figure 2-13 Stress state in soil.

By applying vertical load to a soil sample, the sample deforms laterally with a strain of h

and vertically with a strain of v. The installation of horizontal reinforcements in soil

layers generates interaction or adhesion between the soil and adhesion caused by friction

and others. Since the reinforcement is stiff, the particles of soil are nearly rigidly
64

connected to the reinforcement by friction action. Thus, the gain in the strength of a

reinforced soil mass can be credited to an increase in confining pressure applied to the

soil (Figure 2-14). It is obvious that as vertical stress increases v, the lateral stress also

increases H. Therefore the stress circle for the reinforcement condition always lies below

the rupture curve (Figure 2-13). As soon as reinforcement ruptures or the adhesion

between the reinforcement gives way, the soil fails.

Figure 2-14 Soil & reinforcement.

The lateral stress (H = vK0 ) is equal to the force transferred from the unit of soil into

the reinforcement. Consequently, the tensile stress for each reinforcement is equal to

lateral stress divided by cross sectional area of the reinforcement (T= vK0/ar). The strain

in the reinforcement can be calculated by dividing tensile stress by elastic modulus of the

reinforcement (r = r = (vK0) / (arEr) ). As the stiffness of the reinforcement is increasing,

the lateral strain decreases; as a result, lateral strain can approach zero by increasing the

stiffness of the reinforcement. As lateral strain increases, K0 approaches Ka.


65

2.2.8. Failure Modes of Soil

As illustrated in Figure 2-15, there are three failure modes for reinforced slopes:

1. Internal is where the failure plane passes through the reinforcing elements.

2. External is where the failure surface passes behind and underneath the reinforced mass.

3. Compound is where the failure surface passes behind and through the reinforced soil

mass.

Figure 2-15 Failure modes for reinforced soil slopes (Allen and Bathurst, 2001).

1. Internal Failure Modes

Some examples of internal failure can be pullout failure, tensional failure on

reinforcement, and long term failure due to environmental effects such as corrosion,

creep, chemical and biological. The soil transfers the load to reinforcement without

rupture, and reinforced soil usually separates to two sections which are active zone and
66

resistant zone. Active zone tends to move from resisting zone; therefore, pullout

resistance is developed in the resisting zone. For that reason, the maximum tensile load

in reinforcement occurs between active zone and resisting zone.

Figure 2-16 Reinforced soil wall.

The maximum tensile stress per unit width of the wall in sheet of reinforcement can

calculated as follow:

Tmax=K v S = h S = K v SvSh (2-1)

h= Horizontal Stress
67

K= Lateral Earth Pressure Coefficient

Sv= Vertical Spacing between Reinforcement Layer and for Strip Reinforcement

Sh=Horizontal Spacing between Strip Reinforcement

In most cases, before rupture of reinforcement, the reinforcement may elongate and

deform. However, elongation and rotation of the reinforcement may be insignificant for

stiff inextensible reinforcement but may be significant with geosynthetics. As a result,

the component force in the direction of the failure surface would increase and the normal

component may increase or decrease.

2. External Failure Modes

External stability of a reinforced soil mass rests on on the capability of the mass to act as

a stable block and withstand all external loads without failure. Possible failure methods

are shown in Figure 2-17 and they include sliding, deep-seated overall instability, local

bearing capacity failure at the toe (lateral squeeze type failure), as well as

disproportionate settlement from both short-term and long-term conditions.


68

Figure 2-17 External failure modes for reinforced soil slopes (Allen and Bathurst, 2001).

2.3. Factors Effecting Soil Reinforcements

Some elements that influence the behavior and performance of reinforced soil include the

strength of reinforcement, soil particle size and density of soil. Table 2-1 shows these and

other factors that influence this behavior, and the following sections describe each of

these factors.

Reinforcement
Soil Soil State Reinforcement Distribution Construction

Geometry of
Paticle Size Density Form Location Structures

Surface
Grading Overburden Properties Orientation Compaction
69

Mineral Construction
Content State of Stress Dimensions Spacing System

Index Degree of
Properties Saturation Strength Aesthetics

Stiffness Durability

Table 2-1 Factors Influence the Soil Reinforcement.

2.3.1. Reinforcement

The formation of constant rupture surfaces through the soil is inhibited by reinforcement

which increases the stiffness and shear strength. Reinforcement when introduced into

soil and aligned with the tensile strain arc disturbs the uniform pattern of strain that

would develop if the reinforcement is not present.

Shear Test

1.6
S hear Lo ad Txy? S ig m a y

1.4
1.2
1
Unreinforced
0.8
Reinforced
0.6
0.4
0.2
0
0 2 4 6 8 10
Shear Displacement X (mm)

Figure 2-18 Load-displacement result for shear test (Noorzad & Mirmoradi, 2010).
70

From Figure 2-18 it can be that initially the reinforcement has no effect on soil until the

reinforcement begins to deform. As soon the soil starts deforming, it creates strength in

the soil to resist the shear loads, and then soil strain causes strain in the reinforcement.

Therefore, this causes an increase in strength in the reinforced soil. Strength is improved

until a limiting value is achieved, with further shear displacement the improvement

remains constant. Dyer (1985) using photo-elasticity, which showed the influence of the

presence of the reinforcement on the state of stress in the soil sample, as presented in

Figure 2-19 (a)(c). In these experiments, crushed glass was used in substitution of soil

and photo-elasticity techniques were employed. The normal stress was applied to the

sample top by a rigid plate. The photograph in Figure 2-19 (a) was taken during an

unreinforced test, whereas those in Figure 2-19 (b) and (c) were taken during tests on

samples reinforced by vertical and inclined steel grids, respectively. Bright regions in the

photographs are regions of high compressive stresses, whereas dark regions are regions

of low stress levels.

Figure 2-19 Results of photo-elastic studies in the direct shear test. (a) Unreinforced; (b) Vertical
reinforcement; (c) Inclined reinforcement(Dyer, 1985).
71

Form

One of the procedures to improve the performance of soil reinforcement is the form of

reinforcement that adheres to the soil. The deformation in soil introduces strain in the

reinforcement. There are a variety of reinforcement forms which dependent on the

application as discussed in section 2. The most common forms are sheets, bars, strips,

grids, and anchors (Figure 2-20). These types of forms create a friction bond between

soil and the reinforcement. Typically grids and anchors introduce a higher bond between

reinforcement and soil.

Figure 2-20 Common forms of reinforcement.

Surface Properties

The coefficient of friction is one of the most critical properties between soil and

reinforcement. The higher the friction the more efficient the reinforcement and using a

rougher surface is therefore significantly better than using a smooth surface. Rougher

surfaces are seen on reinforcement by grooves, ribs or embossing a pattern. Roughened

surfaces will tend towards the ideally rough condition depending on the depth and
72

spacing of grooves, ribs and, embossing and also depends on grading and particle size of

the soil. In a test (Palmeira, 2009), the frontal face of the wall consisted of a steel plate

with an anti-rust coating. In another other test, the wall had a smooth painted surface

which was lubricated with double layers of plastic films and oil. The results are shown in

Figure 2-17.

Figure 2-21 Rough and smooth surface (Palmeria, 2009).

Dimensions

Based on location within a wall and conditions, different size and dimensions of

reinforcement can be used. Several pull out tests were performed by Khedkar and

Mandal on different reinforcements and dimensions. They used different reinforcements

such as a sheet (Sh), C3 (3 mm high cellular reinforcements), C25, C30, C40, and C50.

Figure 2-17 denotes the comparison between load versus normalized displacement curves

for numerous heights of reinforcements under the normal pressures of 75 kPa. Pullout
73

displacement is normalized with the longitudinal cell dimension. From these results, it is

clear that the dimensions of reinforcements have an effect on soil reinforcement.

Figure 2-22 Different size of reinforcement (Khedkar and Mandal, 2009).

Figure 2-23 Pullout load versus displacement curve comparison for various reinforcement (Khedkar
and Mandal, 2009).

Strength

The strength of reinforcements is one of the most important characteristics in soil

reinforcements, and there are different types of reinforcement offered with different
74

strengths. Andrejack and Wartman (2010) have created the Multi-Axial Tension Test for

Geosynthetics, a new large-diameter experimental device capable of applying multi axial,

out-of-plane loading to relatively large multiples geosynthetic specimens (48 cm

diameter). Figure 2-24 shows representative pressure vs. strain data for three geotextiles.

This data has been corrected for the presence of a bladder. Both GT-1 and GT-2 have

similar ultimate strains of approximately 20%, while GT-3 has an ultimate strain of 15%.

The ultimate pressure at failure is markedly different for each of the geotextiles.

Figure 2-24 Pressure vs. strain data for various geotextiles (Andrejack & Wartman, 2010).

The ultimate tensile stress of reinforcements has a direct effect on the shear strength of

soil reinforcement. Therefore, losing tensile strength due to failure of reinforcement or

environmental effects like intensification in temperature (Figure 2-25) on reinforcement

would abruptly reduce the shear strength of the reinforced soil to shear strength of the

soil shown at an equal displacement.


75

Figure 2-25 Rupture stress-time-temperature relationship of PE geomembrane under axisymmetric


loading (Duvall, 1993).

Stiffness

In the design of geomembrane applications, the stressstrain curve needs to be known for

different types of reinforcement. Laboratory tests on geomembranes are usually

performed according to standards at a constant strain rate in order able to determine the

stiffness of the geomembranes (Wesseloo, Visser, and Rust, 2004). The stiffness of soil

reinforcement is governed by the reinforcement in the soil; which are the elastic modulus

and the effective cross sectional area and is called longitudinal stiffness (EaT). The

reinforcement in soil is placed in the direction of tensile strain in soil to restrict

deformation in soil reinforcement. This deformation creates forces in the reinforcement.

However, bending stiffness, which is related to the elastic modulus and the moment of
76

inertia, does not have any effect on the performance of reinforced soils except on very

soft soil.

2.3.2. Reinforcement Distribution

Location

For the most economical design the failure mechanisms and planes occurs when an

applied stress is larger than the failure stress of the structure. Stresses are separated into

the two categories of normal and shear stress. Normal stress is defined as stress

perpendicular to a plane in a material, while shear stress occurs parallel to a plane and

they are predicted to define the strain field. Therefore, reinforcements are placed in the

locations that have the most deformation or greatest tensile strains.

Orientation

Peak reinforcement load is proportional to the vertical spacing of the reinforcement.

However, very small vertical spacing or very large vertical spacing could result in a

deviation from this linear relationship. For example, at small spacing, it may be possible

to derive additional benefit (such as reduced reinforcement loads) as a result of improved

confinement of the soil. At some larger spacing, it is acceptable to assume that the

reinforced backfill will cease to function as a coherent mass and will require refinement

of the methodology. Analysis of full-scale structures and analytical models calibrated

against full-scale data where spacing of the reinforcement is the primary variable is

needed to clearly establish this relationship. The horizontal distribution of the

reinforcement is currently considered in the methodology as a modifier to the available


77

reinforcement tensile strength given as a force per unit width of wall. However, the use of

Rc in this way does not account for the possible effect of the reinforcement to confine the

soil (Allen and Bathurst, 2001).

2.3.3. Soil

The state of soil and the properties of the soil have a large influence on the behavior of

soil reinforcement. There are different types of soils used in soil reinforcement, and it

depends on the condition of the job site, type of geotextile used, and other conditions. In

some conditions, reinforcement is used to improve a weak soil such as a soft soil,

granular, and waste disposal material.

There are several particle sizes that exist in soil mechanics such as poorly graded (a soil

is uniformed-graded has most of its particles are about the same size), gap graded (a

soil is gap-graded has at least one particle size that is missing), and well graded

(a soil sample that has all sizes of material present from the No. 4 sieve to the No.

200 sieve).

The optimal particle size for a soil is well graded, because it drains well, provides long

term durability, is stable during construction, and has good physiochemical properties.

However, there is not always a possibility to reach well graded soil. In some countries

such as Japan, well graded soil is limited, and poor graded soil is used in soil

reinforcement. In this case, drainage systems are used to create negative pore water

pressures to provide the necessary stability. Therefore, it leads to the concept of

combined reinforcement and drainage. In the past two decades geotextile reinforcement
78

has been used extensively in many soil improvement situations. (Lorenzo et al., 2004;

Chu et al., 2004, 2006; Sarsby, 2007; Kazimierowicz-Frankowska, 2007; Bergado and

Teerawattanasuk, 2008; Basudhar et al., 2008; Brianon and Villard, 2008; Li and Rowe,

2008; Rowe and Taechakumthorn, 2008; Chen et al., 2008; Chattopadhyay and

Chakravarty, 2009; McCartney and 2010).

2.3.4. Soil State

Density

Dissimilar soil states will have different influences on soil reinforcement, and varied

densities of soils have a direct effect on stress and strain relationships in reinforcement

soil. Phanikumar, Prasad, and Singh (2009) address load test results obtained on different

types of sand beds such as fine, medium and coarse sand, reinforced with geogrids that

are 120 mm in diameter. The improvement in the loadsettlement response was

considered.
79

Figure 2-26 Effect of number of geogrids, n, on the load required for a settlement of 0.5 mm
(Phanikumar, Prasad, and Singh, 2009).

Figure 2-26 shows the effect of the amount of geogrids on the vertical compressive load

necessary for a settlement of 0.5 mm in different types of sand which has a different

density. The amount of geogrids (n) was varied as n =1, 2 and 3. The spacing between

the geogrids (s) was constant at 10 mm. In the case of tests with n=1 and n=2, the depth

to the top geogrid (u) from the base of the test plate was varied from 10 mm and 20 mm.

Table 2-2 summarizes the vertical compressive loads (N) required to be applied in

different cases to reach a settlement of 0.5 mm.


80

z
Table 2-2 Vertical compressive loads (N) for different cases (d=0.5 mm) (Phanikumar, Prasad, and
Singh, 2009).

Overburden

Overburden pressure has an effect on the friction angle between soil and reinforcement in

soil reinforcement. As overburden pressure increases, the coefficient of friction

decreases; thus, the peak angle of shearing stress of a granular soil also decreases with the

increase in normal stress.

The results of some tests are provided in Figure 2-27. It is obvious from the data that the

index failure angle is influenced by the overburden stress, although to a varying degree.

All of the testing was performed at an overburden stress of 1.6 kPa and interface 1,

interface 4, and interface 7 are S-HDPE, HDPE, and NW-NP-NH geotextile, respectively

(Narejo, 2003).
81

Figure 2-27 Influence of overburden stress on index friction angle (Narejo, 2003).

State of Stress

The state of stress within a reinforced structure is dissimilar with increasing height. The

void ratio decreases as the height of the soil increases because of the increase in normal

stress. The relative strain in the soil decreases and effective lateral stress inclines to the

active condition. In practice, at the top face, the coefficient of lateral earth pressure is at

rest, and lower down, the coefficient lateral pressure gets nearer to active pressure.

Ahmadabadi and Ghanbari (2009) examined active earth pressure distribution for

different cohesion strengths on a retaining wall which is 10 m high. A sample wall with

cohesive-frictional backfill and the effect of cohesion strength on the distribution of

active earth pressure is shown in (Figure 2-28). This distribution has a non-linear

association and the tension crack zone spreads as the cohesion strength of the soil

increases.
82

Figure 2-28 Active earth pressure distribution vs. height of wall (Ahmadabadi, and Ghanbari, 2009).

Degree of Saturation

A problem related to saturated soil is usually fine grained material and cohesive soils

which are usually poor in drainage have an effective stress transfer that may not be

immediate. Consequently, there would be an impermanent decrease in shear strength

which reduces the construction rate in order to stabilize the soil.

Raisinghani and Viswanadham (2010) show a dissimilarity of equivalent coefficient of

permeability with normal stress for different soils along with geosynthetic layers. With an

increase in normal stress, a decrease in equivalent coefficient of permeability in the soils

can be seen (Figure 2-29).


83

Figure 2-29 Coefficient of permeability vs. normal soil (Raisinghani and Viswanadham, 2010).

2.4. Durability

2.4.1. Introduction

The subject of the durability of geotextiles is one of the most quarrelsome issues in civil

engineering today. One of the most important issues, for soil reinforcement to be stable,

is durability of reinforcement. Soil does not provide the best environments for

reinforcement; therefore, proper consideration must be considered during design for

environmental hazards. One of the major problems with corrosion in soil is that it is

difficult to monitor the corrosion until a failure happens. Agaiby and Colin (1996),
84

identify three categories of structures, based on design life, and the relative importance of

durability and the rate of corrosion of the materials forming the structure in Table 2-3.

Design life and the importance of: Durability and rate of corrosion

Permanent structures: Major consideration

60-100 yrs. US

120 yrs. UK

Short-life structures: Minor consideration

1-20 yrs.

Temporary structures: No problem

1-100 wk.

Table 2-3 Durability (Agaiby and Colin, 1996).

2.4.2. Corrosion

Corrosion is the deterioration of a metal or its properties by chemical or electrochemical

reactions with the environment. When a large surface is affected it can be viewed as

general corrosion and estimated by an average fictitious uniform rate of corrosion per

year. If limited to small points so that definite indentations form in the metal surface, it is

denoted as pitting corrosion and generally described as maximum pit depth per year

(Christopher, Berg and Elias, 2001). Corrosion is essentially a return of metals to their

native state as oxides and salts. Only the more noble metals and copper exist in nature in

their metallic state. Other metals are refined by applying energy in the form of heat.
85

Unless protected from the environment, these metals revert by the corrosion process,

which is irreversible, from their temporary state to a more natural state.

Corrosion is an electrochemical process, it just occurs in metals, and it does not occur in

glass or plastic. Corrosion occurs by potential difference between two points which are

electrically connected, and this difference happens due to the existing difference of salt

and oxygen concentration in the soil. Corrosion occurs when metal is transferred in

solution in the form of positive ions or cations. The cathode reaction relates to the

electrons remaining in the parent metal. Hydrogen evolution and oxygen reduction are

two predominate cathodic reactions in reinforced soil.

2.4.3. Degradation

Different activities such as ultraviolet, light, high energy radiation, oxidization,

hydrolysis and chemical reaction will degrade polymeric materials, and increasing in

temperature would increase the rate of degradation (Richaud, Farcas, Divet and Benneton,

2008). These actions cause the reinforcement to become brittle; therefore, it decreases

the deformation along the reinforcement.

Effect of Ultraviolet light

Ultraviolet light generates degradation by reaction with the covalent bonds of organic

polymers which cause yellowing embrittlement. However, the effect of ultraviolet light

on geotextile can be overlooked by covering the reinforcement in the soil (ASTM D4355,

2007). Though, in some cases such as walls, face slope, or even during shipment and

storing, the reinforcement is positioned facing the sunlight for a period of time; therefore,
86

geotextiles must resist the effect of UV light. Goetextiles with a larger thicknesses are

more resistant to UV light than thin fibrous material.

Effect of oxidation

Degradation due to oxidation occurs as a result of an increase in pressure and temperature,

and in some cases exposure to UV light. Oxidation usually occurs on polyethylene and

polypropylene, but it does not affect polyester. Hsuan and Li (2005), attempted to

evaluate the oxidative resistance of high density polyethylene (HDPE) geogrids at

different temperatures and pressures over period of time. Figure 2-30 and Table 2-4

shows standard oxidation induction time (OIT) with different temperature and pressure.

As temperature and pressure increase, the rate of oxidation also increases.

Figure 2-30 Relationship among OIT; temperature and pressure (Hsuan and Li, 2005)
87

Table 2-4 Relationship among OIT; temperature and pressure (Hsuan and Li, 2005)

Effect of hydrolysis

The decrease of the allowable tensile strength in reinforcement is contingent on short-

term effects like installation damage, which reduces the maximum tensile strength but

does not affect the long-term properties like creep and aging by hydrolysis, which result

in long-term strength harm (Hufenus R., Regger, Flum and Sterba, 2005). The

hydrolysis occurring under the influence of water can represent a possible risk for

installed Geosynthetics, and water molecules react with polymer molecules which

reduces molecular weight and strength. Hydrolysis is a slow reaction. Its rate can

increase by influence of temperature and humidity. The reaction which takes place is

shown in Figure 2-31.


88

Figure 2-31 Chemical reaction.

2.5. Fiber reinforcement polymer (FRP)

2.5.1. Introduction

Composite materials (or composites for short) are engineered materials made from two or

more constituent materials with considerably different physical or chemical properties which

remain separate and distinct on a macroscopic level within the finished structure. The

definition of a composite material is usually limited to those which contain two base

materials in roughly equal amounts, with different mechanical characteristics. Alloys which

have only trace amounts of other metals are not composites, but Fiber Reinforced Polymers

(FRP) are composites and they are the main material in this study. Fiber Reinforced Polymer

(FRP) is a composite material comprising of a polymer matrix reinforced with fibers. The

fibers are usually glass, carbon, or aramid, while the polymer is usually an epoxy, vinylester

or polyester thermosetting resin. FRPs are often used in the aerospace, automotive, marine,

and construction industries. These materials have found prevalent acceptance in recent years
89

especially in the aerospace field and are special because of their high strength to weight ratio,

durability, and ability to form complex shapes.

Since composites are synthetic, the mechanical properties can be designed to suit the

particular application. For instance, space shuttles are made of carbon because carbon can

resist in high temperate without changing its material properties. Aramid fiber composites are

used to absorb ballistic impact in the protective vests worn by police officers. In the early

90s, Fiber Reinforced Polymer composite (FRP) began to be used more for rehabilitation

purposes in Civil Engineering. These days, FRP is used more because of its high elastic

modulus, high strength, light weight, corrosion resistance, and workability. Several structural

applications which have traditionally required concrete, masonry, or steel may provide new

opportunities to use fiber composites for retrofitting purposes.

A fiber-reinforced polymer (FRP) is a composite material comprising a polymer matrix

reinforced with fibers. The fibers are usually carbon, glass, and/or aramid, while the polymer

is usually an epoxy resin, vinylester or polyester thermosetting plastic. A composite structural

material contains high-strength fibers embedded in a resin matrix which together develops

mechanical properties greatly superior than those of the individual components. The fiber

reinforced polymer (FRP) industry today is experiencing substantial growth as more products

are made from reinforced plastic for greater durability, strength and life. Thousands of

products are now manufactured from reinforced plastics including building materials,

sporting equipment, appliances, automotive/aircraft parts, boat and canoe hulls, and bodies

for recreational vehicles.


90

2.5.2. Fiber

The strength of the Fiber Reinforced Polymer composite comes from the layers of fiber.

These are superior to larger cross sections because there are fewer places for weaknesses to

occur. In any large cross section of glass or any material, there would be several voids, small

cracks, small imperfections, and/or contamination from other materials.

The probability is high that any cross section of this size will have some microscopic defect.

These small irregularities develop points of stress concentration, as soon as the cross section

is subjected to a load. The cracks will start from there, and eventually the failure will occur.

One of the differences between the theoretical method of calculating strength of material and

actual strength of material are these small microscopic defects. If the piece of glass is drawn

into an extremely thin fiber, a space as small as a few microns, the statistical probability of a

microscopic flaw in any cross section is much less. Any such weakness would cause the fiber

to fail at that point, so the actual strength of a continuous fiber is extremely close to its

theoretical strength.

The fibers are usually available in the market as tows and yarns. A tow is a collection of

filaments laid out straight. Tows are usually specified by the number of filaments which

varies from 500 to 150,000. Yarn is almost the same as tow, and the only difference is that

yarn is slightly twisted into a more independent strand. Tows and yarns are laid on a backing

of continuous fiber sheets. This backing may be a randomly oriented scrim layer of a lower

modulus material, or a polymer which impregnates the area between the fibers. The latter

product is called a prepreg and requires special storage to keep the matrix from curing

prematurely. Generally the fibers are available in rolls, and these are usually separated by
91

paper backing sheet. Fabrics are made of yarns mechanically connected into a geometric

form. The wide variety of fabrics available in roll form is limited only by market demand and

includes unidirectional, layer stitched, and woven fabrics. Some fabrics have all layers of

fibers in one direction, and they are called unidirectional fabrics. The width of these fibers

varies from a few inches all the way to several feet. These fibers are usually named by their

weight per unit area and are stitched together with glass or polyester fibers for handling.

Unidirectional fabrics may be stitched together in layers having different fiber contents and

orientations relative to the roll direction. In general, the cross plies are laid 90 or 45 to the

base fabric, but any angle is possible to have depending on application. Depending on

application, the amount of fiber, the direction of fiber, the type of fiber, and the angle of fiber

varies. Advantages include versatility in orientation when applied to flat or singly curved

surfaces, and there are usually thin and readily penetrated by the matrix polymer (fragment).

Woven fabrics are similar to cloth fabrics in that the fiber bundles go over and under the

individual cross plies. These are usually narrow to 0/90weaves, but the amounts of fiber and

the types of fiber varies. One of the disadvantages of these fibers being curved out of plane

away from their primary orientation, but they are easy to handle and readily conform to three

dimensional curved molds without buckling. Some fabrics are woven by running the bundles

under several cross plies and then over one to straighten the fibers as much as possible. If this

pattern is staggered across the roll, the fiber deviation is reduced, and some of the desirable

properties of a woven remain (Fiberite, 1992).


92

2.5.3. Glass Fiber

Glass fiber is a generic term applied to a group of fiber with wide range of chemical

composition. The major part of all these fiber is silica (50% to 60% SiO2). The most common

type of glass is called E-glass. E standing for electrical since this type of glass has an

extremely well electrical insulating properties. More than 90% of all glass produced is E-

Glass with only a very small fraction being used for electrical purposes. The other type of

glass is C-glass. C-glass has a higher percentage of SiO2 with a much better chemical

resistance than E-Glass.

Glass fibers have the advantages of low cost and high strength in tension. Glass has a high

ratio of strength to weight. These fibers have an elastic modulus of about 10500 ksi and a

tensile strength of approximately 500 ksi. The disadvantages of glass are low fatigue

resistance, poor abrasion resistance, a tendency to degrade in alkaline environment, the

modulus is about a third that of steel so it is not as useful for increasing stiffness, and they

perform poorly in compression (Fiberite, 1992).

2.5.4. Carbon Fiber

Carbon or graphite is an extremely light element that can exist in a variety of crystalline

forms. Carbon, like glass is a generic term applied to a family of products with a wide range

of properties. Carbon fibers have a high modulus of elasticity and high strength in both

tension and compression. Carbon has a high ratio of strength to weight, and better

environmental resistance than glass. Composed almost entirely of carbon atoms, they are

similar to diamonds in that they gain their strength from their molecular configuration.
93

However, carbon fiber molecules are arranged in long hexagonal layers rather than a three

dimensional matrix. One of the disadvantages of carbon is electrical conductivity; therefore,

carbon and steel rust in long term. The other disadvantage is the high cost of manufacturing

carbon.

Polyacrylonitrile (PAN), rayon fiber, or pitches are some of the precursors to manufacturing

Carbon fiber. These processes are spun into fibers if needed, and then Carbon fiber is

stretched and heated up to 1500C in an inert atmosphere where the non-carbon elements

volatilize and the carbon molecules assume their chained orientation. By finishing the

process, the fabric contains around 85% carbon. To improve the molecular orientation and

increase the carbon to 99%, graphite fibers are added and the mixture is reheated again up to

3000C (Agarwal, 1990).

2.5.5. Matrix Materials

Since adhesion between individual fibers is limited, something additional is needed to

transmit forces between fiber and the applied loads. This commonly takes the form of a

polymer or plastic matrix. Polymers are structurally much more complex than metals or

ceramics. Polymers have the advantages of low cost, ease of workability, and good resistance

to some environmental effects. However they have lower strength and modulus, lower

temperature, and ultraviolet light resistance. Polymer matrix materials are broadly classified

by their behavior into thermoset and thermoplastic resins.

Thermoplastic resins are weak in molecular interaction, but extremely strong in chemical

bonds. This group includes polyethylene and nylon. When a thermoplastic polymer is heated
94

above its glass transition temperature, the molecules deform randomly and the plastic melts.

It can then be reshaped before it cools and reused. This is generally a desirable quality for

recycling of materials.

Another type of resin is Themosets. Thermosets are illustrated by crosslinking and covalent

bonds between the molecules, once chemical curing takes place. After Thermosets are heated

above their glass transition temperature, they will break down, and their molecules will break

down and macroscopic properties will be changed. The advantages of Thermosets are they

are an extremely easy process and have extremely well chemical resistance. Examples of

thermosets include polyesters and epoxies which are the most commonly used polymer

matrix materials for FRPs. Themosets are usually made of two components, resin and

hardener. These two components are usually mixed by different ratios, and then the fibers are

saturated by them. The curing time varies depending on the temperature and type of resin and

hardener. (Agarwal, 1990).

2.5.6. Advantages

In the past few decades, Fiber Reinforced Polymer (FRP) composites have been developed

and utilized in civil engineering, construction, automotive, marine and aerospace industries,

just to name a few. Research and applications have shown that FRP composites are far more

efficient and highly advantageous for retrofitting techniques than conventional materials.

Some common applications are as follows:

1. Increasing the structural integrity and strength to withstand underestimated loads or

correcting design errors/omissions


95

2. Increasing the load bearing capacity

3. Compensating for lost materials due to deterioration

4. Eliminating premature failure due to inadequate detailing

5. Well environmental resistance

It has been observed and universally accepted that replacing conventional materials with FRP

composites eliminates many issues associated with strengthening concrete and steel

structures due to such properties as the high strength to weight ratio, resistance to corrosion,

excellent fatigue strength, versatility, and economics. However, there are issues that affect

the material properties of the FRP; primarily the problems associated with environmental

conditions, e.g. the effects of thermal loading on the effective lifespan and mechanical

properties of FRP, along with the FRP/substrate interfacial bond strength. Carbon fibers

have a high elastic modulus and an ultra-high tensile strength. The tensile strength of

individual fibers is in excess of 700 ksi. Composed almost entirely of carbon atoms, the

fibers are generally available as bundles of 500-150,000 filaments of approximately five

microns in diameter called yarn. These are then assembled directly into FRP products or

into intermediate forms such as continuous fiber sheets or fabrics. Continuous fiber sheets are

made of parallel yarns attached to a flexible backing tape for handling. Fabrics are made of

yarns stitched into a geometric form. The yarns may run unidirectional like the continuous

fiber sheets, or be woven at different angles into a fabric. Since there is no adhesion between

individual fibers, a polymer or resin matrix is used to transmit forces between the fibers.

Polymers, which include the epoxy used in this study, have the advantages of low cost, ease
96

of workability, and good resistance to environmental effects. With tensile of strength several

times that of steel, modulus of elasticity of more than 20,000 ksi, and exceptional durability,

CFC can be used very effectively in many areas of geotechnical engineering.

The durability of this type of CFC fabrics in aggressive environments has been studied. In

particular, accelerated aging tests were conducted in various chemical solutions with minimal

loss of mechanical properties after 20,000 hours of direct exposure. Samples of carbon fiber

composites were placed in various chemical solutions and tested under accelerated aging

conditions (Tavkolizadeh and Saadatmanesh, 2005 and Tannous, 1997). Eight different

environments were simulated for accelerated exposure. The selected environments were: 1)

water H2O at 25o C; 2) saturated Ca(OH)2 solution with pH of 12 at 25oC; 3) saturated

Ca(OH)2 solution with pH of 12 at 60oC; 4) HCI solution with pH of 3 at 25oC; 5) NaCl 3.5

percent by weight solution at 25oC; 6) NaCl+CaCl2 (2:1) 7 percent by weight solution at 25oC;

7) NaCl+MgCl2 (2:1) 7 percent by weight solution at 25oC; and 8) ultraviolet radiation at

30.0 x 10-6 J/sec/cm2 .

Typically, seawater contains about 3.5 percent soluble salts by weight. The average weight of

the constituent salts are NaCl (27 g/l), MgCl2 (3.2 g/l), MgSO4 (2.2 g/l), and CaSO4 (1.3 g/l),

and KCI (0.2 g/l). Since NaCl has the highest percentage by weight (80 percent), seawater

was prepared by dissolving 35 grams of sodium chloride in 500 grams of distilled water, then

water, and then was added until the total weight of the solution was 1000 grams. The

resulting solution was 3.5 percent by weight of soluble NaCl with a chloride concentration of

0.6 mol/l.
97

Figure 2-32 shows results of exposure tests of the fabric to the various aggressive

environments simulated through the chemical solutions described above. In each figure two

curves are shown: the solid line indicates the behavior for glass reinforced fabric and the

dotted line shows the behavior for the carbon fiber reinforced fabric. As can be seen, there is

very little change in the properties (ultimate strength) of carbon fiber fabric after 20,000 hrs

of exposure, while glass based fabric shows significant losses. This is a testimony to the

superior long-term durability of carbon fiber reinforced membranes in highly aggressive

environment. The results for glass fiber reinforced membrane were provided here for

comparison, since this type of fabric has been used for certain industrial applications, but

because of its durability problems is not recommended for application in the soil environment.
98
99

Figure 2-32 FRP to the various aggressive environments (Tavkolizadeh and Saadatmanesh, 2005)
100

Ultraviolet and Microorganism exposures: In addition to the environments described above,

the materials were also tested under UV exposure and moist soil with microorganisms. For

the UV radiation, a series of 15 watt black light fluorescent tubes were used to simulate

ultraviolet radiation between 300 and 400 nm with a peak of 340 nm and intensity of 30 x 10-

6 J/sec/cm2 . Tubes and specimens were 25 cm apart from each other. After exposure, very

little change was observed in the mechanical properties of samples tested. In fact, some

samples showed a slight increase in the mechanical properties. The UV exposure seemed to

have helped curing of the carbon polymer fabric and resulted in improved properties. The

microorganisms and moist soil appear to have no impact on the mechanical properties of the

composite membrane during and after the 20,000 hrs of exposure.

2.6. Interface Behavior

2.6.1. Introduction

One of the most important characteristics in soil reinforcement is understanding that

design is associated with the shear behavior of the interface between reinforcement layers

and soil. In many designs it is suggested to use coarse grained soils for backfills (Elias

et al., 2001 and AASHTO, 2002). For the design of a drainage system, some current

industry guidelines such as NCMA (2002) allows the use of fine grained soil up to 35%.

Other design codes, such as the British Standard (BS8006, 1995), allows cohesive-

frictional soils, which usually is soil with greater than 15% passing 63 m sieve, to use as

a backfill.
101

Other guidelines for building reinforced embankments and slopes allow the use of up to

50% fine grained soils (passing sieve #200) for backfills (Elias et al., 2001). However,

due to the lack and high cost of better quality backfill soil in some areas, lower quality

backfill soils have been used in constructing slopes and embankments (Powel et al., 1999

and Musser and Denning, 2005). Using low amounts of fine graded soil (it can be low as

10%) can decrease the permeability of soils significantly (BS8006, 1995, Elias et al.,

2001 and Koerner, 2005). Often times these structures are constructed on unsaturated soil.

As soon as the degree of saturation increases, it creates a decrease in the soil

reinforcement interface shear strength. A rise in degree of saturation can occur by

variation of the ground water table and seasonal precipitation, and these occurrences

create changes in the soil properties such as soil moisture condition, suction, and thus the

interface behavior. In many instances backfill soils are exposed to heavy rainfalls or

compacted to reach wet optimum, and the failure acquired due to increase of pore water

pressure and decrease in matric suction. (Mitchell and Zornberg, 1995, Christopher et al.,

1998, Koerner, 2005, Sandri, 2005, Lawson, 2005 and Stulgis, 2005). Matric suction in

the soil is pore air pressure subtract to pore water pressure (Fredlund and Rahardjo, 1993

and Lu and Likos, 2004).

Pullout tests (ASTM D6706, 2007) and interface shear teats (ASTM D5321, 2008) are

two laboratory tests to find the soil and reinforcement interface strength on soil

reinforcement specimens. Based on the standard proctor test AASHTO T-99, in order to
102

simulate field conditions during construction, the soil sample needs to compact at

optimum moisture content and 95% of maximum dry density.

As previously mentioned, interface behavior for two different materials is very important

for many geotechnical engineering problems such as the interface between soil and pile

foundation and reinforcement and soil. Desai (1981) describes the effect of interface

behavior between soil and structures in a building foundation system. The location

between two different materials which relative deformation may transpire is known as

interface. There is more than one type of material used and the deformation would be

more complicated than simple frictional sliding. Four possibilities of deformation at an

interface can transpire according to Desai (1984), Desai and Nagaraj (1988), and they are

as follows:

a. No slip: When there is no movement under shear stress and the normal stress

stays compressed

b. Slip: When shear stress passes a failure stress and the normal stress remain

constant

c. Debonding: when shear stress and normal stress are zero or tensile for the

debonded part

d. Rebonding: during rebonding the interface returns to the bonded state with

nonzero values of normal stress and shear stress.


103

Figure 2-33 Modes of deformations at interface (Desai et al., 1984).

The friction solid by theory of adhesive friction was described by Bowden and Tabor in

1950, and the method was true for the friction between solid surface and sand.

Afterwards, the interface friction in a form similar to Mohr Coulomb theory was

expressed by Potyondy in 1961. Later on, the sliding behavior of structures was studied

by Fujino et al. (1987), Watanabe & Tochigi (1986), and Zhuan-zhi (2003). The radial

displacement of retaining walls was explored by Uwabe (1987), and Ahmadabadi &

Ghanbari (2009). Finally, the properties of the interface zone between the soil and wall

influences the earth pressure acting on the retaining wall as studied by Nakai (1985), El-

Naggar & Kennedy (1997) and Villemus Morel & Boutin (2007).
104

2.6.2. Pullout Test

When a pullout test takes place, both halves of the shear box are restrained, and a small

gap is left between their interfacing surfaces so that the reinforcement placed between the

two halves can be pulled freely. The top and bottom halves may contain the same soil

type or two different soils. A normal load (N) is applied on the top soil surface, and the

reinforcement is pulled at a constant rate until failure transpires. The force (F) required to

pull the sample is recorded versus horizontal movement. The test is repeated under

different normal loads representing different confining pressures. Typically, more reliable

design values are obtained from the pullout test than the direct shear test.

Pullout tests are used to mimic the behavior of reinforcement in soil to obtain the

interface properties of reinforced soil, and use these characteristics for designing such as

anchorage strength of reinforcements (Figure 2-34). There are other laboratory tests by

which the interface properties can be found such as direct shear test (Nataraj et al., 1995)

which will be explored in the following section. However, pullout tests are not a

standard test to calculate the interface properties including boundary effects, variation in

testing procedure, soil placement, compaction schemes, (Juran et al., 1988) complex

geometry of such materials and influence of effects such as soil dilation and

reinforcement loadstraintime characteristics.


105

Figure 2-34 Interaction mechanisms of reinforced soil wall (modified from Palmeira and Milligan,
1989a).

Palmeira (1987), Abramento (1993), Perkins and Cuelho (1999), and Sugimoto (2001),

performed pullout tests with different setups and procedures which change the typical

boundary conditions of the pullout box, and they are follow as (Figure 2-35):

a. The rigid front face.

b. Minimizing the friction along the front face by using oil, plastic films, grease,

((Palmeira, 1987; Abramento, 1993) or use of sleeves (Perkins and Cuelho, 1999)

c. The effect length of reinforcement to be a distance from the front wall (Palmeira,

1987).

d. Using flexible or movable frontal face (Sugimoto et al., 2001)


106

Figure 2-35 Boundary of pullout tests.

The experiments are limited due to issues of simulating the job site, cost and time

consuming, numerical methods, such as finite element method (FEM) and finite

difference method (FDM), are used to perform larger scale complex problems, and

improve some factors that may influence test results. One of the factors that can affect

the results from Pullout tests is the size of the apparatus. These affects to the results can

be substantial. Dais (2003) used the finite element method to explore the responses of

reinforced soil in pullout test by varying the size of the box. The box was varied from 0.3

m to 1 m. The length of the box and the length of the reinforcements remained constant

at 2 m and 0.5 m, respectively. The reinforcements were located at mid height, and they

simulated as linear elastic material. The soil, on the other hand, was modeled as an

elastic-plastic material. In his modeling, he assumed a rigid front face with 6 interface

friction angle as well as along the other internal faces of the apparatus. Based on the test
107

results, the lowest box height produced a stiffer pull-out response and a higher maximum

pull-out load. Slight influence of the box height was observed for heights greater than the

reinforcement length.

A pullout test is challenging to perform, since there are less controlled operational

conditions in the field in contrast to those in the laboratory. Clamps for the reinforcement

and reaction for the applied pull-out load have to be designed carefully. In tests with

actual reinforced structures, he potential influence of the conditions of the wall frontal

face on the test results should be studied. Tests with the reinforcement buried in

embankments can be performed; nevertheless, caution should be taken in the

interpretation of the results from these types of tests because of the usually low stress

levels on the reinforcement required for the test to be practical. Also, different and

sometimes impractical failure mechanisms may transpire, depending on the thickness of

soil above the reinforced layer, uniformity of fill material properties and boundary

conditions.

2.6.3. Direct shear test

During a direct shear test, the bottom half of the box is usually fully restrained against

movement while the top half of the box can slide over the bottom half, or vice versa. In

the meantime, a normal load (N) is applied uniformly on the surface of the soil. The soil

is then sheared to failure while horizontal and vertical forces and movements are
108

recorded. This arrangement yields a shearing force (F) distributed over the horizontal

area of the box. The test is completed when the horizontal force remains constant or starts

to drop. Numerous tests are performed under different normal loads to determine the

ultimate shear strength and the angle of internal friction () of the soil. Direct shear tests

are performed in combination with pullout tests since soil/reinforcement interface

properties depend on the shear strength of the soil.

The properties of the interface between soil and reinforcement are typically determined

by using direct shear tests. There are other tests available to define the properties of soil

and reinforcement such as tilt table test (Wu, 2008) or the cyclic multi degree of freedom

(CYMDOF) (Desai, 1981). However, direct shear test is one of the most popular tests to

define the properties of soil reinforcement. There is a large amount of research available

that studies the interface shear strength of soil with other type reinforcement such as

geogrids, tire shreds, rubber chips, geofoam (Bergado (2006), and Palmeira (2009). A

Mohr-Coulomb model is often used to simulate the interface behavior of linear elastic

materials. However, most of the data from a direct shear test shows a nonlinear

relationship between force and displacement. Therefore, it is better to use nonlinear

modeling.

2.7. Different Models

MohrCoulomb theory is a mathematical model describing the response of brittle

materials such as concrete, or rubble piles, to shear stress as well as normal stress. Most
109

of the traditional engineering materials somehow follow this rule in at least a portion of

their shear failure envelope. Normally the theory applies to materials in which the

compressive strength far surpasses the tensile strength. In geotechnical engineering, it is

used to define shear strength of soils and rocks at different effective stresses. In

structural engineering it is used to determine failure load as well as the angle of fracture

of a displacement fracture in concrete and similar materials. Coulomb's friction

hypothesis is used to define the combination of shear and normal stress that will cause a

fracture of the material. Mohr's circle is used to define which principal stresses will

produce this combination of shear and normal stress, and the angle of the plane in which

this will transpire. According to the principle of normality the stress introduced at failure

will be perpendicular to the line describing the fracture condition.

A material failing according to Coulomb's friction hypothesis will show the displacement

presented at failure forming an angle to the line of fracture equal to the angle of friction.

This makes the strength of the material determinable by comparing the external

mechanical work introduced by the displacement and the external load with the internal

mechanical work introduced by the strain and stress at the line of failure. The Mohr

Coulomb failure criterion characterizes the linear envelope that is obtained from a plot of

the shear strength of a material versus the applied normal stress. This relation is

expressed as:

(2-2)
110

where is the shear strength, is the normal stress, c is the intercept of the failure

envelope with the axis, and is the slope of the failure envelope. The parameter c is

often termed the cohesion and the angle is called the angle of internal friction.

Compression is assumed to be positive in the following discussion. If compression is

assumed to be negative then should be replaced with .

Formulations relate to the assumptions of perfect plasticity as to those of a hardening

material, permitting also the analysis of particular conditions such as cyclic loading and

softening. Along with a new definition of the kinematic state variables, stresses are

related to displacements, and the theoretical structure of these models remains

indistinguishable to that of the bulk of the elasto-plastic models recommended for the

description of soil behavior (De Gennaro and Frank (2002)).

This constitutive relationship can be simply implemented in finite elements programs and

then used for analysis of the behavior structures composed of multilayered materials

presenting slipping risk at the constituents interface. The interface can represented by

{d} = [C] {d} for nonlinear elastic model, which {d} is the elastic incremental stress,

{d} is the elastic incremental strain and [C] is the elastic constitutive. For most elasto-

plastic models, it is presumed that the material is plastic up to a yielding point and then

plastic.
111

Summary

This chapter briefly discusses different types of soils, soil stabilization and different types

of soil reinforcement. The various tests to examine the interface relationship between

soils and reinforcement are also investigated. Different types of modeling techniques are

also briefly explained. An introduction and the various characteristics of fiber reinforced

polymer (FRP) are also presented. The following chapters discuss some experimental,

theoretical and numerical methods for the study and evaluation of the interaction between

soils and Geo-Composites, with particular reference to the applications of these materials

in soil reinforcement.
112

CHAPTER 3. LABORATORY RESULTS

In this chapter the results from the laboratory tests are described. Figure 2-34 shows

several possible failure or deformation mechanisms of a reinforced soil wall depending

on the region and loading conditions considered. In region A (Figure 2-34), sliding of the

soil mass on the reinforcement surface can occur. Direct shear tests can be used to

measure the soilreinforcement bond under these conditions. In region B, soil and

reinforcement can deform horizontally. A plane strain test similar to the in-soil tensile

test can be used in this case. Region C shows a state where soil and reinforcement are

sheared. The direct shear test with the reinforcement inclined to the shear plane can be

used. In region D, the reinforcement is being pulled-out, and pull-out tests would be

relevant in this situation. It should be noted that all these tests have limitations in

simulating the actual conditions found in a reinforced soil structure (Palmeira and

Milligan, 1989a; Mendes et al., 2007).

The interaction between soils and geosynthetics has been modeled theoretically and

experimentally in numerous different ways for the last three decades. The main objective

of this chapter is to discuss experimental tools for the understanding and evaluation of

soil-geosynthetic. This discussion is appropriate to most applications of geosynthetics.


113

3.1. Direct Shear Test

Direct shear test is a laboratory test, mostly conducted by geotechnical engineers to find

the shear strength parameters of soil such as friction angle. Soil shear strength parameters

are essential for stability because every physical structure imposes some load on the soil

that supports the foundations. In a direct shear test, the failure of the soil sample in shear

is caused along a pre-determined plane. In this test, the normal load, strain and shearing

force are measured directly.

3.1.1. Direct shear machine

In this research, the results are obtained by using direct shear machine of Soil Test

D120A8 as shown in Figure 2-34. The shape of inner box is square with dimension of

5x5 cm. The shear testing apparatus/machine manufacturer provides strain controlled on

the horizontal axis, and the rate of the displacement can be changed by using different

gears that are provided with the direct shear machine. The normal force is applied by

using different weights that are applied at the top of the box on the sample. The Soil

Test D120A8 machine works manually; the data is collected by humans and to minimize

human error all the data such as tangent displacement and tangent load are recorded by

video camera, and then imported into excel for further analyses.
114

Figure 3-1 Direct shear machine.

3.1.2. Calibration direct shear test

The tangent load defines by deformation on a ring (Model: SoilTest 15002), and the

deformation on ring is measured by a gage (Model: SoilTest LC-2) with accuracy of

0.0001 as shown in Figure 3-1. For calibration of the ring, different weights applied on

the ring, and the corresponding deformations for each weight are collected. Linear

relation is assumed between each weight; therefore, during a test, by reading the

deformation of the ring, the weight (or the load) can be calculated. Tangent deformation

is measured by using a gage (Model: SoilTest LC-8) with accuracy of 0.001, which the

gage is also calibrated.

A particular soil with recognized friction angle is tested with direct shear machine of

Soil Test D120A8, and based on the results obtained from this test, the results were
115

higher than the actual values; therefore, the direct shear box is modified. To reduce the

friction between the two steel plates when the two inner boxes are placed on top of each

other and also adjusting the height of CFRP with soil, three steel sphere balls with

diameter of 5 mm are used between the two steel boxes as shown in Figure 3-2.

Figure 3-2 Using steel balls on direct shear boxes.

Figure 3-3 and Figure 3-4 show the result of direct shear test on loose sand with and

without using steel ball, respectively. Based on the result, as the normal force is

increasing, there is a larger difference between the shear stresses. The friction angle for

each results are calculated and there is a 5.37 degree reduction between the old shear box

and the modified shear box as shown Figure 3-5.


116

OldShearBox
120 125.6KPa

100 62.8KPa

80
Shearstress(kPa)

31.4KPa

60 15.7KPa

40 7.87KPa

20
0KPa

0
0 1 2 3
Sheardisplacement(mm)

Figure 3-3 Results on direct shear test on old-shear box (soil-soil).

NewShearBox
120
125.6KPa

100
62.8KPa

80
Shearstress(kPa)

31.4KPa

60
15.7KPa

40
7.87KPa

20
0KPa
0
0 1 2 3
Sheardisplacement(mm)

Figure 3-4 Results on direct shear test on new shear box (soil-soil).
117

120

100
=0.797
80
(kPa)

oldshearbox
60
newshearbox

40

20
=0.654
0
0 50 100 150
(kPa)

Figure 3-5 Friction angle for two different shear boxes.

3.1.3. Soil

Jupar sand is used in this study. This is a native sand from Jupar roadway of Kerman

province situated in the southeast part of Iran. Jupar sand is abundantly available in this

region. Visual classification showed the Jupar sand is subround to round. This soil

sample based on Unified Soil Classification System (USCS) is classified as poorly-

graded sand (SP), and ASTM D422 is used for the procedure of this lab. Figure 3-6

shows the graph of the particle size of the soil using sieve analysis for coarse portion of

the soil and hydrometer analysis for fine portion of the soil. A compaction test on this soil

is accomplished based on ASTM D698. Compaction curve for this soil is shown in

Figure 3-7, and the index properties of the soil are presented in Table 3-1.

The DST results presented in Figure 3-8 show the shear stress-shear strain behavior for

Jupar sand over a range of normal loads (ASTM D-3080 (2004)).


118

ParticleSizeDistributionCurve

100
90
80
PercentofPassing

Hydrometer
70
Test
60 SieveTest
50
40
30
20
10
0
1 10 100 1000
GrainSize(mm)

Figure 3-6 Particle size analysis.

CompactionCurve

16.2
DryDensity(kN/m)

16.1
16
15.9
15.8
15.7
15.6
15.5
8 10 12 14 16 18
WaterContent(%)

Figure 3-7 Compaction curve for sand soil.


119

Water content (%) 0%

Specific Gravity (Gs) 2.64

Maximum void ratio (emax) 0.64

Minimum void ratio (emin) 0.60

Coefficient of Curvature (CC) 1

Uniformity Coefficient (UC) 3.4

Friction angle () 35

Cohesion (c) 0.0

Table 3-1 Index properties of investigated soil.

120

100
n= 125.6 kPa
ShearStress, (kPa)

80

60

n= 62.8 kPa
40
n= 31.4 kPa

20 n= 15.7kPa
n= 7.9kPa

0
0 1 2 3 4 5 6 7 8 9 10
HorizontalStrain, (%)

Figure 3-8 Shear Stress vs. Horizontal Displacement.

3.1.4. Fiber Reinforcement Polymer

Seven samples of Carbon Fiber Reinforcement Polymer (CFRP) were tested in tension to

obtain the properties of CFRP. Samples were one inch wide with a 30 cm in length, and
120

they were tested in tension with a displacement rate of 1.3 mm/min base on ASTM

D3039. The data collection rate is one point every second. Figure 3-9 shows the stress

vs. strain curve of one the samples out the seven samples. The average and the standard

deviation of the maximum stress as well as then corresponding strain, modulus of

elasticity and load per width of these seven samples are shown in Table 3-2 below.

500

450

400

350
Tension Stress (MPa)

300

250

200

150

100

50

0
0 0.2 0.4 0.6 0.8 1 1.2
Tension Strain, (%)

Figure 3-9 One layer of carbon fiber.

AVG

Max
Stress 426.5 MPa

Co Strain 0.013867 mm/mm

Load/W 403 KN/mm

E 42.4 GPa
121

Stdev

Max
Stress 46.7 MPa

Co Strain 0.008064 mm/mm

Load/W 39.3 KN/mm

E 5.93 GPa

Table 3-2 One layer of carbon fiber.

Figure 3-10 Failure of CFRP.

3.1.5. Preparation

One sheet of unsaturated carbon fiber is cut with dimensions of 5 cm by 5 cm as the

dimension of the inner box on direct shear machine. The fiber sheet is saturated with

epoxy resin which is formed from a 2:1 ratio of resin and hardener. Resin and hardener

are measured by volume using a measuring cup. They are then mixed using an electric
122

drill for three minutes. Then the epoxy resin is applied to one side of the fiber with a

squeegee. The squeegee is moved in a direction parallel to the fibers. After saturating one

side, the fiber sheet is flipped over and the same process is applied as shown in Figure

3-11.

Figure 3-11 Saturated carbon fiber.

Then, saturated carbon fiber is placed on top of a stone block with cross sections of 5x5

cm. After that, the block with fiber is placed in a mold so that soil can be placed on top

of the fiber with an initial normal force. It takes about 48 hrs for curing; however,

depending on temperature, the curing can vary. Then, the sample is ready to be placed in

the direct shear box (Figure 3-12). However, the height of the block with CFRP is

adjusted in order to place it exactly between the two shear boxes (Figure 3-13).
123

Figure 3-12 Block and CFRP.

Figure 3-13 Schematic diagram of modified shear box, used in experiments.

3.1.6. Effects of Initial normal force

In this part, one side of CFRP is cured on top of the block and the other side of CFRP is

cured with different height of sand on top of CFRP until the CFRP is cured. The heights

of the soil on top of the CFRP are varied, such as 0 cm, 10 cm, 20 cm, 30cm, 40cm, 50cm,
124

and 100cm, to determinate the possibility of increasing friction angle between CFRP and

soil. Then, the block is placed in direct shear box, and three different normal stresses as

0.5 kg/cm, 1.0 kg/cm, and 2.0 kg/cm (49 kPa, 98 kPa, and 196 kPa) are applied on the

sample. The friction angle between CFRP without any soil on top and soil is

measured by using the three different normal stresses.

Figure 3-14 shows results of direct shear test with an initial 0 cm soil on top of the CFRP

(i= 0 kPa); however, it is close to zero, because the surface of the CFRP is covered with

a small portion of the soil. As it was mentioned before, three different normal stresses

were applied on top of the sample, and the peak tangent stresses with corresponding

normal stress are graphed as shown in Figure 3-15. Then, the best linear curve passing

through the three data sets and the friction angle is calculated as 28.88 degrees.

PrecastedCFRPSandi=0kPa(h=0m)

120

100
ShearStress(kPa)

80
196KPa
60 98KPa
49KPa
40

20

0
0 1 2 3 4 5 6
Sheardisplacement(mm)

Figure 3-14 Direct shear test with 0 m compaction depth.


125

PrecastedCFRPSandi=0kPa(h=0m)
120

100
MaxShearStress(kPa)

80

60

40

20 =0.5515n

0
0 50 100 150 200 250
NormalStress(kPa)

Figure 3-15 Variation of maximum shear stress vs. normal stress for 0 m compaction depth.

Based on Figure 3-16, the friction angle between the soil and CFRP with initial 10 cm

soil embankment (i= 1.6 kPa) is 35.17 degrees.

PrecastedCFRPSandi=1.6kPa(h=0.1m)

160
140
120
ShearStress(kPa)

100 196KPa
80 98KPa
60 49KPa
40
20
0
0 1 2 3 4 5 6
Sheardisplacement(mm)

Figure 3-16 Direct shear test with 0.1 m compaction depth.


126

PrecastedCFRPSandi=1.6kPa(h=0.1m)
160
140
MaxShearStress(kPa)

120
100
80
60
=0.7046n
40
20
0
0 50 100 150 200 250
NormalStress(kPa)

Figure 3-17 Variation of maximum shear stress vs. normal stress for 0.1 m compaction depth.

Figure 3-18 shows the friction angle between the soil and CFRP with initial 20 cm soil

embankment (i= 3.22 kPa) is 37.7 degrees.

PrecastedCFRPSandi=3.2kPa(h=0.2m)

180
160
140
ShearStress(kPa)

120
196KPa
100
98KPa
80
49KPa
60
40
20
0
0 1 2 3 4 5 6
Sheardisplacement(mm)

Figure 3-18 Direct shear test with 0.2 m compaction depth.


127

PrecastedCFRPSandi=3.22kPa(h=0.2m)
160
140
MaxShearStress(kPa)

120
100
80
60 =0.7729n

40
20
0
0 50 100 150 200 250
NormalStress(kPa)

Figure 3-19 Variation of maximum shear stress vs. normal stress for 0.2 m compaction depth.

Figure 3-20 shows the friction angle between the soil and CFRP with initial 30 cm soil

embankment (i= 4.83 kPa) is 38.68 degrees.

PrecastedCFRPSandi=4.83kPa(h=0.3m)

180
160
140
ShearStress(kPa)

120
100 196KPa
98KPa
80
49KPa
60
40
20
0
0 1 2 3 4 5 6
Sheardisplacement(mm)

Figure 3-20 Direct shear test with 0.3 m compaction depth.


128

PrecastedCFRPSandi=4.8kPa(h=0.3m)
180
160
MaxShearStress(kPa)

140
120
100
80
60
=0.8005n
40
20
0
0 50 100 150 200 250
NormalStress(kPa)

Figure 3-21 Variation of maximum shear stress vs. normal stress for 0.3 m compaction depth.

Figure 3-22 shows the friction angle between the soil and CFRP with initial 40 cm soil

embankment (i= 6.44 kPa) is 39 degrees.

PrecastedCFRPSandi=6.44kPa(h=0.4m)

160
140
120
ShearStress(kPa)

100 196KPa
80 98KPa
60 49KPa
40
20
0
0 1 2 3 4 5 6
Sheardisplacement(mm)

Figure 3-22 Direct shear test with 0.4 m compaction depth.


129

PrecastedCFRPSandi=6.44kPa(h=0.4m)
180
160
MaxShearStress(kPa)

140
120
100
80
=0.8098n
60
40
20
0
0 50 100 150 200 250
NormalStress(kPa)

Figure 3-23 Variation of maximum shear stress vs. normal stress for 0.4 m compaction depth.

Figure 3-24 shows the friction angle between the soil and CFRP with initial 50 cm soil

embankment (i= 8.05 kPa) is 39.25 degrees.

PrecastedCFRPSandi=8.05kPa(h=0.5m)

180
160
140
ShearStress(kPa)

120
196KPa
100
98KPa
80
49KPa
60
40
20
0
0 1 2 3 4 5 6
Sheardisplacement(mm)

Figure 3-24 Direct shear test with 0.5 m compaction depth.


130

PrecastedCFRPSandi=8.05kPa(h=0.5m)
180
160
MaxShearStress(kPa)

140
120
100
80 =0.8098n
60
40
20
0
0 50 100 150 200 250
NormalStress(kPa)

Figure 3-25 Variation of maximum shear stress vs. normal stress for 0.5 m compaction depth.

Figure 3-26 shows the friction angle between the soil and CFRP with initial 100 cm soil

embankment (i= 16.1 kPa) is 39.25 degrees.

PrecastedCFRPSandi=16.1kPa(h=1.0m)

160
140
120
ShearStress(kPa)

100 196KPa
80 98KPa
60 49KPa
40
20
0
0 1 2 3 4 5 6
Sheardisplacement(mm)

Figure 3-26 Direct shear test with 1 m compaction depth.


131

PrecastedCFRPSandi=16.1kPa(h=1.0m)
180
160
MaxShearStress(kPa)

140
120
100
80 =0.8171n
60
40
20
0
0 50 100 150 200 250
NormalStress(kPa)

Figure 3-27 Variation of maximum shear stress vs. normal stress for 1 m compaction depth.

From Figure 3-14 to Figure 3-27, it can be concluded that as the initial stresses increase

the friction angle until it reaches the maximum friction angle without considering

adhesion, and it can also be concluded that the minimum height to obtain the maximum

friction angle is 50 cm. Figure 3-28 shows the summary of the results obtained for

different heights of soil on top of CFRP.


132

Friction Angle vs. Compaction Depth w/o


Adhesion
41
39
FrictionAngle(degree)

37 FrictionAngleb/w
35 Soil&CFRP
33 SoilFrictionAngle
31
29 FrictionAngleb/w
27 Soil&FRPSand
25
0 0.2 0.4 0.6 0.8 1 1.2
CompactionDepth(m)

Figure 3-28 Friction angle vs. compaction depth.

3.1.7. Friction angle between soil and CFRP

In this part, the friction angle between the soil and CFRP without any soil sticking to

CFRP is calculated by using three different normal stresses such as 49 kPa, 98 kPa, and

196 kPa. Figure 3-29 shows the result of the direct shear test between CFRP and sand,

and based on the results, the friction angle is calculated as 27.21 degrees.
133

120

100
ShearStress(kPa)

80
196KPa
60 98KPa

40 49KPa

20

0
0 2 4 6
ShearDisplacement(mm)

Figure 3-29 Direct shear test between CFRP and Soil.

120

100
MaxShearStress(kPa)

80
=0.5142n
60

40

20

0
0 50 100 150 200 250
NormalStress(kPa)

Figure 3-30 Variation of maximum shear stress vs. normal stress for CFRP and sand.

Figure 3-31 to Figure 3-33 compare the different behavior of reinforced soil to the sample

that was cast in place CFRP cured with 100 cm soil on top and the other one with no soil
134

sticking to it. Figure 3-28 summarizes the results of the friction angle with different

initial stress. It can be concluded that existing soil on CFRP has a strong influence on

increasing friction angle.

50
45
40
35
CFRP&Soil
Shearstress(kPa)

30
25
CFRPSand&Soil
20
15
10
5
0
0 1 2 3 4 5 6
Sheardisplacement(mm)

Figure 3-31 CFRP w/ soil & CFRP w/o soil with normal stress of 49 kPa.
135

80
70
60
50 CFRP&Soil
Shearstress(kPa)

40
CFRPSand&Soil
30
20
10
0
0 1 2 3 4 5 6
Sheardisplacement(mm)

Figure 3-32 CFRP w/ soil & CFRP w/o soil with normal stress of 98 kPa

180
160
140
Shearstress(kPa)

120
100
80 CFRP&Soil

60 CFRPSand&Soil

40
20
0
0 1 2 3 4 5 6
Sheardisplacement(mm)

Figure 3-33 CFRP w/ soil & CFRP w/o soil with normal stress of 196 kPa.
136

3.1.8. Precast preparation and results

In this part, the different behavior of soil reinforcement for cast in place and precast

is investigated, and the results are compared. All the results obtained in previous labs are

categorized as precast CFRP. To simulate cast in place, The block is placed at the

bottom of the direct shear box, and the saturated fiber carbon is placed on top of it. Then,

the soil is placed on top of the CFRP, and a normal load which is equal to 100 cm of soil

is applied on top of the sample for 48 hrs. After removing the initial load, the three

different normal loads (such as 0.5 kg/m, 1.0 kg/m, and 2.0 kg/m) are applied to obtain

the friction angle. Figure 3-34 shows the result of the three tests, and the calculation of

the friction angle which is 39.38 degrees. Figure 3-36 to Figure 3-38 compares the

behavior of the two types of soil reinforcements which are CFRP Precast and CFRP

Cast in Place. Both types use 100 cm of soil placed on top of the soil during curing of

epoxy resin. It can be concluded that there wouldnt be a large difference between the

results. Figure 3-39 shows the ratio of the max shear strength to normal strength for cast-

in-place & precast CFRP-Sand.


137

180
160
140
ShearStress(kPa)

120
100 49KPa
80 98KPa
60
196KPa
40
20
0
0 2 4 6
Sheardisplacement(mm)

Figure 3-34 Soil & cast in place CFRP w/ 100 cm soil in top i=16.1 kPa.

180
160
140
MaxShearstress(kPa)

120
100
80 =0.8208n
60
40
20
0
0 50 100 150 200 250
Normalstress(kPa)

Figure 3-35 Variation of maximum shear stress vs. normal stress on Sand & cast in place CFRP w/
100 cm soil in top i=16.1 kPa.
138

PrecastFRP&CastinPlaceFRP(100cmsoilintop)
NormalStress49kPa
50
45
40
Shearstress(kPa)

35
30 castinplaced
25
20 precasted
15
10
5
0
0 2 4 6
Sheardisplacement(mm)

Figure 3-36 Precast CFRP & cast in placed CFRP with normal stress of 49 kPa.

PrecastFRP&CastinPlaceFRP(100cmsoilintop)
NormalStress98kPa
90
80
70
Shearstress(kPa)

60
50 castinplaced
40 precasted
30
20
10
0
0 2 4 6
Sheardisplacement(mm)

Figure 3-37 Precast CFRP & cast in placed CFRP with normal stress of 98 kPa.
139

PrecastFRP&CastinPlaceFRP(100cmsoilintop)
NormalStress196kPa
180
160
140
Shearstress(kPa)

120
100 castinplaced
80 precasted
60
40
20
0
0 2 4 6
Sheardisplacement(mm)

Figure 3-38 Precast CFRP & cast in placed CFRP with normal stress of 196 kPa.

0.95
NormalStesstoShearStressRatio

0.9

0.85
Cast in Placed
0.8 Precasted

Cast in Placed Resistance Manner


0.75
Precasted Resistance Manner

0.7
0 50 100 150 200 250
NormalStress(kPa)

Figure 3-39 Comparing ratio the of max shear strength to normal strength b/w cast in place &
precast.
140

3.1.9. Discussion

A summary on the direct shear tests is presented below:

It is a good test for determining interface shear strength.


Boundary conditions may influence test results, particularly for small shear boxes.
It is difficult to assess soilgeotextile interface shear stiffness, which is relevant
for numerical simulations of reinforced soil structures such as walls and steep
slopes.
The interpretation of tests with inclined reinforcements is very difficult.

Two specimen preparation methods of pre-casting and casting in place had almost no

significant influence on the measured interface friction angle. Sand-CFRP specimens

have a higher interface friction coefficient (twice larger) than FRP specimens, since the

FRPs have a smooth surface compared to Sand-CFRPs that have a rough surface. Due to

rougher surface of CFRP-Sand than CFRP and because of grains on the interface surface

are less likely to be rearranged during shear, CFRP-Sand interface behaviors have no

peak.

3.2. Pull out test

The pull-out test device was designed and assembled using a triaxial loading device and a

direct shear device. In the pull-out test, the normal force applied by the triaxial loading

and pull out force is applied by a direct shear device. CFRP samples were prepared in lab.

Precast and cast-in-place samples were tested. The pull-out force and corresponding

displacements of each of the materials were recorded and compared.


141

3.2.1. Test materials:

3.2.1.1. Sand

The same soil is used in section 3.1.3.

3.2.1.2. Epoxy

The same epoxy is used as section 3.1.5

3.2.1.3. CFRP

The same CFRP is used as section 3.1.4.

3.2.2. Test equipment

The Pull-out box was made of a steel rectangular shape of 25x25x25cm which had a hole

in the middle on one side in order to pull out the CFRP-Sand samples Figure 3-40. The

normal load was applied by a triaxial loading machine, and CFRP samples were pulled

by a direct shear machine as shown in Figure 3-41.


142

Figure 3-40 Pull out box.

Figure 3-41 Test apparatus for pullout test.


143

3.2.3. Test procedure

In this study, two methods are used to make the CFRP-Sand samples. The methods are

precast CFRP-Sand and cast-in-place CFRP-Sand. For a large scale project, CFRP-Sand

can be made in a precast factory, and CFRP-Sand can be prepared at job site which

would be synonymous to cast-in-place. The results of the pull-out test for both cases are

compared to define the most economical and efficient method.

3.2.3.1. Precast tests

Eight groups of three specimens with dimensions of 35x5cm were made in which 20cm

of the specimens were covered by sand as shown in Figure 3-42. The first group was

made of only CFRP and epoxy resin, and the other groups were prepared under seven

different normal loads during the curing period. The curing period is the time it takes the

epoxy resin to cure. The normal loads were equivalent to the height of the soil at the top

of the saturated CFRP as follows: 1, 10, 30, 50, 70, 100 and 150cm during the curing

time. The curing period for each group was 48 hours which is in accordance with the

manufacturer. The tests were carried out by pulling the CFRP-Sand samples out of dry

sand with a displacement rate of 1mm/min with three different normal stresses as follow,

25, 50 and 100 kPa.


144

Figure 3-42 Twenty-four CFRP-Sand specimens prepared for precast pull-out test.

3.2.3.2. Cast-in-place tests

In this section, CFRP-Sand samples were saturated, loaded, cured in the pull-out box and

then tested. The amount of load that was used during the curing period was based on the

optimum results obtained from the precast CFRP-Sand samples. However, the procedure

of making and testing the samples was the same as the precast samples.

3.2.4. Test results and discussion

3.2.4.1. Precast mode

The pull-out test results of precast CFRP-Sand specimens were shown in Figure 3-43,

Figure 3-44, and Figure 3-45. The results indicate that the lowest maximum pull-out

force was taken by the CFRP samples without sand attached to the samples. The

maximum pull-out force was obtained for the 100 and 150cm soil heights. Therefore, it

can be concluded that increasing the normal load during the curing period traps more soil
145

particles in the saturated CFRP and causes a rougher surface which in turn increases the

shear resistance force. However, when comparing the results of the 100cm and 150cm

soil heights, it can conclude that both results produced similar results. This indicates that

increasing the normal load for soil heights greater than 100cm during the curing period

does not have an effect on the result of the pull-out forces.

Figure 3-43 Pull out test on Precast CFRP-Sand with normal stress 25kPa.
146

Figure 3-44 Pull out test on Precast CFRP-Sand with normal stress 50kPa.
147

Figure 3-45 Pull out test on Precast CFRP-Sand with normal stress 100kPa.

Figure 3-46 shows the friction angle of CFRP-Sand versus the normal stress during the

curing period. Based on the results, it can be seen that the internal friction angle

increases almost linearly with the normal load during the curing period of the 100cm

sample. Eventually the friction angle ceases to increase and stays constant. In other
148

words, samples made under a normal load equal to 100 cm and 150cm soil height have

the same friction angle which shows that increasing the normal load during the curing

period time more than 100cm does not have any effect on the friction angle. The

maximum friction angle obtained from the 150cm soil height specimens is 11.25% higher

than the friction angle of the 50cm soil height specimens and 11.8% higher than friction

angle of soil sample. Based on the following results, the normal load equal to 100cm soil

height was chosen as the optimum burden for producing CFRP-Sand.

Figure 3-46 Friction angle of CFRP-Sand with soil.

3.2.4.2. Cast-in-place mode

According to the precast CFRP-Sand test results, the normal load equal to 100cm soil

height was chosen as the optimum burden for producing CFRP-Sand; therefore, six
149

samples were made using the cast-in-place method and tested. The results were shown in

Figure 3-47.

Cast-in-place under 100cm soil


140
120
Pull-out force (kg)

100
80
60
Cast in place (1) n = 100
40 kPa
Cast in place (2) n = 100
20 kPa
Cast in place (1) n = 50
0
kPa
0 1 2 3 4 5 6 7 8
Displacement of FRP (mm)

Figure 3-47 Cast-in-place CFRP-Sand specimens pull-out test results.

3.2.5. Discussion

For n=25 kPa, the cast-in-place samples were able to tolerate a maximum tensile force

twice as large as the precast samples as shown in Figure 3-48. When n=50 kPa, the

maximum tensile force in cast-in-place CFRP-Sand was 10% more than precast CFRP-

Sand and shown in Figure 3-49. Finally, when n=100 kPa, both samples in the two

modes tolerated the same tensile forces (Figure 3-50).


150

n = 25kPa
140
Precasted under 100cm
120 Soil
Casted in Place under
Pull-out force (kg)

100 100cm Soil


80
60
40
20
0
0 1 2 3 4 5 6 7 8
Displacement of CFRP (mm)

Figure 3-48 Comparison results of pull-out test for cast-in-place and pre-casted for height of 25 cm.

n = 50kPa
140
120
Pull-out force (kg)

100
80
60
Precasted under 100cm
40 Soil
20 Casted in Place under
100cm Soil
0
0 1 2 3 4 5 6 7 8
Displacement of CFRP (mm)

Figure 3-49 Comparison results of pull-out test for cast-in-place and pre-casted for height of 50cm.
151

n = 100kPa
140
120
Pull-out force (kg)

100
80
60 Casted in Place under
100cm Soil
40
Precasted under 100cm
20 Soil

0
0 1 2 3 4 5 6 7 8
Displacement of CFRP (mm)

Figure 3-50 Comparison results of pull-out test for cast-in-place and pre-casted for height of 100cm.

The comparison between the tolerated forces in the two modes indicates that in the lower

normal stresses, cast-in-place CFRP-Sand samples show higher pull out resistance force.

Cast-in-place CFRP-Sand Samples were made in the pull-out box and they became more

unified with the soil inside the box. The particles of CFRP and the soil had a stronger

bond in comparison to precast samples at lower normal stresses for preparation. However,

for both preparation techniques, the bond between the soil particles and CFRP during

curing time was the same at higher normal pressure.

The internal friction angle of the cast-in-place samples was 36.3 which is approximately

6.5% higher than the precast samples.

One of the disadvantages of CFRP is used in direct shear test and pull out test is the

brittle resin which is used for saturating the Carbon fibers. CFRPs using brittle resin are
152

breakable in small deformations; therefore, a more flexible resin is used in next section.

However, this type of test is usually difficult to perform, because of less controlled

operational conditions in the field in comparison to those in the laboratory. Clamps for

the reinforcement and reaction for the applied pull-out load have to be carefully designed.

In tests in real reinforced structures one has also to consider the possible influence of the

conditions of the wall frontal face (if any) on the test results. Tests with the reinforcement

buried in embankments can be performed, however, due care should be taken in the

interpretation of the results from these types of tests because of the usually low stress

levels on the geosynthetic required for the test to be practical. Besides, different and

sometimes unrealistic failure mechanisms may occur, depending on the thickness of soil

above the geosynthetic layer, uniformity of fill material properties and boundary

conditions.

3.3. CYMDOF Test

3.3.1. CYMDOF machine

The cyclic multi degree of freedom (CYMDOF), as shown in shear Figure 3-51, device

was originally designed and developed by Desai and was used by Desai and his

coworkers (Drumm 1983, Desai and Rigby 1997, Fishman 1988, and Desai et. al. 2005).

The present device which was conceived and redesigned over the period of ten years was

modified and fabricated by Desai and Rigby (1997). The modified shear device was built

with a provision for undrained interface testing.


153

The CYMDOF shear device is intended for the testing of interfaces such as those

between concrete and soil, steel and soil, joints in rock masses, and any other materials as

long as the top specimen is relatively rigid. The device is capable of simulating both

direct and simple shear modes of deformation. For direct shear testing, the two materials

are placed one above the other in the two halves of the box. For simple shear testing a

provision for allowing simple shear deformation is made by confining the geologic

specimen by annular smooth confining rings, which are coated with Teflon. Under

shearing action a thin interface zone develops between the two materials. Horizontal

displacements of this zone are measured by two pairs of small LVDTs (Linear Variable

Differential Transformers). This enables relative or slip displacements at the interface to

be determined.

A single hydraulic pump supplies pressure through hydraulic lines to any of three

separate testing stations (Figure 3-52). The oil pressure is used to operate hydraulic

actuators which generate the load or displacement needed for performing tests. At every

station a servovalve is connected to each actuator in order to regulate the oil flow that

controls the actuator. The servovalve, in turn, is controlled by MTS electronics.


154

Figure 3-51 Cyclic multi degree of freedom machine.

Figure 3-52 Hydraulic pump.


155

3.3.2. Soil

The backfill soil is collected by El-Hoseiny (1999) from Tanque Verde wash in Tucson

AZ in 1999 as shown in Figure 3-53. The properties of the soil is defined by El-Hoseiny

(1999); however, some of the properties such as classification of soil and maximum dry

density are tested in laboratory to identify the same soil is using as El-Hoseiny (1999).

Backfill Soil

Figure 3-53 Tanque Verde Wash.

The natural soil is available in constitutive modeling laboratory, and to be able to use it in

CYMDOF machine, the entire soil sample is passed through sieve number 4 (Figure

3-54). By following up the ASTM D421 and ASTM D422, the particle size distribution

of the soil is defined. Figure 3-55 shows the particle size distribution of the soil of El-

Hoseiny in 1999 and the same soil in 2010. The uniformity coefficient is calculated to be

3.4 (El-Hoseiny obtained 3.62) and the coefficient of concavity is calculated to be 1.18

(El-Hoseiny obtained 1.19).


156

Based on Unified Soil Classification System (USCS), the soil is classified as poorly

graded sand (SP).

Figure 3-54 Tanque Verdes soil.

Sieve Opening % %
Number Size Remaining Remaining Passing
4 4.75 0 0.00 100.00
10 1.68 406.3 40.67 59.33
20 0.85 375.1 37.54 21.79
40 0.425 156.8 15.69 6.10
50 0.3 31.5 3.15 2.94
100 0.15 19.7 1.97 0.97
200 0.075 3.9 0.39 0.58
Pan 0 5.8 0.58 0.00
Sum 999.1 100
Total 1000 g
Table 3-3 Particle size distribution curve for the soil.
157

100
90
80 Test Result (2010)
70
El-Hoseiny (1999)
% Passing

60
50
40
30
20
10
0
0.01 0.1 1 10
Opening Size (mm)

Figure 3-55 Particle size distribution curve for the soil.

Based on ASTM D698, the standard compaction test (Figure 3-56) is performed on the

soil, and the relation between the dry unit weights of the soil and the corresponding water

content is graphed. The optimum moisture content (OMC) for the soil is determined as

8.41% (El-Hoseiny (1999) obtained 8.0%) and the corresponding maximum dry unit

weight was 18.82 KN/m (El-Hoseiny (1999) obtained 18.85 KN/m) (Table 3-4).
158

Figure 3-56 Standard compaction test.

WSatSoil d d
(lbs) (lbs/ft) (lbs/ft) (KN/m)

Test 1 3.8 121.8036 4.38 116.70 18.52

Test 2 3.92 125.65 6.60 117.87 18.71

Test 3 4.01 128.5349 8.41 118.56 18.82

Test 4 4.02 128.8554 10.25 116.88 18.55

Test 5 4.06 130.1375 12.97 115.20 18.29

Table 3-4 Compaction curve data.

Compaction Curve

18.9

18.8
Dry Density (KN/m^3)

18.7

18.6

18.5

18.4

18.3

18.2
0 2 4 6 8 10 12 14
Water Content (%)

Figure 3-57 Compaction curve.

The results from sieve test and maximum dry density that El-Hoseiny (1999) obtained in

1999 are very close to the result that were obtained in 2010; therefore, the rest of the

results are assumed to be as same as El-Hoseiny (1999).


159

Property Value Units

Unified Soil Classification SP -

Specific Gravity 2.64 -

D10 0.48 mm

D30 1 mm

D60 1.75 mm

Cu 3.64 -

Cc 1.19 -

Max Void Ratio, emax 0.71 -

Min Void Ratio, emin 0.37 -

Max Dry Density 18.84 KN/m

Min Dry Density 15.35 KN/m

Optimum Moisture
Content 8 %

Table 3-5 Properties of the soil.

3.3.3. Samples Preparation

Three samples in this investigation are designated Sample A, Sample B, and Sample C.

Sample A:

This sample is made of CFRP (Carbon Fiber Reinforcement Polymer) and the layers of

fibers are saturated by epoxy resin. Fiber layer cut 25x25 cm (8x8 in), and Rhino Linings

company saturated the layers of CFRP to simulate the same process as the field. The
160

curing time vary between 15 sec to 10 min; it depends on temperature and the epoxy resin

product (Figure 3-59). The product used is called TuffGrip which is used in the back of

the trucks or boats. The advantages are excellent slip resistance, good impact resistance,

excellent abrasion resistance, excellent corrosion resistance, good chemical resistance,

and full color range available with UV top coat. After complete curing time, the sample

was cut in circular shape with diameter of 6.5 (Figure 3-58), and the sample was ready

for test.

Figure 3-58 Cutting CFRP.


161

Figure 3-59 Sample A.

Sample B:

Sample B is the same as Sample A, since the same fiber, and epoxy resin were used.

However, during the curing time, one layer of a soil (Tanque Verdes backfill soil) was

installed on the top of the fiber (Figure 3-60). In the field, it is possible to make the

sample like this, since the fibers can be saturated in the field and compacted with the

available soil in field.


162

Figure 3-60 Sample B.

Sample C:

This sample is made of Carbon FRP, it saturated with medium epoxy resin, and the

curing time for the sample is about 24 hrs. The sample is made in structures lab, and

during the curing time, one layer of soil (Tanque Verdes soil) was installed on top of the

fiber (Figure 3-61). The reason for using different epoxy resin is the cost and longer time

of curing.
163

Figure 3-61 Sample C.

3.3.4. Installing CFRP for Direct Shear Test Using CYMDOF Machine

There are two parts for running a direct shear test using CYMDOF. One is a hardware

procedure, which focuses on the installation of CFRP in CYMDOF machine. The second

one is a software procedure, which focuses more on how to control the CYMDOF

machine. This section focuses on these procedures for the installation of CFRP for a

direct shear test using CYMDOF machine.

Hardware:

This section covers all the steps that are necessary to install CFRP in CYMDOF machine.

Step1: Hydraulic pump needs to warm up for 15 minutes at 0 zero pressure before

starting any test.


164

Step 2: CYMDOF valve has to be opened, and MTS valve has to be closed to be able to

transfer the hydraulic oil to CYMDOF machine (Figure 3-62).

Figure 3-62 CYMDOF valve.

Step 3: Turn on the two cooling systems for the pump which are water cooling and fan

cooling.

Step 4: A teflon ring is placed at the lower box, and then lower sample holder on top of it.
165

Figure 3-63 Placing Teflon (a) and sample holding container(b).

Step 5: The soil is placed in the lower sample holder in five layers, and each layer is

compacted with about 25 blows. Then, the top surface of the soil sample is leveled.

Figure 3-64 Soil placed in lower sample holder.


166

Step 6: The CFRP sample is placed at the bottom of the upper guide box by six bolts that

screw to the upper guide box as shown in Figure 3-65.

Bolts

CFRP

Figure 3-65 Using bolts to attach CFRP to bottom of upper guide box.

Step 7: The upper guide box is placed on the soil sample, and then an aluminum block is

placed on top of the upper guide box as shown in Figure 3-66.


167

Figure 3-66 Aluminum block is placed on top of the upper guide box.

Software:

The CYMDOF is controlled by a computer program called TestStar, which was

originally made by MTS. In this research more than 27 displacement controlled shear

tests are conducted on CFRP and soil. The normal compression tests are conducted in

such a manner to evaluate the parameters needed for the Hierarchical Single Surface

(HISS) plasticity model for sand and CFRP interface.

The shear tests are performed under a wide range of normal stresses such as 17.5 kPa,

35kPa, 70 kPa, 140 kPa, 210 kPa, 350 kPa, 525 kPa, 875 kPa, and 1050 kPa, and the final

horizontal displacement is 10 mm. These tests are conducted at a displacement rate of

0.02 mm/sec. The loading, unloading to zero shear stress, and reloading are performed

monotonically to the final shear or horizontal displacement assigned for the test. Normal

load is applied on the sample by load control until reaches the desire normal load. Then,
168

horizontal load applied on the sample until reaches 2 mm, and then, the horizontal load

reaches zero. The horizontal load applies again until it reaches 4mm, and then the

horizontal load reaches zero again. The details of computer programming for each of the

normal stresses are as follow:


169

Normal Stress 17.5 kPa Normal Stress 35 kPa

Normal Load Rate 50 N/Sec Normal Load Rate 100 N/Sec

Max Load 375 N Max Load 748 N

Data Collected Every 1 sec Data Collected Every 1 sec

Hold 20 sec Hold 20 sec

Data Collected Every 1 sec Data Collected Every 1 sec

1. Horizontal Load Rate 0.02 mm/sec 1. Horizontal Load Rate 0.02 mm/sec

1. Max Displacement 2 mm 1. Max Displacement 2 mm

1. Horizontal Reloading Rate 50 N/sec 1. Horizontal Reloading Rate 100 N/sec

2. Horizontal Load Rate 0.02 mm/sec 2. Horizontal Load Rate 0.02 mm/sec

2. Max Displacement 4 mm 2. Max Displacement 4 mm

2. Horizontal Reloading Rate 50 N/sec 2. Horizontal Reloading Rate 100 N/sec

3. Horizontal Load Rate 0.02 mm/sec 3. Horizontal Load Rate 0.02 mm/sec

3. Max Displacement 6 mm 3. Max Displacement 6 mm

3. Horizontal Reloading Rate 50 N/sec 3. Horizontal Reloading Rate 100 N/sec

4. Horizontal Load Rate 0.02 mm/sec 4. Horizontal Load Rate 0.02 mm/sec

4. Max Displacement 8 mm 4. Max Displacement 8 mm

4. Horizontal Reloading Rate 50 N/sec 4. Horizontal Reloading Rate 100 N/sec

5. Horizontal Load Rate 0.02 mm/sec 5. Horizontal Load Rate 0.02 mm/sec

5. Max Displacement 10 mm 5. Max Displacement 10 mm

5. Horizontal Reloading Rate 50 N/sec 5. Horizontal Reloading Rate 100 N/sec

Data Collected Every 4 sec Data Collected Every 4 sec

Normal Unload Rate 100 N/sec Normal Unload Rate 100 N/sec

Min Unloading Rate 0 N Min Unloading Rate 0 N

Data Collected Every 1 sec Data Collected Every 1 sec

Table 3-6 TestStar programming for 17.5 kPa and 35 kPa normal stresses
170

Normal Stress 70 kPa Normal Stress 140 kPa

Normal Load Rate 100 N/Sec Normal Load Rate 100 N/Sec

Max Load 1497 N Max Load 2993 N

Data Collected Every 1 sec Data Collected Every 1 sec

Hold 20 sec Hold 20 sec

Data Collected Every 1 sec Data Collected Every 1 sec

1. Horizontal Load Rate 0.02 mm/sec 1. Horizontal Load Rate 0.02 mm/sec

1. Max Displacement 2 mm 1. Max Displacement 2 mm

1. Horizontal Reloading Rate 100 N/sec 1. Horizontal Reloading Rate 50 N/sec

2. Horizontal Load Rate 0.02 mm/sec 2. Horizontal Load Rate 0.02 mm/sec

2. Max Displacement 4 mm 2. Max Displacement 4 mm

2. Horizontal Reloading Rate 100 N/sec 2. Horizontal Reloading Rate 50 N/sec

3. Horizontal Load Rate 0.02 mm/sec 3. Horizontal Load Rate 0.02 mm/sec

3. Max Displacement 6 mm 3. Max Displacement 6 mm

3. Horizontal Reloading Rate 100 N/sec 3. Horizontal Reloading Rate 50 N/sec

4. Horizontal Load Rate 0.02 mm/sec 4. Horizontal Load Rate 0.02 mm/sec

4. Max Displacement 8 mm 4. Max Displacement 8 mm

4. Horizontal Reloading Rate 100 N/sec 4. Horizontal Reloading Rate 50 N/sec

5. Horizontal Load Rate 0.02 mm/sec 5. Horizontal Load Rate 0.02 mm/sec

5. Max Displacement 10 mm 5. Max Displacement 10 mm

5. Horizontal Reloading Rate 100 N/sec 5. Horizontal Reloading Rate 50 N/sec

Data Collected Every 4 sec Data Collected Every 4 sec

Normal Unload Rate 100 N/sec Normal Unload Rate 100 N/sec

Min Unloading Rate 0 N Min Unloading Rate 0 N

Data Collected Every 1 sec Data Collected Every 1 sec

Table 3-7 TestStar programming for 70 kPa and 140 kPa normal stresses
171

Normal Stress 210 kPa Normal Stress 350 kPa

Normal Load Rate 200 N/Sec Normal Load Rate 200 N/Sec

Max Load 4490 N Max Load 7484 N

Data Collected Every 1 Sec Data Collected Every 1 sec

Hold 20 Sec Hold 20 sec

Data Collected Every 1 Sec Data Collected Every 1 sec

1. Horizontal Load Rate 0.02 mm/sec 1. Horizontal Load Rate 0.02 mm/sec

1. Max Displacement 2 mm 1. Max Displacement 2 mm

1. Horizontal Reloading Rate 200 N/sec 1. Horizontal Reloading Rate 200 N/sec

2. Horizontal Load Rate 0.02 mm/sec 2. Horizontal Load Rate 0.02 mm/sec

2. Max Displacement 4 mm 2. Max Displacement 4 mm

2. Horizontal Reloading Rate 200 N/sec 2. Horizontal Reloading Rate 200 N/sec

3. Horizontal Load Rate 0.02 mm/sec 3. Horizontal Load Rate 0.02 mm/sec

3. Max Displacement 6 mm 3. Max Displacement 6 mm

3. Horizontal Reloading Rate 200 N/sec 3. Horizontal Reloading Rate 200 N/sec

4. Horizontal Load Rate 0.02 mm/sec 4. Horizontal Load Rate 0.02 mm/sec

4. Max Displacement 8 mm 4. Max Displacement 8 mm

4. Horizontal Reloading Rate 200 N/sec 4. Horizontal Reloading Rate 200 N/sec

5. Horizontal Load Rate 0.02 mm/sec 5. Horizontal Load Rate 0.02 mm/sec

5. Max Displacement 10 mm 5. Max Displacement 10 mm

5. Horizontal Reloading Rate 200 N/sec 5. Horizontal Reloading Rate 200 N/sec

Data Collected Every 4 sec Data Collected Every 4 sec

Normal Unload Rate 200 N/sec Normal Unload Rate 200 N/sec

Min Unloading Rate 0 N Min Unloading Rate 0 N

Data Collected Every 1 sec Data Collected Every 1 sec

Table 3-8 TestStar programming for 210 kPa and 350 kPa normal stresses
172

Normal Stress 525 kPa Normal Stress 875 kPa

Normal Load Rate 500 N/Sec Normal Load Rate 1000 N/Sec

Max Load 11226 N Max Load 18709 N

Data Collected Every 1 sec Data Collected Every 1 sec

Hold 20 sec Hold 20 sec

Data Collected Every 1 sec Data Collected Every 1 sec

1. Horizontal Load Rate 0.02 mm/sec 1. Horizontal Load Rate 0.02 mm/sec

1. Max Displacement 2 mm 1. Max Displacement 2 mm

1. Horizontal Reloading Rate 500 N/sec 1. Horizontal Reloading Rate 1000 N/sec

2. Horizontal Load Rate 0.02 mm/sec 2. Horizontal Load Rate 0.02 mm/sec

2. Max Displacement 4 mm 2. Max Displacement 4 mm

2. Horizontal Reloading Rate 500 N/sec 2. Horizontal Reloading Rate 1000 N/sec

3. Horizontal Load Rate 0.02 mm/sec 3. Horizontal Load Rate 0.02 mm/sec

3. Max Displacement 6 mm 3. Max Displacement 6 mm

3. Horizontal Reloading Rate 500 N/sec 3. Horizontal Reloading Rate 1000 N/sec

4. Horizontal Load Rate 0.02 mm/sec 4. Horizontal Load Rate 0.02 mm/sec

4. Max Displacement 8 mm 4. Max Displacement 8 mm

4. Horizontal Reloading Rate 500 N/sec 4. Horizontal Reloading Rate 1000 N/sec

5. Horizontal Load Rate 0.02 mm/sec 5. Horizontal Load Rate 0.02 mm/sec

5. Max Displacement 10 mm 5. Max Displacement 10 mm

5. Horizontal Reloading Rate 500 N/sec 5. Horizontal Reloading Rate 1000 N/sec

Data Collected Every 4 sec Data Collected Every 4 sec

Normal Unload Rate 500 N/sec Normal Unload Rate 500 N/sec

Min Unloading Rate 0 N Min Unloading Rate 0 N

Data Collected Every 1 sec Data Collected Every 1 sec

Table 3-9 TestStar programming for 525 kPa and 875 kPa normal stresses.
173

Normal Stress 1050 kPa

Normal Load Rate 1000 N/Sec

Max Load 22452 N

Data Collected Every 1 sec

Hold 20 sec

Data Collected Every 1 sec

1. Horizontal Load Rate 0.02 mm/sec

1. Max Displacement 2 mm

1. Horizontal Reloading Rate 1000 N/sec

2. Horizontal Load Rate 0.02 mm/sec

2. Max Displacement 4 mm

2. Horizontal Reloading Rate 1000 N/sec

3. Horizontal Load Rate 0.02 mm/sec

3. Max Displacement 6 mm

3. Horizontal Reloading Rate 1000 N/sec

4. Horizontal Load Rate 0.02 mm/sec

4. Max Displacement 8 mm

4. Horizontal Reloading Rate 1000 N/sec

5. Horizontal Load Rate 0.02 mm/sec

5. Max Displacement 10 mm

5. Horizontal Reloading Rate 1000 N/sec

Data Collected Every 4 sec

Normal Unload Rate 500 N/sec

Min Unloading Rate 0 N

Data Collected Every 1 sec

Table 3-10 TestStar programming for 1050 kPa normal stress.


174

3.3.5. Test Result

Interface laboratory results between Sample A and backfill soil:

Normal Stress 17.5 kPa

60

50
Shear Stress (kPa)

40

30

20

10

0
0 2 4 6 8 10
-10
Shear Displacement (mm)

Figure 3-67 Interface result for backfill soil and sample A with n=17.5 kPa.

Normal Stress 17.5 kPa

0.25
Normal Displacement (mm)

0.2

0.15

0.1

0.05

0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-68 Tangential displacement vs. normal displacement for backfill soil & sample A with
n=17.5 kPa.
Note: Normal displacements (normal displacement vs. vertical displacement figures) are
plotted in the negative direction.
175

Normal Stress 35 kPa

80
70
Shear Stress (kPa)

60
50
40
30
20
10
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-69 Interface result for backfill soil and sample A with n=35 kPa.

Normal Stress 35 kPa

0.3
Normal Displacement (mm)

0.25

0.2

0.15

0.1

0.05

0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-70 Tangential displacement vs. normal displacement for backfill soil & sample A with n=35
kPa.
176

Normal Stress 70 kPa

120

100
Shear Stress (kPa)

80

60

40

20

0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-71 Interface result for backfill soil and sample A with n=70 kPa.

Normal Stress 70 kPa

0.35
Normal Displacement (mm)

0.3
0.25
0.2
0.15
0.1
0.05
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-72 Tangential displacement vs. normal displacement for backfill soil & sample A with n=70
kPa
177

Normal Stress 140 kPa

200
180
160
Shear Stress (kPa)

140
120
100
80
60
40
20
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-73 Interface result for backfill soil and sample A with n=140 kPa.

Normal Stress 140 kPa

0.35
Normal Displacement (mm)

0.3
0.25
0.2
0.15
0.1
0.05
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-74 Tangential displacement vs. normal displacement for backfill soil & sample A with
n=140 kPa.
178

Normal Stress 210 kPa

300

250
Shear Stress (kPa)

200

150

100

50

0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-75 Interface result for backfill soil and sample A with n=210 kPa

Normal Stress 210 kPa

0.35
Normal Displacement (mm)

0.3
0.25
0.2
0.15
0.1
0.05
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-76 Tangential displacement vs. normal displacement for backfill soil & sample A
withn=210 kPa
179

Normal Stress 350 kPa

450
400
Shear Stress (kPa)

350
300
250
200
150
100
50
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-77 Interface result for backfill soil and sample A with n=350 kPa.

Normal Stress 350 kPa

0.6
Normal Displacement (mm)

0.5

0.4

0.3

0.2

0.1

0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-78 Tangential displacement vs. normal displacement for backfill soil & sample A with
n=350 kPa
180

Normal Stress 525 kPa

600

500
Shear Stress (kPa)

400

300

200

100

0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-79 Interface result for backfill soil and sample A with n=525 kPa

Normal Stress 525 kPa

0.5
Normal Displacement (mm)

0.45
0.4
0.35
0.3
0.25
0.2
0.15
0.1
0.05
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-80 Tangential displacement vs. normal displacement for backfill soil & sample A with
n=525 kPa
181

Normal Stress 875 kPa

900
800
Shear Stress (kPa)

700
600
500
400
300
200
100
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-81 Interface result for backfill soil and sample A with n=875 kPa.

Normal Stress 875 kPa

0.6
Normal Displacement (mm)

0.5

0.4

0.3

0.2

0.1

0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-82 Tangential displacement vs. normal displacement for backfill soil & sample A with
n=875 kPa.
182

Normal Stress 1050 kPa

1000
900
800
Shear Stress (kPa)

700
600
500
400
300
200
100
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-83 Interface result for backfill soil and sample A with n=1050 kPa.

Normal Stress 1050 kPa

0.7
Normal Displacement (mm)

0.6
0.5
0.4
0.3
0.2
0.1
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-84 Tangential displacement vs. normal displacement for backfill soil & sample A with
n=1050 kPa.
183

Sample A/Backfill

Normal Stress (kPa) Shear Stress (kPa)

17.5 49.97

35 74.69

70 112.75

140 175.32

210 257.66

350 412.04

525 555.02

875 778.84

1050 996.78

Interface Friction
Angle 41.45 Degree

Adhesin 58.04 kPa

Table 3-11 Normal stress vs. max shear stress for sample A and backfill soil.
184

1200

1000
Shear Stress (kPa)

800

600
y = 0.8833x + 58.04
400 R = 0.9922
200

0
0 200 400 600 800 1000 1200
Normal Stress (kPa)

Figure 3-85 Normal stress vs. max shear stress for sample A and backfill soil.
185

Interface laboratory results between Sample B and backfill soil:

Normal Stress 17.5 kPa

70
60
Shear Stress (kPa)

50
40
30
20
10
0
0 2 4 6 8 10 12
Shear Displacement (mm)

Figure 3-86 Interface result for backfill soil and sample B with n=17.5 kPa.

Normal Stress 17.5 kPa

0
Normal Displacement (mm)

0 2 4 6 8 10
-0.1

-0.2

-0.3

-0.4

-0.5

-0.6

-0.7
Shear Displacement (mm)

Figure 3-87 Tangential displacement vs. normal displacement for backfill soil & sample B with
n=17.5 kPa.
186

Normal Stress 35 kPa

80
70
Shear Stress (kPa)

60
50
40
30
20
10
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-88 Interface result for backfill soil and sample B with n=35 kPa.

Normal Stress 35 kPa

0.3
Normal Displacement (mm)

0.25

0.2

0.15

0.1

0.05

0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-89 Tangential displacement vs. normal displacement for backfill soil & sample B with n=35
kPa
Note: Normal displacements (normal displacement vs. vertical displacement figures) are
plotted in the negative direction.
187

Normal Stress 70 kPa

140
120
Shear Stress (kPa)

100
80
60
40
20
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-90 Interface result for backfill soil and sample B with n=70 kPa

Normal Stress 70 kPa

0.4
Normal Displacement (mm)

0.35
0.3
0.25
0.2
0.15
0.1
0.05
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-91 Tangential displacement vs. normal displacement for backfill soil & sample B with n=70
kPa
188

Normal Stress 140 kPa

200
180
160
Shear Stress (kPa)

140
120
100
80
60
40
20
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-92 Interface result for backfill soil and sample B with n=140 kPa.

Normal Stress 140 kPa

0.5
Normal Displacement (mm)

0.45
0.4
0.35
0.3
0.25
0.2
0.15
0.1
0.05
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-93 Tangential displacement vs. normal displacement for backfill soil & sample B with
n=140 kPa.
189

Normal Stress 210 kPa

300

250
Shear Stress (kPa)

200

150

100

50

0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-94 Interface result for backfill soil and sample B with n=210 kPa.

Normal Stress 210 kPa

0.7
Normal Displacement (mm)

0.6
0.5
0.4
0.3
0.2
0.1
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-95 Tangential displacement vs. normal displacement for backfill soil & sample B with
n=210 kPa.
190

Normal Stress 350 kPa

450
400
Shear Stress (kPa)

350
300
250
200
150
100
50
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-96 Interface result for backfill soil and sample B with n=350 kPa.

Normal Stress 350 kPa

0.5
Normal Displacement (mm)

0.45
0.4
0.35
0.3
0.25
0.2
0.15
0.1
0.05
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-97 Tangential displacement vs. normal displacement for backfill soil & sample B with
n=350 kPa.
191

Normal Stress 525 kPa

700
600
Shear Stress (kPa)

500
400
300
200
100
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-98 Interface result for backfill soil and sample B with n=525 kPa.

Normal Stress 525 kPa

0.7
Normal Displacement (mm)

0.6
0.5
0.4
0.3
0.2
0.1
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-99 Tangential displacement vs. normal displacement for backfill soil & sample B with
n=525 kPa.
192

Normal Stress 875 kPa

1000
900
800
Shear Stress (kPa)

700
600
500
400
300
200
100
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-100 Interface result for backfill soil and sample B with n=875 kPa.

Normal Stress 875 kPa

0.9
Normal Displacement (mm)

0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-101 Tangential displacement vs. normal displacement for backfill soil & sample B with
n=875 kPa.
193

Normal Stress 1050 kPa

1200

1000
Shear Stress (kPa)

800

600

400

200

0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-102 Interface result for backfill soil and sample B with n=1050 kPa.

Normal Stress 1050 kPa

0.8
Normal Displacement (mm)

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-103 Tangential displacement vs. normal displacement for backfill soil & sample B with
n=1050 kPa.
.
194

Sample B/Backfill

Normal Stress (kPa) Shear Stress (kPa)

17.5 43.73

35 75.6

70 123.99

140 188.86

210 269.06

350 400.22

525 645

875 894.95

1050 1007.18

Interface Friction
Angle 44.68 Degree

Adhesin 60.43 kPa

Table 3-12 Normal stress vs. max shear stress for sample B and backfill soil.
195

1200

1000
Shear Stress (kPa)

800

600
y = 0.9487x + 60.43
400 R = 0.9889

200

0
0 200 400 600 800 1000 1200
Normal Stress (kPa)

Figure 3-104 Normal stress vs. max shear stress for sample B and backfill soil.
196

Interface laboratory results between Sample C and backfill soil:

Normal Stress 17.5 kPa

50
45
40
Shear Stress (kPa)

35
30
25
20
15
10
5
0
0 2 4 6 8 10

Shear Displacement (mm)

Figure 3-105 Interface result backfill for soil and Sample C with n=17.5 kPa.

Normal Stress 17.5 kPa

0.15
Normal Displacement (mm)

0.1

0.05

0
0 2 4 6 8 10
-0.05

-0.1

-0.15

-0.2

Shear Displacement (mm)

Figure 3-106 Tangential displacement vs. normal displacement for backfill soil & Sample C with
n=17.5 kPa.
197

Normal Stress 35 kPa

60

50
Shear Stress (kPa)

40

30

20

10

0
0 2 4 6 8 10

Shear Displacement (mm)

Figure 3-107 Interface result for backfill soil and Sample C with n=35 kPa.

Normal Stress 35 kPa

0.5
Normal Displacement (mm)

0.45
0.4
0.35
0.3
0.25
0.2
0.15
0.1
0.05
0
0 2 4 6 8 10

Shear Displacement (mm)

Figure 3-108 Tangential displacement vs. normal displacement for backfill soil & Sample C with
n=35 kPa.
Note: Normal displacements (normal displacement vs. vertical displacement figures) are
plotted in the negative direction.
198

Normal Stress 70 kPa

120

100
Shear Stress (kPa)

80

60

40

20

0
0 2 4 6 8 10

Shear Displacement (mm)

Figure 3-Interface result for backfill soil and Sample C with n=70 kPa.

Normal Stress 70 kPa

0.6
Normal Displacement (mm)

0.5

0.4

0.3

0.2

0.1

0
0 2 4 6 8 10

Shear Displacement (mm)

Figure 3-109 Tangential displacement vs. normal displacement for backfill soil & Sample C with
n=70 kPa.
199

Normal Stress 140 kPa

180
160
Shear Stress (kPa)

140
120
100
80
60
40
20
0
0 2 4 6 8 10

Shear Displacement (mm)

Figure 3-110 Interface result for backfill soil and Sample C with n=140 kPa.

Normal Stress 140 kPa

0.6
Normal Displacement (mm)

0.5

0.4

0.3

0.2

0.1

0
0 2 4 6 8 10

Shear Displacement (mm)

Figure 3-111 Tangential displacement vs. normal displacement for backfill soil & Sample C with
n=140 kPa.
200

Normal Stress 210 kPa

250

200
Shear Stress (kPa)

150

100 `

50

0
0 2 4 6 8 10

Shear Displacement (mm)

Figure 3-112 Interface result for backfill soil and Sample C with n=210 kPa.

Normal Stress 210 kPa

0.8
Normal Displacement (mm)

0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0 2 4 6 8 10

Shear Displacement (mm)

Figure 3-113 Tangential displacement vs. normal displacement for backfill soil & Sample C with
n=210 kPa.
201

Normal Stress 350 kPa

400
350
Shear Stress (kPa)

300
250
200
150
100
50
0
0 2 4 6 8 10

Shear Displacement (mm)

Figure 3-114 Interface result for backfill soil and Sample C with n=350 kPa

Normal Stress 350 kPa

0.9
Normal Displacement (mm)

0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0 2 4 6 8 10

Shear Displacement (mm)

Figure 3-115 Tangential displacement vs. normal displacement for backfill soil & Sample C with
n=350 kPa.
202

Normal Stress 525 kPa

700
600
Shear Stress (kPa)

500
400
300
200
100
0
0 2 4 6 8 10

Shear Displacement (mm)

Figure 3-116 Interface result for backfill soil and Sample C with n=525 kPa.

Normal Stress 525 kPa

1.2
Normal Displacement (mm)

0.8

0.6

0.4

0.2

0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 3-117 Tangential displacement vs. normal displacement for backfill soil & Sample C with
n=525 kPa.
203

Normal Stress 875 kPa

700
600
Shear Stress (kPa)

500
400
300
200
100
0
0 2 4 6 8 10

Shear Displacement (mm)

Figure 3-118 Interface result for backfill soil and Sample C with n=875 kPa.

Normal Stress 875 kPa

1
Shear Displacement (mm)

0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0 2 4 6 8 10

Shear Displacement (mm)

Figure 3-119 Tangential displacement vs. normal displacement for backfill soil & Sample C with
n=875 kPa.
204

Normal Stress 1050 kPa

900
800
Shear Stress (kPa)

700
600
500
400
300
200
100
0
0 2 4 6 8 10

Shear Displacement (mm)

Figure 3-120 Interface result for backfill soil and Sample C with n=1050 kPa

Normal Stress 1050 kPa

1
Normal Displacement (mm)

0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0 2 4 6 8 10

Shear Displacement (mm)

Figure 3-121 Tangential displacement vs. normal displacement for backfill soil & Sample C with
n=1050 kPa.
205

Sample C/Backfill

Normal Stress (kPa) Shear Stress (kPa)

17.5 45.23

35 49.75

70 105.23

140 162.41

210 209.5

350 377.85

525 576.09

875 636.08

1050 806.58

Interface Friction
Angle 36.13 Degree

Adhesin 64.92 kPa

Table 3-13 Normal stress vs. max shear stress for sample C and backfill soil.
206

900
800
700
Shear Stress (kPa)

600
500
400
300 y = 0.7286x + 64.921
R = 0.9564
200
100
0
0 200 400 600 800 1000 1200
Normal Stress (kPa)

Figure 3-122 Normal stress vs. max shear stress for sample C and backfill soil.

3.3.6. Discussion

Sample A, Sample B, and Sample C have friction angles of 41.45, 44.68, and 36.13

degrees, respectively. The adhesion for Sample A, Sample B, and Sample C are 58.04

kPa, 60.43 kPa, and 62.94 kPa, respectively.

Based on all the results obtained from CYMDOF, Sample A, which is made of flexible

resin with no soil on the surface, is chosen for further investigations. Although, Sample

B, which is made of flexible resin with soil on the surface, has a higher friction angle, it

is more expensive and harder to prepare.


207

Summary

SoilCFRP interaction has been modeled experimentally in many different ways for the

last three decades. The main objective of this chapter was to discuss experimental tools

for the understanding and evaluation of such interaction, which is relevant to most of the

applications of CFRP. Pull out and direct shear tests were used to investigate the

interface behavior of fine sand and CFRP using epoxy CW200 for a small

scale. CYMDOF test was used to investigate the interface behavior of backfill soil and

CFRP using epoxy Rhino Linings for a larger investigation. Since CFRP using Rhino

Linings are more flexible compared to CFRP using CW 200, CFRP using Rhino Linings

is used for further investigation. For further investigation, between the two soils, backfill

soil was selected since the backfill soil was used in a MSE wall (more detail are provided

in Chapter 6) which was monitored with different sensors at different location of the

MSE wall. Therefore, the Finite element analysis can compare with the field data.
208

CHAPTER 4. CONSTITUTIVE MODELS FOR SOIL AND INTERFACE

4.1. Introduction

A constitutive model describes the behavior of an engineering material when subjected to

external loads. Their definition, mathematical and experimental, is essential for realistic

prediction from solution procedures based on computation mechanics. The power of a

constitutive model depends on the extent to which the physical phenomenon has

simulated. The behavior of engineering materials can be considered at the different

levels, e.g. microscopic, macroscopic. There are a number of constitutive models, and

there is no universal one, yet. Based on the theory of elasticity available in which the

generalized Hookes law is used to define linear or piecewise linear elastic stressstrain

response. However, interfaces between geologic and structural material exhibit a complex

behavior that cannot be modeled by the elastic models. The ideas, concepts and some of

the literary phrasing used in this chapter are adopted and derived from the text by Desai

(2001).

Desai (2005b) reported that A constitutive law or model represents a mathematical

definition for the behavior of a material based on laboratory and/or field tests that includes

significant factors affecting behavior. Hence, a constitutive model should predict the behavior

of a material for all significant factors that influence the actual (field) response of the material.

In this sense, a classical elastic model is a constitutive law only if the material is isotropic,

linear elastic, and is not affected by factors other than loading under a given stress path.

Hence, available models are applicable only for the factors for which they are developed.
209

There have been numerous efforts to develop constitutive relations for soils and

interfaces. The development of these constitutive laws necessitates consideration of

physical laws as well as observing their behavior in laboratory and field conditions.

A more accurate constitutive model depends on the amount to which the physical

phenomenon has been simulated. One disadvantage of constitutive model is that there

are different types constitutive models. Some models may better the behavior of some

material and others may not be able to simulate the other class of material. Many

materials obey Hooke's law as long as the load does not surpass the material's elastic limit.

A material for in which Hooke's law is consider as good approximation is known as a

linear-elastic or "Hookean" material. In its simplest terms, Hooke's law states that strain

is directly proportional to stress.

It should be noted that geological materials such as interface behavior have extremely

complex behavior and the stress and strain behavior is non-linear. These materials can

have dilative behavior under shear loading. In this study, Hierarchical Single Surface

(HISS) plasticity is used to model the behavior of the interface.

The hierarchical single-surface (HISS) plasticity models give a general formulation for

the elastoplastic categorization of the material behavior. These models can allow for

isotropic and anisotropic hardening. They also allow for associated and nonassociated

plasticity characterizations which can be used to signify material response based on the

continuum plasticity theory (Desai, 1994). In this research, special case of HISS mode is

used to simulate the intact behavior of the interface between soil and CFRP. The most
210

basic and simplest version of the HISS-0 that allows for isotropic hardening and

associated response has been used. The following sections briefly address different type

of theories that is used for back predicting the lab results.

4.2. Different Types of Constitutive Models

4.2.1. Elasticity Theory

Elastic material has one to one relationship between stress and strain. Elastic materials

follow the same stress strain path during loading and unloading cycles. The energy

density function can be articulated in terms of the state of current strain, and independent

of the strain history or strain path.

A Material is called linear elastic if the stress-strain relationship is linear, and the material

on reloading returns to original confirmation using same path. Otherwise, it is nonlinear

elastic material (Figure 4-1). The linear stress strain relation is as follow:

ij=Cijklkl (4-1)

where ij is the stress tensor, kl is the strain tensor, and Cijkl is constitutive tensor for

linear elastic materials The nonlinear response is approximated as piecewise linear by

dividing the nonlinear behavior in a number of linear increments. The material

parameters are often assumed to be constant within each increment. The monotonic

response of some nonlinear materials is often simulated by using the piecewise linear

approximation in which the parameters or moduli (Et, vt, Gt, and Kt) vary from increment
211

to increment (Desai, 2001) For a nonlinear (piecewise) material, the stress-strain relation

is

ij=Cij + Cijklkl + Cijklkl + Cijklmnklmn (4-2)

Figure 4-1 Schematic of Stress-Strain Response of linear and nonlinear Elastic Body (Desai, 2001).

4.2.2. Plasticity Theory

In plasticity, there exists deformation in a material that is non-reversible (Lubliner, 2008).

As an example, a solid piece of metal being that is bent or pounded into a new shape

shows plasticity as permanent changes occur within the material itself. The transition

from elastic behavior to plastic behavior is called yield.

Plastic deformation can be observed in many materials including metals, soils, rocks,

concrete, foams, bone and skin (Jirasek et al., 2002; Chen, 2008; Chen, 2007; Leveque et
212

al., 1993). The physical mechanisms that cause plastic deformation can widely differ.

The irrecoverable deformation portion of strain is called the plastic strain, p, and the

recoverable part of strain is known as elastic strain e (Desai, 2001). The yield function

for uniaxial state of stresses as shown in Figure 4-2 can written as

F=-y=0 (4-3)

Figure 4-2 Schematic of Stress-Strain Response in Elastoplastic Deformaion (Desai, 2001).

is the present state of stress of stresses and y is the stress level where the beginning of

yielding happens. For an isotropic material, the yield condition in Equation (4-3) can be

expressed as

F=F(1 , 2 , 3) (4-4)
213

1, 2 and 3 are the principal stresses. For a number soils, the formulation of constitutive

law based on the assumption that the yield function (F) is the same as the plastic potential

formula (Q) is called associative plasticity (Desai and Hashmi, 1985). Hill (1950) and

Mendelson (1968) developed von Mises criterion for the yield in metals, in which the

second invariant of deviatoric stress tensor J2D is equal to a constant k2 (k is the material

parameter, which can be calculated from the laboratory test), the yield function is given

by:

F= J2D - k2=0 (4-5)

The shear strength increases as normal stress increases on the failure plane, and the

Mohr-Coulomb criterion is given by:

= c + n tan (4-6)

The geometric representation of von Mises yield surface is a circle and the Mohr-

Coulomb yield surface is shown as a hexagon (Figure 4-3). Von Mises criterion

demonstrates that the yield strength of material is the same in compression and tension

and it is valid for most metals.


214

Figure 4-3 Representation of yield surface for von Mises and Mohr-Coulomb criteria in the
octahedral (Desai, 2001).

For concrete and geological materials, von Mises yield surface is inadequate, since the

compressive strength and tensile strength are different. Hence, since the Mohr-Coulomb

criterion dependents on confining pressure, it is used often for cohesionless materials. In

other words, it allows for the effect of friction and is suitable for frictional geologic

materials.

Nearly all geologic materials show yielding or hardening from the beginning of loading,

and the yeild stress increases continuously with deformation. However, in classical

plasticity, a yield function is in term of the state stress, and there is no consideration

given to the internal cohesion physical state of the material during deformation (Desai

and Siriwardane, 1984).

Hardening models are used for materials that demonstrate yielding from the beginning of

loading. Roscoe et. al. (1958) developed critical state concept (CSC) for material

yielding from the beginning. Critical State encompasses the conceptual models that
215

demonstrate the mechanical behavior of saturated remolded soils based on the Critical

State concept. The original Critical State concept is an idealization of the observed

behavior of saturated remolded clays in triaxial compression tests, and it is assumed to

apply to undisturbed soils. It states that soils and other granular materials, are

continuously sheared until they flow as a frictional fluid and they will come into a well-

defined critical state. At the beginning of the critical state, shear distortions occur without

any further changes in mean effective stress, deviatoric stress (or yield stress, in uniaxial

tension according to the von Mises yielding criterion), or specific volume. In other words,

a granular material under shear stress passes through progressive deformation states and

eventually approaches the critical state at which no further volume change occurs.

Loose frictional material under shear loading experiences various states of yielding

before attaining the critical state. The volume constantly decreases until it reaches

critical state and the particles are arranged so that there is no longer a volume change

occurring with further shearing. This particular void ratio is called the critical void ratio.

On the other hand, in a dense material, the volume first decreases, and then the material

dilates until the volume reaches a constant value as shown in Figure 4-4 (Casagrande,

1936).
216

Figure 4-4 Loose sand and dense sand.

4.2.3. Hierarchical Single Surface (HISS) Model

HISS models can allow for isotropic and anisotropic hardening. They are able to

characterize material response based on the continuum plasticity theory (Desai, 2001;

Desai, 1980, 1994, and 1995; Desai & Faruque, 1984; Desai et al., 1986; Desai and

Wathugala, 1987). A constitutive model based on the hierarchical single surface (HISS)

approach and the elastoplasticity theory is proposed to illustrate the behavior of interfaces

that are subject to static and cyclic loading situations. In the hierarchical system, a

model describing fundamental characteristics of the interfaces is tailored to include the

increasingly complex behavior of the interfaces. The proposed model can simulate

associative (HISS-0) and nonassociative (HISS-1), and monotonic and cyclic loading
217

(Desai 2001). The model is depend based on interface behavior from laboratory tests

which are shear tests on sand-steel and sand-concrete interfaces.

It is often practical to stress one of the important advantages offered by the hierarchical

nature of the HISS model. Most other available models provide for a specific

characteristic of the material behavior. However, the HISS approach provides for

hierarchical acceptance of models of increasing complexity. For example, going from

linear elastic to nonassociated elastoplasticity to elastoplastic models (Desai, 2001).

The yield function (F) for associated plasticity model for solid in the hierarchical single

surface is given as follow:

F =J 2D-(- J n1 + J 21)(1-Sr)m (4-7)

= J 2D - FbFS=0 (4-8)

J 2D= J2D /p2a (4-9)

where

J2D= the second invariant of the deviatoric stress tensor

pa= the atmospheric pressure constant

J 1= (J1+3R)/pa (4-10)

J1=the first invariant of the total stress tensor

R= Parameter related to the cohesive strength


218

m=-0.5

=Parameter related to the ultimate surface

= related to the shape of F in the 1-2-3 space

Sr= Stress ratio= (27/2).(J3D/J3/22D) (4-11)

where

= Growth or hardening parameter

= (, v, D)

Simple hardening function

=h1/h2=a1/ 1 (4-12)

Hashmis composite hardening function (Desai and Hashmi, 1989):

=h1 exp [-h2 (1-(D/(h3+h4D))] (4-13)

Wathugalas composite hardening function (Desai et. al., 1991):

=h1 / (v+h3Dh4)h2 (4-14)

= (dpij dpij)1/2 (4-15)

where

v= Trajectory volumetric strains

D= Trajectory of deviatoric plastic strains

= v + D= . /
+ 1/3 (4-16)

Fb= - J n1+ J21 represents the shape of the yield surface in the J1-J2D space
219

Fs= (1-Sr)m The shape of the yield surface in the octahedral plane

For the associative model (0), the plastic potential function (Q) is the same as the yield

function (F). Graphical representations of the yield surface for a typical soil in J1-J2D

space are shown in following figures (Figure 4-5, Figure 4-6 and Figure 4-7) for triaxial

plane and octahedral plane.

Figure 4-5 J1-J2D (Desai, 2001).


220

Figure 4-6 Octahedral plane (Desai, 2001).

Figure 4-7 1-2-3 space (Desai, 2001).


221

Figure 4-8 Decomposition of behavior for observed stress-strain as normally consolidated soil and of
part causing overconsolidaton (Desai, 2001).

The relationship between stress and strain with no disturbance can be written as

daij= Ciijkl dikl (4-17)

iij is strain tensor, and Ciijkl can be define by using elastic, elastoplastic theories etc., and

they are characterized as follow,

. .
/ (4-18)
.

In chapter 5 HISS model is described in more detail, and the parameters for HISS model

are defined from the laboratory results by using the above equations for two dimensional

interfaces. HISS model is used to drive the shear stress vs. shear deformation for CFRP-

sand.
222

Summary

Engineering materials are difficult to characterize in their initial natural or artificially

manufactured states. Characterization of their behavior under a variety of possible

forces-natural, mechanical, and environmental also poses a challenging problem. Human

understanding of the behavior of materials, which are a mixture of continuous and

discontinuous particle system at the same time, involves mental, physical, and

mathematical models; the latter are often used to develop numerical models for solution

by the artificial mind, which is the modern computer (Desai, 2001). Different models

are introduced by researches to predict the behavior of materials under loading. This

chapter briefly describes three different theories (Elasticity, Plasticity, and Hierarchical

Single Surface Model). To study and evaluate the interaction between soils and CFRP

numerically, HISS model is used. In chapter 5, HISS model is described in more detail,

and the parameters for HISS model are defined from the lab results by using the

equations from this chapter for two dimensional interfaces. HISS model is used to predict

shear stress vs. shear deformation for CFRP-sand.


223

CHAPTER 5. DETERMINATION OF MATERIALS PARAMETERS AND


MODEL VERIFICATION

5.1. Introduction

Constitutive models are often used to predict the complex behavior of a geologic material

such as interface behavior under loading. Based on experimental data, the model needed

to be calibrated, before using the constitutive models to simulate the behavior of the

material. In order to calibrate the model, the necessary constants of the material need to

be found. The theories of elasticity and plasticity are used to investigate the constitutive

models for geologic materials. Plasticity theory requires a defined yield surface, a plastic

potential surface, and a hardening law. Therefore, it requires some material constants.

In this research, the parameters that are related with the HISS model using different states

of a material deformation such as elastic, plastic related to hardening, phase change from

compression to dilation and ultimate behavior are found. The ideas, concepts and some

of the literary phrasing used in this chapter are adopted and derived from Mechanics of

Materials and Interfaces text (Desai, 2001).

There are nine parameters for the material properties in HISS model, and they are as

follows:

1. Elastic Constant

E,

2. Plasticity Constant

Constants associated with ultimate state

, , m
224

Constants associated with the transition from contractive to dilative

Constants related to growth function

a1 , 1

The non-associative constant

5.2. Soil

For all CYMDOF test, the same soil is used as was used by El-Hoseiny (1999). El-

Hoseiny verified the model with respect to the laboratory test data of sand using CTC

tests by using HISS-1 and DSC models and back-prediction of the triaxial compression

tests of backfill soil. Then, he used finite element analysis for laboratory test involving

conventional triaxial tests; a typical comparison is as shown in Figure 5-1.


225

Material Constants Symbol Value

E 62X1030.28
Elastic Constants
0.3

0.12
Plasticity Constants
0.45

Phase Change Parameter n 2.56

a1 3.00E-05
Growth Parameters
0.98

Non-associative Constant 0.2

Du 0.93

Disturbance Parameters A 0.37

Z 1.6

Table 5-1 Material Constants for Tanque Verde Pantano Wash Sand.
226

Figure 5-1 Typical comparisons between predicted and observed responses for soil and interface
(Desai and El-Hoseiny, 2005).

5.3. Interface

One of the definitions of an interface is when two bodies made of different materials are

in contact. An example would be FRP and backfill soil (Figure 5-2) and the existence of

a finite smeared zone between the two that behaves as an interface. In other words, the

interface may be defined as a thin layer between two contact surfaces. The contact

surface can be extremely rough to medium-rough to smooth. The surface depends on the

characteristics of irregular asperities. In macro level and/or micro level, smooth contact

involves asperities, and an ideally smooth surface is essentially only hypothetical.


227

Figure 5-2 Interface between CFRP and backfill soil.

A model can be determined for the bulk interface or joint that can provide a practical

simulation of relative motions between two surfaces. The interface zone is referred to as

a thin layer with a thickness, t, that can be treated as equivalent or a smeared zone

between two materials (Desai et al, 1984; Sharma and Desai 1992).

The relative displacements of the interface are normal displacement (v) and tangential

displacement (u) (Desai and Ma, 1992).

u= ue+u p+us (5-1)

v=ve+v p+vs (5-2)

e, p and s denote elastic, plastic and slip deformations, respectively; however, the plastic

deformation and slip deformation can be combined into one as follows:

u= ue+up (5-3)

v=ve+vp (5-4)
228

Interface stresses for a two-dimensional interface (Figure 5-2) are given by:

=T/A (5-5)

n=N/A (5-6)

and n are shear stress and normal stress, respectively, and A is the nominal area of the

interface; T and N are shear force and normal force, respectively. In this research, the

total normal stresses are the same as the effective normal stresses, since the condition is

dry.

Based on the results for the interface between dense sand and FRP using CYMDOF, no

softening is observed; therefore, the disturbance value is zero. For that reason, HISS-0 is

used for back prediction of the lab results. The elastoplastic 0 version for the response

of interfaces involves the following parameters (their equations are given later):

Model Constants Comment

Elastic ks, kn Shear and normal stiffness

a, b, n, Straight ultimate envelope


Plasticity
a, b, n, , q Curved ultimate envelope

None Zero strength

Critical State Kn Normal strength

c0, c1, c2, vr0, Critical state

Table 5-2 HISS-0 constant parameters.


229

5.3.1. Elastic Parameters for HISS Model

The notation for shear stiffness is ks, and it is defined as the unloading slopes of vs. ur

curve. kn is normal stiffness, and it is the unloading slopes of n vs. vr curve. If they are

considered to be variable in a nonlinear elastic model, they can be expressed as functions

of factors such as shear stress and normal stress. Normal stiffness can be obtained from

compression test as shown in Figure 5-3. Based on Figure 5-3, it can see that the slope of

CFRPSand, normal stiffness, varies with different confining pressure; therefore, normal

stiffness is defined in terms of normal stress.

1000
NormalStress(kPa)

800

600

400

200

0
0 0.0005 0.001 0.0015 0.002 0.0025
NormalDisplacement(m)

Figure 5-3 Normal Displacement vs. Normal Stress.

After finding slopes for corresponding to normal stresses, the results of normal stiffness

vs. normal stresses are plotted in Figure 5-4. Then, the best fit curve is passed through

the normal stresses and normal stiffness (Figure 5-4), and corresponding equation are

shown as:
230

kn= -0.2588n2 + 986.96n + 129355 (5-7)

1.0E+06
9.0E+05
NormalStiffness(kPa/m)

8.0E+05
7.0E+05
6.0E+05
5.0E+05
4.0E+05
3.0E+05
2.0E+05
1.0E+05
0.0E+00
0 200 400 600 800 1000 1200
NormalStress(kPa)

Figure 5-4 Variation of Normal Stiffness with Normal Stress for Interface.

Shear stiffness (ks) is determined from the data of one way cycling of shear stress vs.

tangential displacement. For all different normal stresses, the second unloading is used to

determine the shear stiffness as shown in Figure 5-5.


231

Normal Stress 875 KPa

900
800 Ks
Shear Stress (KPa)

700
600
500
400
300
200
100
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 5-5 Determination of interface shear stiffness, ks.

Figure 5-6 shows shear stiffness versus normal stress, based on the lab result as shown in

chapter 3 for FRP-Sand for different normal stress. Then, the best curve fitting passing

through the following data, and corresponding equation for the best curve fitting are

shown as follows:
232

4.0E+05
3.5E+05
ShearStiffness(kPa/m)

3.0E+05
2.5E+05
2.0E+05
1.5E+05
1.0E+05
5.0E+04
0.0E+00
0 200 400 600 800 1000 1200
NormalStress(kPa)

Figure 5-6 Variation of shear stiffness vs. normal stress

ks = -0.2096 n2 + 432.95 n + 95299 (5-8)

5.3.2. Elastoplastic Parameters for HISS Model

As mentioned before, HISS plasticity version 0 is considered. Desai (1995) showed that

for a two dimensional interface, the yield function, F (Equation(4-7)) can be shown as

F=2 + *n-n*q (5-9)

n* =R+ n (5-10)

R is the intercept along (when n = 0), it is related to adhesion (Figure 5-8). For the

linear ultimate envelope, R can be found as


233

R=c0/ (5-11)

is the growth function, as given,

=a/b (5-12)

a and b are the hardening parameters and is the trajectory of plastic shear and normal

displacements:

=(duprdupr + dvprdvpr)1/2 (5-13)

is computed at various points on the vs. ur and n vs. vr curves for stresses at those

points. Then the values on a and b can also be found by following equation:

ln a b ln = ln (5-14)

The parameter b denotes rate of hardening, which is the slope of average curve, and a is

denoted as the value of when is equal to one which is intercept when ln is equal to

zero (Figure 5-7).


234

Figure 5-7 Determination of hardening parameters (Desai, 2001).

The parameter is determined from the slope of the ultimate envelope in shear stress vs.

normal stress space (Figure 5-8).

u = nq/2 (5-15)

= u2 / nq (5-16)
235

Figure 5-8 Yield and ultimate surface (Desai, 2001).

The value of q is for the curved envelope, when shear stress is the ultimate shear stress.

Therefore, is equal to zero. For most cases, the value of q is equal to two, which

represents a straight line. However, the value of q is determined from the following

equation:

ln u = ln + q/2 ln n (5-17)

The slope of graph ln u verses ln n represent the value of q from the slope of the

average line. The n value (phase or state change parameter) is determined from state

of stress at the transition point (Figure 5-9), which is changes from compression to

dilation. The value of n can be determined from the following:

n= q/(1-2/nq) (5-18)
236

and q are the stresses at the transition point.

Figure 5-9 State of stress at transition point.


237

For back calculating, the ultimate yield surface or envelope that represents the asymptotic

stress states will be an average straight line with slope equal to , where q is a parameter.

The value of q equals two for the straight line envelope. Also, the ultimate yield surface

or envelope that represents the asymptotic stress states will be a non-linear envelope with

slope equal to = u2 / nq. For both of these cases, the parameters are shown as follows:

Value q assumed to be equal to two

Friction angle and adhesion

Figure 5-10 shows ultimate shear stress verses normal shear stress, and best linear curve

fitting are shown in following figure. Based on the equation from best curve fitting, the

interface friction angle is equal to 41 degrees, and cohesion is equal to 58 kPa.

Backfill and sample A

1200
y = 0.8833x + 58.04
Shear Stress (kPa)

1000
R = 0.9922
800
600
400
200
0
0 200 400 600 800 1000 1200
Normal Stress (kPa)

Figure 5-10 Ultimate shear stress vs. Normal stress.

Determination parameters q and


238

The ultimate yield surface that represents the asymptotic stress states will be an average

straight line with slope equal to the square root of gamma, where q is a parameter. Its

value equals two for the straight line envelope, as noted before is zero at the ultimate

shear stress. The ultimate shear stress of the intact behavior is unknown. The ultimate

shear stress of the intact behavior can be taken as a value approximately ten percent

higher than the ultimate or peak shear stress from the shear test. Figure 5-11 shows the

square ultimate shear stress vs. square normal stress. It shows the best linear curve fitting

with a slope of 1.0658, which is the value of .

Backfill and Sample A with 10% increase in


ultimate shear
1.4E+06
1.2E+06
y = 1.0658x
1.0E+06 R = 0.985
(kPa)

8.0E+05
6.0E+05
4.0E+05
2.0E+05
0.0E+00
0.0E+00 5.0E+05 1.0E+06 1.5E+06
n (kPa)

Figure 5-11 1.1*2 vs. n2.

The phase change parameter (n)

The phase change parameter (n) is based on the state of stresses when the interface

transitions from compaction to dilation behavior. The line connecting the peaks of the
239

yield envelope is called the phase change line. At these points of maximum value dF /

dn is equal to zero. The value of parameter, n, can be determined as follows:

n= q/ (1-t/n) (5-19)

Shear stress in this equation is shear stress at the transition, and Figure 5-12 shows

transition shear stress vs. normal stress, and the slope of the line is equal to 0.5257 which

is t / n. By knowing this information; the phase change parameter is equal to 3.95.

Backfill and Sample A at Transition


y = 0.5257x
600 R = 0.9963
Shear Stress (kPa)

500
400
300
200
100
0
0 200 400 600 800 1000 1200
Normal Stress (kPa)

Figure 5-12 Transition shear stress vs. normal stress


240

Determination of , a, and b

To determine the growth function (), the hardening parameters (a and b) need to be

defined. As previously mentioned, the parameter b notes the rate of hardening which is

the slope of the average curve and a is denoted as the value of when is equal to one

which is the intercept when ln is equal to zero. The following figures show the ln vs.

ln for calculating hardening parameters.

n=70kPa
ln
6
8.6 8.5 8.4 8.3 8.2 8.1 8 7.9
7

y=4.1364x 43.5 8
R=0.8352

ln
9

10

11

12

Figure 5-13 Determination of hardening parameters for n=70 kPa.


241

n =140kPa
ln
9.4
9 8.8 8.6 8.4 8.2 8 9.6 7.8
9.8
y=1.5802x 23.456
10
R=0.9281
10.2

ln
10.4
10.6
10.8
11
11.2
11.4

Figure 5-14 Determination of hardening parameters for n=140 kPa.

n =210kPa
ln 10
8.8 8.6 8.4 8.2 8 7.8
10.5
y=2.0854x 28.476
R=0.9111
11
ln

11.5

12

12.5

Figure 5-15 Determination of hardening parameters for n=210 kPa.


242

n=350kPa
ln
11.2
8.8 8.6 8.4 8.2 8 11.4 7.8
11.6
y=1.6097x 25.429 11.8
R=0.9189
12

ln
12.2
12.4
12.6
12.8
13

Figure 5-16 Determination of hardening parameters for n=350 kPa.

n=525kPa
ln
11.5
9 8.8 8.6 8.4 8.2 8 7.8 7.6 7.4 7.2
12
y=1.329x 23.657
R=0.9324 12.5
ln

13

13.5

14

14.5

Figure 5-17 Determination of hardening parameters for n=525 kPa.


243

n=875kPa
ln
13
8.8 8.6 8.4 8.2 8 7.8 7.6 7.4 7.2
13.2
y=0.6797x 18.976
R=0.952 13.4

ln
13.6

13.8

14

14.2

Figure 5-18 Determination of hardening parameters for n=875 kPa.

n=1050kPa
ln
13.2
8.8 8.6 8.4 8.2 8 7.8 7.6 7.4 7.2
13.4

y=0.9302x 21.442 13.6


R=0.942 13.8
ln

14
14.2
14.4
14.6
14.8

Figure 5-19 Determination of hardening parameters for n=1050 kPa.

After averaging the values of b and a from the entire test, the rate of hardening is equal to

1.764 and when =1, is equal to 9.07 E-10.


244

Material Constants Symbol Value

kn kn= -0.25882 + 986.96 + 129355


Elastic Constants
ks ks = -0.2096 n + 432.95 n + 95299

1.0658

Plasticity Constants 0

q 2

Phase Change Parameter n 3.95

a 9.07E-10
Growth Parameters
b 1.764

Table 5-3 Material constants for soil-FRP interface.

Calculated the actual value of q and following material parameters

Determination parameters q and

The parameter gamma is determined from the slope of the ultimate envelope in shear

stress vs. normal stress space. The value of q is found by Equation (5-17), which the plot

of ln u vs. ln n provides the value of q from the slope of the average line. When natural

log of normal stress is equal to zero, gamma can be determined. According to Figure 5-20,

the value of and q are 23.42 and 1.5386, respectively. Note, the ultimate shear stress of

the intact behavior can be taken as a value about ten percent higher than the ultimate or

peak shear stress from the shear test.


245

lnu vs.lnn
7.5
7 y=0.7693x+1.5768
Ln(ultimateshearstress)

R=0.9945
6.5
6
5.5
5
4.5
4
2 3 4 5 6 7 8
Ln(normalstress)

Figure 5-20 Plot for determination of and q.

The phase change parameter (n)

The phase change parameter (n) is based on the state of stresses when the interface

transits from compaction to dilation behavior. The value of parameter, n, can be

determined using Equation (5-19). The slope of shear and normal stresses at the

transition point is determined from Figure 5-21, which is equal to 4.5, and then the value

of n is equal to 1.9.
246

200000 At Transition

160000
t (KPa)

120000

80000 y = 4.4967x
R = 0.9952
40000

0
0 10000 20000 30000 40000
qnt(KPa)

Figure 5-21 Shear stress vs. normal stress at transition.

Determination of , a, and b

To determine the growth function (), the hardening parameters (a and b) need to be

defined. As mentioned before, The parameter b denotes rate of hardening which is the

slope of average curve and a is denoted as the value of when is equal to one which I

thes intercept when ln is equal to zero. The following figures show the ln vs. ln for

calculating hardening parameters.


247

NormalStress70kPa
1.6

1.5

1.4

ln
1.3

1.2
y=0.46x 2.408
R=0.976 1.1

1
8.6 8.5 8.4 8.3 8.2 8.1 8 7.9
ln

Figure 5-22 Determination of hardening parameters for n=70 kPa (q=1.54).

NormalStress140kPa

1.4
1.3
1.2
ln

1.1
y=0.4416x 2.562 1
R=0.9847 0.9
0.8
0.7
0.6
9 8.8 8.6 8.4 8.2 8 7.8
ln

Figure 5-23 Determination of hardening parameters for n=140 kPa (q=1.54).


248

NormalStress210kPa
1.2

0.8

ln
y=0.669x 4.674 0.6
R=0.9784
0.4

0.2
8.8 8.6 8.4 8.2 8 7.8
ln

Figure 5-24 Determination of hardening parameters for n=210 kPa (q=1.54).

NormalStress350kPa

1.2

0.8
ln

0.6

0.4
y=0.8165x 6.1648
R=0.9608 0.2

0
8.8 8.6 8.4 8.2 8 7.8
ln

Figure 5-25 Determination of hardening parameters for n=350 kPa (q=1.54).


249

NormalStress525kPa

1
0.8
0.6

ln
0.4
0.2

y=0.8633x 6.6585 0
R=0.9683 0.2
0.4
9 8.5 8 7.5 7
ln

Figure 5-26 Determination of hardening parameters for n=525 kPa (q=1.54).

NormalStress875kPa

0.8

0.6

0.4
ln

0.2
y=0.7209x 5.5268 0
R=0.9489
0.2

0.4
8.8 8.6 8.4 8.2 8 7.8 7.6 7.4 7.2
ln

Figure 5-27 Determination of hardening parameters for n=875 kPa (q=1.54).


250

NormalStress1050kPa
1.2

0.8

0.4

ln
0
y=1.1842x 9.5064
R=0.9249 0.4

0.8

8.8 8.6 8.4 8.2 8 7.8 7.6 7.4 7.2 1.2


ln

Figure 5-28 Determination of hardening parameters for n=1050 kPa (q=1.54).

After averaging the value of b and a from the entire test, the rate of hardening is equal to

0.7365 and when =1, is equal to 0.0263.


251

MaterialConstants Symbol Value

2
kn Kn= -0.2588n + 986.96 n+ 129355
ElasticConstants
2
ks Ks = -0.2096 n + 432.95 n + 95299

23.4

PlasticityConstants 0

q 1.5386

PhaseChangeParameter n 1.9

a 0.0263
GrowthParameters
b 0.7365

Table 5-4 Material constants for soil-FRP interface

5.3.2.1. A Comparison between q=1.54 and q=2

All the parameters corresponding to each q from Table 5-3 and Table 5-4 are put in

Equation (5-17) to compare the lab results with the two models as shown in Figure 5-29.

Equation (5-15) is used to compare ultimate shear stress and normal stress for two

different qs and the lab results are as shown in Figure 5-30.


252

7.5
7
6.5
6
5.5
Ln(u)

LabTest
5
q=1.5386
4.5
q=2
4
3.5
3
2 3 4 5 6 7 8
Ln(n)

Figure 5-29 Comparison between natural log of u vs. n of two different qs and lab results.

1200

1000

800
u (kPa)

600 LabTests
Theoryw/q=1.54
400
Theoryw/q=2
200

0
0 500 1000 1500
n (kPa)

Figure 5-30 Comparison between u vs. n of two different qs and lab results.
253

5.3.3. Computer Programming for Back Prediction

After obtaining the HISS model parameter, a computer program using MATLAB is

operated for back predicting the shear stress vs. shear deformation for interface between

CFRP-Sand. The procedure is as follow:

Derivative respect to , n, and from yield function (Equation (5-9)):

df/d=2 (5-20)

df/n=n(n +R)n+1 q(n +R)q-1 (5-21)

df/d=df/d.d/d=-ba(n +R)n/ b+1 (5-22)

Define Ce in matrix form

(5-23)

Define dF/d in matrix form

(5-24)

Define Ci in matrix form



(5-25)

254

To model a direct shear test using HISS model, first normal stress increment is
applied, and corresponding normal deformation is calculated using Equation
(4-17), with no shear stress until reaching the desired normal stress. Then, shear
stress increment is applied, and corresponding shear deformation is calculated,
with constant normal stress.

The computer program is as follow,

gamma=1.0658

b=1.764

a=9.077*10^(-10)

n=3.95

c=39.915

r=c/(gamma)^0.5;

kn=(-0.2588*0^2+986.96*0+129355);

ks=-0.2096*0^2+432.95*0+95200;

ce=[ks 0;0 kn];

ci=ce;

dsigman=0.5;

kesi=0.0000001;

tav=0;

counter=1;

uvect=zeros(2,1);

sigmahist=0;

uvecthist=0;

for sigman=0:dsigman:1050

uvecthist(counter,1)=uvect(1);

uvecthist(counter,2)=uvect(2);

sigmahist(counter,1)=0;
255

sigmahist(counter,2)=sigman;

if counter>1

alfa=(gamma*(sigman+r)^2-tav^2)/((sigman+r)^n);

kesi=(a/alfa)^(1/b)

kn=(-0.2588*sigman^2+986.96*sigman+129355);

ks=-0.2096*sigman^2+432.95*sigman+95200;

ce=[ks 0;0 kn];

dfds=[dfdtav(tav);dfdsigman(a,n,kesi,b,sigman,gamma,r)];

ci=ce-(ce*dfds*dfds'*ce)/(dfds'*ce*dfds-

dfdkesi(sigman,n,b,a,kesi,r)*(dfds'*dfds)^.5);

end

duvect=inv(ci)*[0;dsigman];

uvect=uvect+duvect;

dup=duvect(1)-0/ks;

dvp=duvect(2)-dsigman/kn;

counter=counter+1;

end

figure(1)

plot(uvecthist(:,2),sigmahist(:,2))

%figure(2)

%plot(uvecthist(:,1),sigmahist(:,1))

du=0.00001;

sigmahist=0;

uvecthist=0;

sigmavect=[0;sigman];
256

counter=1

dtav=1

for u=0:du:0.01

uvecthist(counter,1)=uvect(1);

uvecthist(counter,2)=uvect(2);

sigmahist(counter,1)=tav;

sigmahist(counter,2)=sigman;

if counter>1

alfa=(gamma*(sigman+r)^2-tav^2)/((sigman+r)^n);

kesi=(a/alfa)^(1/b)

kn=(-0.2588*sigman^2+986.96*sigman+129355);

ks=-0.2096*sigman^2+432.95*sigman+95200;

ce=[ks 0;0 kn];

dfds=[dfdtav(tav);dfdsigman(a,n,kesi,b,sigman,gamma,r)];

ci=ce-(ce*dfds*dfds'*ce)/(dfds'*ce*dfds-

dfdkesi(sigman,n,b,a,kesi,r)*(dfds'*dfds)^.5);

end

dsigmavect=ci*[du;0];

dsigmavect(1)=(-ci(1,2)*ci(2,1)/ci(2,2)+ci(1,1))*du;

%dv=-ci(2,1)*

%sigman=sigman+dsigmavect(2);

tav=tav+dsigmavect(1);

uvect(1)=uvect(1)+du;

%dup=duvect(1)-dtav/ks;

%dvp=duvect(2)-0/kn;

counter=counter+1;
257

end

figure(3)

plot(uvecthist(:,1),sigmahist(:,1))

Back prediction for q=2

After defining all the HISS model parameters as shown in

Material Constants Symbol Value

kn kn= -0.25882 + 986.96 + 129355


Elastic Constants
ks ks = -0.2096 n + 432.95 n + 95299

1.0658

Plasticity Constants 0

q 2

Phase Change Parameter n 3.95

a 9.07E-10
Growth Parameters
b 1.764

Table 5-3, HISS model parameters for q=2.

The above computer programming is used for back predicting shear stress vs. shear

deformation for q=2. The following figures compare lab results with HISS-0 model.
258

Normal Stress 17.5 kPa


70

60
Shear Stress (kPa)

50

40

30 Observed
Model
20

10

0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 5-31 Lab data and prediction of interface test (n=17.5 kPa).

Normal Stress 35 kPa


80
70
Shear Stress (kPa)

60
50
40
Observed
30
Model
20
10
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 5-32 Lab data and prediction of interface test (n=35 kPa).
259

Normal Stress 70 kPa


120

100
Shear Stress (kPa)

80

60
Observed
40 Predicted

20

0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 5-33 Lab data and prediction of interface test (n=70 kPa).

Normal Stress 140 kPa


200
180
160
Shear Stress (kPa)

140
120
100
Observed
80
60 Model
40
20
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 5-34 Lab data and prediction of interface test (n=140 kPa).
260

Normal Stress 210 kPa


300

250
Shear Stress (kPa)

200

150
Observed
100 Model

50

0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 5-35 Lab data and prediction of interface test (n=210 kPa).

Normal Stress 350 kPa


450
400
350
Shear Stress (kPa)

300
250
200 Observed
150 Model
100
50
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 5-36 Lab data and prediction of interface test (n=350 kPa).
261

Normal Stress 525 kPa


600

500
Shear Stress (kPa)

400

300
Observed
200 Model

100

0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 5-37 Lab data and prediction of interface test (n=525 kPa).

Normal Stress 875 kPa


900
800
700
Shear Stress (kPa)

600
500
400 Observed
300 Model
200
100
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 5-38 Lab data and prediction of interface test (n=875 kPa).
262

Normal Stress 1050 kPa


1000
900
800
Shear Stress (kPa)

700
600
500
Onserved
400
300 Model
200
100
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 5-39 Lab data and prediction of interface test (n=1050 kPa)
263

Back prediction for q=1.54

Based on q=1.54, the following table shows HISS models parameters.

Material Constants Symbol Value

kn kn= -0.25882 + 986.96 + 129355


Elastic Constants
ks ks = -0.2096 n + 432.95 n + 95299

1.0658

Plasticity Constants 0

q 2

Phase Change Parameter n 3.95

a 9.07E-10
Growth Parameters
b 1.764

Table 5-3 HISS model parameters for q=1.54.

The same computer program is used for back prediction of shear stress vs. shear

deformations for different normal stresses. The following figures compare lab results

with HISS-0 model q=1.54.


264

Normal Stress 17.5 kPa


60

50
Shear Stress (kPa)

40

30
Observed
20 Model

10

0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 5-40 Lab data and prediction of interface test (n=17.5 kPa).

Normal Stress 35 kPa


80
70
Shear Stress (kPa)

60
50
40
Observed
30
Model
20
10
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 5-41 Lab data and prediction of interface test (n=35 kPa).
265

Normal Stress 70 kPa


140

120
Shear Stress (kPa)

100

80

60 Observed
Model
40

20

0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 5-42 Lab data and prediction of interface test (n=70 kPa).

Normal Stress 140 kPa


250

200
Shear Stress (kPa)

150

Observed
100
Model
50

0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 5-43 Lab data and prediction of interface test (n=140 kPa).
266

Normal Stress 210 kPa


300

250
Shear Stress (kPa)

200

150
Observed
100 Prediction

50

0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 5-44 Lab data and prediction of interface test (n=210 kPa).

Normal Stress 350 kPa


450
400
350
Shear Stress (kPa)

300
250
200 Observed
150 Model
100
50
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 5-45 Lab data and prediction of interface test (n=350 kPa)
267

Normal Stress 525 kPa


600

500
Shear Stress (kPa)

400

300
Observed
200 Model

100

0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 5-46 Lab data and prediction of interface test (n=525 kPa).

Normal Stress 875 kPa


900
800
700
Shear Stress (kPa)

600
500
400 Observed
300 Model
200
100
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 5-47 Lab data and prediction of interface test (n=875 kPa).
268

Normal Stress 1050 kPa


1000
900
800
Shear Stress (kPa)

700
600
500
Observed
400
300 Model
200
100
0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 5-48 Lab data and prediction of interface test (n=1050 kPa).

For independent validations, interface shear tests were performed under two different

initial normal stresses such as 280 kPa and 700 kPa on backfill and sample A, with the

final horizontal displacement of approximately 10 mm. These tests are conducted at a

displacement rate of 0.02 mm/sec. Then, the test data were compared with the HISS

model (q=1.54) to determine if the HISS parameters are suitable for general interface

behavior which involve CFRP and backfill. Based on Figure 12 (a and b), it can be

concluded that predictions from the HISS model provide very good correlation with the

test behavior.
269

Normal Stress 280 kPa


350
300
ShearStress(kPa)

250
200
150
Observed
100
HISSModel
50
0
0 2 4 6 8 10
ShearDisplacement(mm)

Figure 5-49 Independent laboratory tests and predictions of the HISS model with n=280 kPa.

Normal Stress 700 kPa


700

600
Shear Stress (kPa)

500

400

300
Observed
200 HISS Model
100

0
0 2 4 6 8 10
Shear Displacement (mm)

Figure 5-50 Independent laboratory tests and predictions of the HISS model with n=700 kPa.
270

Summary

In this chapter, the HISS model is used to back predict the shear stress vs. shear

deformation. Two different q values are compared; q=2 is used for linear relationship

between normal stress vs. shear stress. q=1.5 is used for nonlinear relationship between

normal stress vs. shear stress. It can be concluded that using q=1.5 is a better predictor

between shear stress vs. shear deformation. Next chapter, a case study is performed and

Plaxis is used to create a hardening model. For independent validations, interface shear

tests were performed under two different initial normal stresses such as 280 kPa and 700

kPa on backfill and sample A Then, the test data were compared with the HISS model

(q=1.54) to determine if the HISS parameters are suitable for general interface behavior

which involve CFRP and backfill. It can be concluded that predictions from the HISS

model provide very good correlation with the test behavior.


271

CHAPTER 6. NUMERICAL MODELING OF MSE WALL USING PLAXIS

6.1. Introduction

Limit equilibrium methods and finite element methods are two methods two that are

commonly used to design reinforced soil structures. Limit equilibrium is an easy

approach that has been used for years to design reinforced soil structures. Nevertheless,

the limit equilibrium design approach does not account for deformations or stress

distributions in soil or reinforcement (Rowe and Ho, 1992). In addition, limit equilibrium

methods underestimate the failure of two geosynthetics reinforced soil walls using

granular and cohesive backfills (Wu, 1992).

Finite element analysis has been widely used in geotechnical problems to create a

distribution of strains and stresses by Seed et al., 1986; Adib et al., 1990; Chew et al.,

1990; Bathurst et al., 1992; Pal, 1997, and Desai and El-Hoseiny, 2005. The advantage

of using Finite Element analysis is the ability to provide information for design purposes

and use in creating and confirming design methods. Nevertheless, the finite element

method requires calibration by some type of observed behavior. The simulation of

reinforced soil structures using finite element analysis can take into account the stress

deformation behavior and modeling discontinuous behavior at the interface between soil

and reinforcement.

In this study, a CFRP reinforced earth retaining wall has been chosen for the numerical

simulation and the validation of the finite element models using Plaxis Hardening-Soil

model (HS).
272

6.2. Mechanical Stabilized Earth Walls (MSE)

Mechanically stabilized earth or MSE is soil constructed with reinforcement. MSE walls

can be used for retaining walls, bridge abutments, dams, seawalls, and dikes. The basic

ideas of MSE have been used throughout history and MSE was developed in its current

form in the 1960s. The elements used for reinforcing can differ but they include steel and

geosynthetics. The first MSE walls date to between 5000 B.C and 2500 B.C. The

Babylonians constructed ziggurats (large towers) of earth that were reinforced with

horizontal sheets of reed laid at regular intervals. The most notable remains of these

ziggurats are those at Aqar Quf near Baghdad, Iraq. The overall height of the ziggurats

were about 87 m (Retaining wall, 2012). MSE walls have become very popular mainly

due to their flexibility, large bases, and cost effectiveness compared to rigid retaining

walls. MSE walls are used for an array of retaining structures. Metal strips, geotextiles,

or geogrids are used to reinforce the soil mass. Other names for the materials that are

used for reinforcement are geosynthetics and geocomposities. A geotextile is any

permeable textile material used in geotechnical applications. Geosynthetics are

polymeric, planar materials. Geocomposite is a product made from a combination of

geosynthetics, and a geogrid is a polymeric product formed by joining intersecting ribs

(Budhu, 2008). The first MSE wall in the United States was constructed in 1971 on State

Route 39 near Los Angeles, California. Since 1997, approximately 23,000 MSE walls

have been constructed in the world. The highest MSE wall built in the United States is 30

m (98 ft) high (Retaining wall, 2012).


273

6.3. Numerical Modeling

Plaxis is a two-dimensional finite element computer program used to perform

deformation and stability analyses for different types of geotechnical applications. Real

situations may be modeled either by a plane strain or a axisymmetric model. Plaxis uses

a graphical user interface which allows users to rapidly generate a geometry model and

finite element mesh based on representative vertical cross section of the (Plaxis, 2010).

Plaxis uses different models, and the main models for simulation for this research are

Mohr coulomb model (MC) and Hardening soil model (HS). The elastic plastic Mohr-

Coulomb model uses five input parameters, which are E and v for soil elasticity; and c

for soil plasticity and as an angle of dilatancy. This Mohr coulomb model represents a

first order analysis of soil or rock behavior. It is recommended to use this model for a

first analysis of the problem considered. For each layer onward Plaxis estimates a

constant average stiffness. Due to this constant stiffness, computations tend to be

relatively fast and one obtains a first impression of deformations.

The hardening soil model is an intricate model for the simulation of soil behavior. As for

the Mohr Coulomb model, limiting states of stress are described by means of the friction

angle, cohesion and dilatancy angle. Soil stiffness is described much more correctly by

using three unlike inputs for stiffness: the triaxial loading stiffness, E50, the triaxial

unloading stiffness, Eur, and oedometer loading stiffness, Eoed. In contrast to the Mohr-

Coulomb model, the hardening soil model also utilizes the stress-dependency of the

stiffness moduli. This means that all stiffnesses increase with pressure. Therefore, all
274

three input stiffnesses relate to a reference stress, being usually 100 kPa. However, the

hardening model does not account for softening due to soil dilatance and debonding

effects. In fact, it is an isotropic hardening model so that it models neither hysteretic and

cyclic loading nor cyclic mobility. The details below are adopted from Plaxis manual

The Hardening soil (HS) model uses two failure criteria and plastic potentials. A family

of Mohr- Coulomb criterion for deviatoric failure and an elliptic cap for volumetric

failure are used in HS model. Some properties of the HS model is as follow:

- Stress dependent stiffness according to a power law with input parameter m

- Plastic straining due to primary deviatoric loading with input parameter

- Plastic straining due to primary compression with input parameter

- Elastic unloading / reloading with input parameters , vur

- Failure according to the Mohr-Coulomb model with parameters c, , and

A basic idea for the formulation of the Hardening-Soil model is the hyperbolic

relationship between the vertical strain and the deviatoric stress, q, in primary triaxial

loading. Standard drained triaxial tests tend to yield curves that can be described by:

(6-1)

The ultimate deviatoric stress, , and the quantity in Equation (6-1) are defined as:

(6-2)
275

The above relationship for is derived from the Mohr-Coulomb failure criterion, which

involves the strength parameters c and . When q= , the failure criterion is satisfied

and perfectly plastic yielding occurs. The failure ratio Rf is ratio between and

which is often equal to 0.9.

The stress strain behavior for primary loading is highly nonlinear. The parameter is

the confining stress dependent stiffness modulus for primary loading. Instead of the

initial modulus, , for small strain, is used as a tangent modulus (Figure 6-1). It is

given by the equation:

(6-3)

is a reference stiffness modulus corresponding to the reference stress pref. The

actual stiffness depends on the minor principal stress, , which is the effective confining

pressure in a triaxial test, and the amount of stress dependency is given by the power m.

Janbu (1963) reported values of m around 0.5 for Norwegian sands and silts. As secant

modulus is determined from a triaxial stress-strain curve for a mobilization of 50%

of the maximum shear strength qf.

For unloading and reloading stress paths, another stress-dependent stiffness modulus is

used.

(6-4)

is a youngs modulus for unloading and reloading (Figure 6-1); hence, E is the

reference Youngs modulus for unloading and reloading to the reference pressure .
276

is a tangent stiffness modulus; Hence, E is a tangent stiffness at vertical stress of

Figure 6-1 Hyperbolic stress-strain relation in primary loading for a standard drained triaxial test
(Plaxis manual, 2006).

(6-5)
277

Figure 6-2 Definition of Eoedreff in oedometer test results (Plaxis manual, 2006).
For the triaxial case the two yield functions and are defined according to Equation

(6-6) and (6-7). Here the measure of the plastic shear strain according to Equation

(6-8) is used as the relevant parameter for frictional hardeining:

(6-6)

(6-7)

With the definition

(6-8)

In reality, plastic volumetric strains will never be precisely equal to zero, but for hard

soils plastic volume changes tend to be small when compared with the axial strain, so that
278

the approximation in Equation (6-8) will be generally be accurate. For a given constant

value of the hardening parameter, , the yield condition 0 can be visualized

in plane by means of a yield locus. When plotting such yield loci, one has to use

Equations (6-6) and (6-7) as well as Equations (6-3) and (6-4) for and

respectively.

Shear yield surfaces do not explain the plastic volume strain that is measured in isotropic

compression. A second type of yield surface must therefore be introduced to close the

elastic region in the direction of the p-axis. The triaxial modulus largely controls the

shear yield surface and the oedometer modulus controls the cap yield surface. The

definition of the cap yield surface is as follow (Plaxis, 2006):

(6-9)

Where is an auxiliary model parameter that relates to which is the value for normal

consolidation. p and are related to mean and devitoric stresses.

The yield surface of a hardening plasticity model can expand due to plastic straining

(Figure 6-3). Distinction is made between two main types of hardening, namely shear

hardening and volumetric hardening. Shear hardening is used to model irreversible strains

due to deviatoric loading. Compression hardening is used to model irreversible plastic

strains due to compression in oedometer loading and isotropic loading (Schanz, 1998).

Further information about the Hardening model can be found in Plaxis material manual

(Plaxis, 2006).
279

Figure 6-3 Representation of total yield contour of the Hardening soil model in principal stress space
for cohesionless soil (Plaxis manual, 2006).

Interface elements are typically modeled by using a bilinear Mohr-Coulomb model.

When a more advanced model is used for the matching cluster material data set, the

interface element will only pick the relevant data for the Mohr Coulomb model.

Therefore the interface stiffness is taken to be the elastic soil stiffness. Hence, E=Eur

where Eur is stress level dependent, following a power law with Eur proportional to m

(Plaxis Manuals, 2010).

Interface elements are often modeled by means of the bilinear Mohr-Coulomb model. In

such case, the interface stiffness is taken to be the elastic stiffness. Hence, E= , where

is stress level dependent, following a power law with proportional to (Plaxis,


280

2006). The Mohr-Coulomb yield condition is an extension of Coulombs friction law to

general state of stress. This condition ensures that Coulombs friction law is obeyed in

any plane within a material element. The full Mohr-Coulomb yield condition consists of

six yield functions when formulated in terms of principal stresses (Smith and Griffith,

1982). The two plastic model parameters appearing in the yield functions are the well-

known friction angle and cohesion c. The yield functions together represent a

hexagonal cone in principal stress space. An elastic-plastic model is used to describe the

behavior of interfaces for the modeling of soil-structure interaction. The coulomb

criterion is used to distinguish between elastic behavior, where small displacements can

occur within the interface, and plastic interface behavior when permanent slip may occur.

For the interface to remain elastic the shear stress is given by:

| | (6-10)

and for plastic behavior is given by:

| | (6-11)

where and are the friction angle and cohesion of the interface. The strength

properties of interfaces are linked to the strength properties of a soil layer. Each data set

has an associated strength reduction factor for interfaces ( ). The geogrid

reinforcement is simulated by a linear elastic - perfectly plastic model (Plaxis, 2006).


281

6.4. MSE Wall in Tucson Arizona

Before modeling a case study using CFRP for reinforcing soil, the model needs to

calibrated and in order to ensure that it will provide reasonable results. Therefore, a MSE

wall (located in Tucson, AZ) is modeled in Plaxis using Hardening soil model, since the

wall has sensors at different depths of soil to provide the vertical stress, horizontal and

vertical strains. The results of the model can be compared to the results of the sensors.

Forty three MSE walls were constructed at Tanque Verde Road to grade separated

interchanges on the Tanque Verde Wrightstown-Pantano roads projects in Tucson, AZ

between November 1984 and October 1985. The project used geogrid reinforcement in

mechanically stabilized earth (MSE) retaining walls in a large transportation related

project in North America (Tensar, 1984). In this study, wall panel 26-32 is simulated

using finite element method analysis. This panel is a two dimension instrumented wall

and can be used for comparing analytical results with the instrumented filed data. The

wall system was considered as an alternative to a conventional cantilevered retaining wall

system or a reinforced earth retaining wall system utilizing Tensar SR2 as reinforcement.

The wall height is 16 ft, and the reinforced soil mass has 6 in thick and 10 ft wide precast

reinforced concrete panels. Soil reinforced geogrids were mechanically connected to the

concrete facing panels at the elevations shown in Figure 6-4 and reached to a length of 12

ft. At the top of the wall fill, a pavement structure is constructed that is composed of 4 in

base coarse covered by 9.5 in of Portland cement concrete. More details of the

configuration are given by Berg et al., 1986.


282

Figure 6-4 Location of instruments for wall panel 26-32 (Berg et al., 1986).

The soil reinforcement used was Tensars SR2 structural geogrid. The Tensar SR2

geogrid is a uniaxial product that is made from high density polyethylene (HDPE)

stabilized with about 2.5% carbon black to provide resistance to attack by ultraviolet light

(Tensar, 1984). It is known to have resistance to chemical substances that exist in soils

(Fishman and Desai, 1991). The geogrids have a maximum tensile strength of 5400 lb/ft
283

and secant modulus in tension at 2 % elongation of 75000 lb/ft. The allowable long term

tensile strength based on creep considerations is reported to be 1986 lb/ft at 10% strain

after 120 years. The foundation soil was excavated to a depth equal to the lesser of the

particular wall panel height and compacted to at least 90% of maximum dry density

based on the proctor test. As backfill proceeded, geogrids were secured to the wall face

at specified elevations and pretensioned. Granular fill was placed and spread up to the

first geogrid level by a front end loader and then compacted. Compaction near the wall

face was carried out by hand-operated, lightweight vibratory device. Further from the

wall, an operator driven, 1 ton, vibratory steel drum roller was utilized.

6.5. Comparing the Plaxis (HS) Results with Field Results

A hardening model in Plaxis is used to model the soil behind the panels 26-32 in an MSE

wall in Tucson. In contrast to an elastic perfectly plastic model, the yield surface of a

hardening plasticity model is not fixed in principal stress space, but it can expand due to

plastic straining. A distinction can be made between two types of hardening, namely

shear hardening and compression hardening. Shear hardening is used to model

irreversible strains due to primary deviatoric loading. Compression hardening is used to

model irreversible plastic strains due to primary compression in oedometer loading and

isotropic loading. Both types of hardening are contained in the present model. The

hardening soil model is an advanced model for simulating the behavior of different types

of soil, both soft soils and stiff soils (Schanz, 1998). When subjected to primary

deviatoric loading, the soil shows a decreasing stiffness and also irreversible plastic
284

strains develop. In a drained triaxial test, the observed relationship between the axial

strain and the deviatoric stress can be approximated by a hyperbola. This relationship was

first formulated by Kondner (1963) and later used in the hyperbolic model (Duncan and

Chang, 1970). The hardening soil model is a much better model by far in comparison to

the hyperbolic model. First, by using the theory of plasticity rather than the theory of

elasticity. Second, by including soil dilatancy and thirdly by introducing a yield cap.

All the parameters used are found from the lab results and from the manufacturer and

they are inserted into the Plaxis program as shown in Table 6-1. The parameters for

backfill soil obtained from El-Hoseiny (1999) triaxial results. Then the results from

Plaxis were compared with data from the field. Field data was collected for about two

years, and dissimilar values were collected due to change is temperature. The minimum

and maximum values of the field data are shown in the following figures. For the

purposes of modeling, the side boundaries were placed at a distance of 2.5 times the

length of the reinforcement, and the bottom boundary was placed at a distance of 3.125

times the height of the wall. Such distances and the assumed boundary conditions are

considered to approximately simulate the semi-infinite extent of the system as shown in

Figure 6-5. In the report, it is mentioned that at the job site the concrete near the wall

was not straight and it is possible that the data close to wall does not provide accurate

results.
285

a) Parameters for backfill and interface

Interface Interface
MaterialContant Symbol Backfill (Tensar) (CFRP)
Adhesion(kPa) C 66 58
Frictionangle/interfaceangle(degree) / 40 34 41.5
Secantstiffnessinstandarddrained
triaxialtest(MPa) E50(ref) 42.8
Tangentstiffnessforprimaryoedometer
loading(MPa) Eoed(ref) 34.3
Unloading/reloadingstiffness(MPa) Eur(ref) 399.8
PowerforStressleveldependencyof
stiffness M 0.5
Lateralcoefficient(normalconsolidation) K0nc 0.4
Failureratio Rf 0.9
Referencestressforstiffnesses(kPa) pref 100

b) Parameters for geogrid and concrete wall

MaterialContant Symbol Tensar CFRP ConcretePanel


Axialstiffness(KN/m) EA 2250.4 11675.1*
Maxaxialtensionforce(KN/m) Np 19.0 186.8*
Thickness(mm) T 4.572 0.635 152.2
Tensile/compressionmodulus
E 1.50E+06 6.09E+07 2.10E+07
ofelasticity(kPa)
Poissonratio V 0.3 0.34 0.15
*EA and Np are reduced by factor of 0.4, since 0.4 ft width of CFRP required for one foot
per length.
Table 6-1 Parameters for Plaxis analysis.
286

a) Boundary condition of the MSE wall b) Detail of the MSE wall

Figure 6-5 FE Mesh for Tucson MSE wall panel 26-32 in Plaxis.

Elevation:5ft
16
14
VerticalStress(psi)

12
10
8 minfielddata
6 maxfielddata
4 Plaxis(HS)
2
0
0 50 100 150
DistancefromWallFace(in)

Figure 6-6 Vertical stress vs. distance from wall face for panel 26-32 face at elevation 5.0 ft.
287

Elevaton:3.5ft
0.9
0.8
0.7
VertSoilStrain(%)

0.6
0.5
VertSoilStrain(Min)
0.4
0.3 VertSoilStrain(Max)
0.2 TensarSR2
0.1
0
0 10 20 30 40
DistanceFromWallFace(in)

Figure 6-7 Vertical soil strain vs. distance from wall for panel 26-32 face at elevation 3.5 ft.

Elevation:0.5ft
0.8
0.7
GeogridStrain(%)

0.6 minBisonGages
fielddata
0.5
0.4 foiltyperesistance
straingagefielddata
0.3
MaxBisonGages
0.2
fielddata
0.1
Plaxis(HS)
0
0 20 40 60 80
DistanceFromWallFace(in)

Figure 6-8 Geogrid strain vs. distance from wall for panel 26-32 face at elevation 0.5 ft.
288

Elevation:1.5feet
0.4
0.35
GeogridStrain(%)

0.3
0.25
0.2 mindatafield
0.15 maxdatafield
0.1 Plaxis(HS)
0.05
0
0 50 100 150
DistFromWall(in)

Figure 6-9 Geogrid strain vs. distance from wall for panel 26-32 face at elevation 1.5 ft.

Elevation:4.5ft
0.45
0.4
0.35
GeogridStrain(%)

0.3
0.25
minfielddata
0.2
0.15 maxfielddata
0.1 Plaxis(HS)
0.05
0
0 20 40 60 80 100
Distfromwall(in)

Figure 6-10 Geogrid strain vs. distance from wall for panel 26-32 face at elevation 4.5 ft.
289

Elevation:11.5ft
0.6

0.5
GeogridStrain(%)

0.4

0.3 maxfielddata

0.2 mindatafield
Plaxis(HS)
0.1

0
0 50 100 150
Distfromwall(in)

Figure 6-11 Geogrid strain vs. distance from wall for panel 26-32 face at elevation 11.5 ft.

Elevation: 8 ft
1.6
1.4
1.2
Soil Strain (%)

1
0.8 min field data
0.6 max field data

0.4 Plaxis

0.2
0
0 50 100 150
Distance from Wall (in)

Figure 6-12 Soil strain vs. distance from wall for panel 26-32 face at elevation 8ft.
290

From above figure, it can be concluded that the result from the field and the result from

HS modeling are slightly different. The error can be due to the sensors not perfectly

installed, error in reading due to temperature, and the HS modeling has limitation as

mentioned before. However, by comparing the HS modeling and field data, it can be

concluded that the following model is acceptable for further analysis. The next step is to

use design guidelines for the same MSE wall using CFRP instead of Tensar SR2.

6.6. Design guideline for MSE wall

To satisfy the design of MSE walls, there are two sets of stability criteria, and they are as

follows:

External stability
Internal stability

The external stability of a MSE wall is found by idealizing a gravity retaining wall with a

vertical face. The internal stability is dependent on the tensile strength of the reinforcing

material and the interface friction of the reinforcing material and the soil. Progressive

collapse of the wall occurs when the tensile failure of the reinforcing material at any

depth, while slip at the interface of the reinforcing material and the soil mass leads to

redistribution of stresses and progressive deformation of the wall. There are two methods

that can be used to analyze the internal stability. The first method is similar to anchored

flexible retaining walls and it is used for reinforcing material with high extensibility. The

Rankine active earth pressure theory is used with the active slip plane inclined at

=45+/2 to the horizontal. The fictional resistance applies over an effective length Le,
291

which is located outside of the active failure zone. The fictional resistance developed on

top and bottom surfaces of the reinforcing material at depth of z is

Pr=2wLe(z+qs)tani ( 6-12)

w = the width of the reinforcing material,

z = the vertical effective stress,

qs = the surcharge,

i = the frictional angle at the soil reinforcement interface.

The tensile force at depth z is

T=KaR (z + qs) SzSy (6-13)

Sz and Sy = the spacing in the Z and Y directions

T = the tensile force per unit length of wall.

The effective length of the reinforcement necessary for limit equilibrium is determined by

setting T=Pr. A factor of safety (FS)t is applied on the tensile force to find the design

effective length as follow:

Le= Ka (z + qs) SzSy(FS)t/(2wLe(z+qs)tani) (6-14)

(FS)t= 1.3-1.5

The total length of reinforcement is L=Le+LR

LR= length of reinforcement within the active failure zone

Since LR is zero at the base of the wall, the calculated length of reinforcement at the base

is often the shortest. This calculated length is often insufficient for translation or bearing
292

capacity. In the design guideline, the length of reinforcement at the base has to be

checked for adequate translation or bearing capacity as follows:

Calculate the maximum lateral active earth force, Pax.

Pax=H0KaC cos (0.5H0+qs) ( 6-15)

Kac= the active lateral pressure coefficient using Coulombs method with wall friction

= the wall friction

Calculate the length of reinforcement required for translation.

T= ni=1Wi tan b (6-16)

Wi=the weight of soil layer i

n= is the number of layers

b= the interfacial friction angle between the reinforcement and the soil at the base

T= H0 L b tan b (6-17)

The length of reinforcement at the base required to prevent translation under long term

loading is

Lb=(Kac)x (0.5H0+qs/)(FS)T/ tan b (6-18)

Calculate the allowable tensile strength per unit width of the reinforcing polymeric

material.

Tall=Tult/(FSID*FSCR*FSCD*FSBD) (6-19)
293

Tall = the allowable tensile strength,

Tult = the ultimate tensile strength,

FS = a factor of safety

ID=installation damage

CR=creep

CD=chemical degradation

BD=biological degradation

Calculate the vertical spacing at different wall heights.

Sz=Tall/(Ka(z+qs)(FSsp)) (6-20)

FSsp=a factor of safety between 1.3-1.5

Determine the length of reinforcement required at the base for external stability.

Determine the total length of reinforcement at different levels.

Determine the external stability (translation and bearing capacity)

For designing the MSE wall using FRP, the same guidelines are used for Panel 30-32

using CFRP, and it is shown as follows:

H=16 feet

H1= 2.5 feet (location of the first FRP from bottom of the wall)

H2=9 feet (location of the second FRP from bottom of the wall)

Tult =Np=32000 lb/ft

=40
294

=130 pcf

Surcharge (SC)=130 psf

Figure 6-13 MSE wall using FRP.

K0= tan 2(45-40/2)=0.22

x(0)=418*0.22=92 psf

x(10.25)=385.15 psf

x(16)=550 psf

F1=10.25*(92+385)/2=2445 lb/ft

F1=5.75*(385+550)/2=2688 lb/ft
295

Figure 6-14 Lateral forces on MSE wall.

Lateral force due to soil pressure=KaH2/2=0.22*130*162/2=3661 lb/ft

Lateral force due to surcharge pressure=qsKa=418*0.22*16= 1471 lb/ft

Moment at the base of the wall due to lateral pressures=

3661*16/3+1471*16/2=31300lbft/ft

MR=L2H/2+(SC) L2/2= 1249 L2

L=5.0 for design L=7 feet

H=Pax=2445+2688=5133 lb

Vn=418+130*16*7=15000 lb

H/Vn=0.34

W=tan-1(0.34)=18.8
296

n=2

i=(1-0.34)n+1=0.29

N=0.1054 exp (9.6 c)=86

qu=0.5 * 130 *7*86*0.29=11400 lb/ft

FS=11400/(130*16)=5.5 ft

LR1=9 tan (45-20)=4.2 ft

Le1=7-4.2=2.8 ft

2.8 w (2 faces) tan 40 *(7*130+418)=2445

W= 0.4 ft

FS1=0.4*32000/2445=5.2

LR2=2.5 tan(45-40/2)=1.16 ft

Le2=7-1.2=5.8 ft

Pr = 5.8 w (2 faces) tan (40) *(13.5*130+418) = 2688

w=0.12 ft

For Design

L=7 feet

W=0.4 feet

FRP 10 oz/ft2

Nevertheless, in Design guidelines based on field observation of existing MSEws, the

prescribed minimum length of the reinforcement is:

. (6-21)
.
297

In this investigation the above equation was not considered in designing the MSE wall

using CFRP, since the CFRP is almost a new material in geotechnical engineering, and

they are not many filed experiment. However, the stability of the MSE wall using CFRP

was checked numerical analysis. The effect of large vertical spacing (> 0.8 m (2.7 ft)) is

not well understood and the use of these higher spacing must be avoided unless full scale

wall data is available (Castellanos, 2010). Since all the properties of the full scale wall

such as backfill, interface, concrete wall and CFRP are available, and the CFPR has

significantly higher stiffness, tensile strength and interface friction angle, it is consider to

have a higher vertical spacing.

6.7. Plaxis (HS) analyses on MSE wall using CFRP

After using the guidelines for designing MSE wall using CFRP, Plaxis (HS) model is

used to simulate the MSE wall. Then, the results of MSE wall using CFRP are compared

with MSE wall using Tensar SR2. The same model is used as the one used in section 6.5

except only two layers of CFRP with length of 7 feet and width of 0.4 feet (based on

design guide) is used instead of Tensar SR2, and they are as follows:
298

Elevation:5feet
12
11
VerticalStress(psi)

10
9
8 TensarSR2(ten
layers)
7
CFRP(twolayers)
6
5
4
0 50 100 150
DistancefromWallFace(in)

Figure 6-15 Vertical stress vs. distance from wall face for CFRP and Tensar SR2 at elevation 5.0 ft.

Elevation:3.5feet
0.35
0.3
VertSoilStrain(%)

0.25
0.2
TensarSR2(ten
0.15 layers)
0.1 CFRP(twolayers)
0.05
0
0 10 20 30 40
DistanceFromWallFace(in)

Figure 6-16 Vertical soil strain vs. distance from wall for CFRP and Tensar SR2 at elevation 3.5 ft.
299

NeartheWallFace Tensar
18 SR2(ten
16 layers)
14 CFRP(two
WallElevation(ft)

12 layers)

10
Linear
8 (Tensar
6 SR2(ten
4 layers))
Linear(CFRP
2 (twolayers))
0
0 1 2 3 4
HorizontalStress(psi)

Figure 6-17 Horizontal stress vs. elevation of the wall for CFRP and Tensar SR2.

Elevation:1.5feetand2.5feet
0.25

0.2
GeogridStrain(%)

0.15
TensarSR2(ten
0.1 layersforEL1.5)
CFRP(twolayersfor
0.05 EL2.5)

0
0 50 100 150
DistFromWall(in)

Figure 6-18 Geogrid strain vs. distance from wall for CFRP and Tensar SR2.
300

Elevation:8feet
0.6

0.5
SoilStrain(%)

0.4

0.3 TensarSR2(ten
layers)
0.2
CFRP(twolayers)
0.1

0
0 20 40 60 80 100
DistfromWall(in)

Figure 6-19 Soil strain vs. distance from wall for CFRP and Tensar SR2 at elevation 8ft.

18
NeartheWallFace
16
14
VerticalElevation(ft)

TensarSR2(ten
12 layers)
10 CFRP(twolayers)
8
Field
6
4
2
0
0 0.5 1 1.5
HorizontalDisplacement(in)
Figure 6-20 Movement of the wall face for CFRP and Tensar SR2.
301

AxialForceinFRPandGeogrid
16
14
12
10
H(ft)

8 TensarSR2(10
Layers)
6
FRP(2Layers)Design
4
2
0
0 500 1000 1500 2000
tensileforce(lbs/ft)

Figure 6-21 Axial force in CFRP and Geogrids.

El-Hoseiny and Desai (2005) used nonlinear soil and interface models for the same wall

and reported the maximum displacement increased to about 42 mm. However, for

present study, the maximum displacement of the wall is around 20 mm (Figure 20). The

main reason for the discrepancy is considered to be possible errors in the measurements.

It is believed that since other measurements compare well with the predictions, the

displacements from the FE prediction can be considered to be reasonable.


302

6.8. Analyzing number of reinforcing layers and vertical spacing

6.8.1. Number of reinforcing layers

Even so the MSE wall is designed based on the design guideline. The effect of the

reinforcement spacing is needed to investigate for any local internal failure (Figure 2-15

Compound III). There is no close form solution available for it; therefore, it requires

finite element analysis. Thus, different spacing between reinforcement is chosen to

investigate the local internal failure and the effect of the reinforcement spacing & wall

deformations. For that reason, panel 26-32 in the MSE wall is used with different spacing,

with a length of seven feet as follows: ten layers, six layers, four layers, and two layers as

shown in Table 6-2 .

10Layers 6Layers 4Layers 2Layers


Elv.0.5 x x x
Elv.1.5 x
Elv.2.5 x x x
Elv.3.5 x x
Elv.4.5 x x
Elv.5.5 x
Elv.7.0 x x x
Elv.9.0 x x x
Elv.11.5 x x x
Elv.14.5 x
Table 6-2 Location of reinforcement in the MSE wall
By investigating the models using hardening model by Plaxis for all the above layers and

investigating the plastic points, no local internal failure was observed, and for all the
303

layers prescribed ultimate state fully reached. Figure 6-22 shows the effect of

reinforcement spacing and wall deformation. Two layers of CFRP is chosen in the

primary design, since there is no local internal failure for two layers of CFRP.

EffectofReinforcementSpacing
18
16
VerticalElevation(ft)

14
12
10 10Layers
8 6Layers
6
4Layers
4
2 2Layers
0
0 0.2 0.4 0.6 0.8
HorizontalDisplacement(in)

Figure 6-22 Effect of reinforcement spacing and wall deformation

6.8.2. Vertical spacing at the center of the wall (CW)

This part evaluates the minimum, maximum and the optimum vertical distance between

CFRP reinforcements (CUF10U) when spacing starts at the center of the wall (CW). For

this investigation, the same wall, which is used previously, with the same parameters of

soil, interface, geogrid, and concrete panel, is used. First, the two layers of CFRP are

placed at the center of the wall (at an elevation of 8 ft). Then, all the plastic points such as

Elastic point, a Plastic point, a Tension point, an Apex point, a Hardening point, or a Cap
304

point are identified. The same procedures are applied for different spacing between the

two layers of CFRP, as the spacing is increased.

Figure 6-23 (a-n) shows the status of the stresses and indicate whether a stress point is an

Elastic point, a Plastic point, a Tension point, an Apex point, a Hardening point, or a Cap

point. An Elastic point is a stress point that is currently not in a state of yielding. A

Plastic point is a stress point where the Mohrs stress circle touches the Coulomb failure

envelope. A Tension point is a stress point that has failed in tension according to the

tension cut-off criterion. An Apex point is a stress point at the apex of the failure

envelope. A Hardening point is a stress point where stress state corresponds to the

maximum mobilized friction angle that has previously been reached. A cap point is a

stress point where the stress state is equivalent to the preconsolidation stress, i.e. the

maximum stress level that has previously been reached.


305
306
307

Figure 6-23 Plastic points for different spacing between reinforcements (CW)

The failure of the wall is not necessarily found from the above figures, since they show

only plastic points at a particular point. Therefore, the maximum wall deformation and

PHI-c reduction factor are analyzed using Plaxis software.

Table 6-3 shows a summary of the results, which are maximum wall deformation and

Msf reduction factor.

Max.wall
Spacing(ft) Msf deformation(ft)
0 <1.0
0.5 <1.0
1 <1.0
2 <1.0
3 <1.0
4 1.204 7.472
5 1.217 4.257
6 1.259 5.242
7 1.314 1.931
8 1.348 1.088
9 1.397 2.215
11 1.461 1.1
12 1.469 1.231
13 <1.0
14 <1.0
Table 6-3 Maximum wall deformation and Msf value for CW MSE wall

Msf multiplier is associated with the Phi-c reduction option in PLAXIS for the

computation of safety factors. The total multiplier Msf is defined as the quotient of the

original strength parameters and the reduced strength parameters and controls the

reduction of tan and c at a given stage in the analysis. Msf is set to 1.0 at the start of a
308

calculation to set all material strengths to their unreduced values. In the phi-c reduction

approach the strength parameters tan and c of the soil are successively reduced until

failure of the structure occurs. The strength of interfaces is reduced in the same way.

The strength of structural components such as plates and geogrids is not influenced by

Phi-c reduction; therefore, the factor of safety for tensile strength of reinforcement is

needed to check separately. The total multiplier Msf is used to define the value of the

soil strength parameters at a given stage in the analysis:

(6-22)

Where the strength parameters with the subscript input refer to the properties entered in

the material sets and parameters with the subscript reduced refer to the reduced values

used in the analysis.

When the spacing between the two CFRP is smaller than 3 ft, the wall fails. Figure 6-23

(a-d) shows the most retaining soil particles reach tension points which creat a tension

path all the way to top of the surface. Based on

Table 6-3, when the spacing of reinforcements varies from 0 to 3 ft, the values of Msf are

less than 1.0. Then, as the spacing increases from 4 ft until it reaches 12 ft, the values of

Msfs increase and it can be seen there is no tension point which creat any path to the top

(Figure 6-23 (e-l)). However, maximum wall deformation decreases until the spacing is

8 ft; then, it starts increasing. After that, by increasing the spacing more than 12 ft, the

wall fails, and based on Figure 6-23 (m & n), the most retaining soil particles reached
309

tension point which creat a path to the top.

Table 6-3 shows Msf less than 1.0 at spacing more than 12 ft. The summary of these

results are shown in Figure 6-24.

Figure 6-24 Effect of vertical spacing on Msf and wall deformation for CW MSE-wall

As seen in Figure 6-24, wall deformation decreases by increasing the spacing until it

reaches 8 ft. Then, wall deformation starts increasing until soil fails. Therefore, spacing

between 6 and 9 ft is the most suitable for this type of the wall when vertical spacing start

increasing from center of the wall. The next part provides a more general application to

investigating internal failure.

6.8.3. Vertical spacing at the centroid of the lateral forces (CF)

This part evaluates the minimum, maximum and the optimum vertical distance between

CFRP reinforcements (CUF10U) when spacing start at center of forces. For this
310

investigation, the same wall, which is used previously, with the same parameters of soil,

interface, geogrid, and concrete panel, is used. However, in this case, the location of the

resultant of force is calculated based on limit equilibrium. Basically, the lateral pressures

on the wall due to soil and surface pressure are calculated. Then, the locations of the

resultant forces of these two pressures are verified. Then, the center of force (CF) of the

two pressures is defined as a reference line which is approximately at an elevation of 5.75

ft. Following this, the reinforcements are placed at the location of center of force with a

zero spacing. Then, all the plastic points such as Elastic point, a Plastic point, a Tension

point, an Apex point, a Hardening point, or a Cap point are identified. The same

procedure is applied for different spacing between the two layers of CFRP; hence, for all

these different spacings, the middle distance of the two layers of the reinforcement is

always the center of force line (5.75 ft).

Figure 6-25 show the status of the stresses, and indicate whether a stress point is an

Elastic point, a Plastic point, a Tension point, an Apex point, a Hardening point, or a Cap

point.
311
312

Figure 6-25 Plastic points for different spacing between reinforcements (CF)

As mentioned before, the failure of the wall is not necessarily found from the above

figures, since they show only plastic points at a particular point. For that reason, the

maximum wall deformation and PHI-c reduction factor are analyzed using Plaxis

software. Table 6-4 shows a summary of the results, which are maximum wall

deformation and Msf reduction factor.


313

Max.wall
Spacing(ft) Msf deformation(ft)
0 <1.0
0.5 <1.0
1 <1.0
2 1.2676 0.7324
3 1.3215 1.6785
4 1.3486 2.6514
5 1.396 3.604
6 1.4428 4.5572
7 1.468 5.532
8 1.4121 6.5879
9 1.4467 7.5533
10 1.3924 8.6076
11 <1.0
Table 6-4 Summary of FE analysis for different spacing (CF)

When the spacing between the two CFRP are equal and smaller than 1 ft the soil

collapsed. Figure 6-25 (a-b) shows the most retaining soil particles tension points.

According to Table 6-4, the values of Msf are less than 1.0, when the spacing of

reinforcements varies from 0 to 1.0 ft. Then, as spacing increases from 2 ft to 9 ft, the

factor of safety increases, and it can be seen there is no tension points along the geogrid

(Figure 6-25 (c-k)). However, maximum wall deformation decreases until the spacing

reaches 5 ft; then, it starts increasing until it fails. By increasing the spacing to 10 ft, the

factor of safety decreased and the wall deformation increased. After that, by increasing

the spacing, the soil collapsed (Figure 6-25 (l)). The summary of these results are shown

in Figure 6-26.
314

Figure 6-26 Effect of spacing on Msf and wall deformation (CF)

As seen in Figure 6-26, wall deformation decreases by increasing the spacing until it

reaches 8 ft. Then, wall deformation starts increasing until soil fails. Therefore, spacing

between 4 and 8 ft is the most suitable for this type of the wall when vertical spacing start

increasing from centroid of forces. The next part provides a more general application to

investigating only internal failure.

6.8.4. Finial analysis

To finalize the possibility of internal failure which includes pullout failure of the

reinforcement and interface failure between soil and reinforcement, a 100 ft tall wall is

modeled by using Plaxis software. In this model, a hardening plasticity model is used

for soil, a Mohr-Column (M-C) model is used for the interface, and a linear elastic model

is used for soil reinforcements and concrete panel. The same soil parameters which are
315

used for the previous wall are used in this wall. In this investigation, the length of the

reinforcement, the type of the CFRP and elevation of CFRP are varied to identify the

most suitable MSE wall. Since the focus of this stage is studying the internal failure and

to eliminate other failure such as soil bearing capacity failure or overturning failure, the

soil foundation is assumed as exceptionally rigid material by using pin at bottom. This

setup is very similar to anchored sheet pile in which the bottom of the wall connected to

pile by pin, and the top of the wall anchored by reinforcements.

In the first part of this study, CFU10U is used with the maximum tensile strength,

corresponding strain, modulus of elasticity and thickness of 134 ksi, 0.93, 8000 ksi, and

0.023 in, respectively. Then, the length and the elevation of the CFU10U are changed

according to Table 6-5 to obtain maximum Msf. Hence, the strength of structural such as

plates and geogrids is not influenced by Phi-c reduction; therefore, the factor of safety for

tensile strength of reinforcement is needed to check separately.

Based on Table 6-5, CUF10U is placed at six different elevations, with fourteen different

lengths, and Msf for each position are calculated. The values of Msf (PHI-C reduction

factor) for elevation equal and smaller than 60 ft are smaller than 1.0. For the ones that

the Msf values are larger than 1.0, the CFRP axial stress reached its ultimate tensile

strength; therefore the soil collapses. Therefore, CUF10 reinforcement is not sufficient

for this type of wall.

Elevation 30 50 60 70 80 90
Length (ft) (ft) (ft) (ft) (ft) (ft)
50 (ft) <1.0 <1.0 <1.0 <1.0 <1.0
60 (ft) <1.0 <1.0 <1.0 <1.0 <1.0
316

70 (ft) <1.0 0.9384* <1.0 0.9402*


80 (ft) <1.0 <1.0 <1.0 <1.0 <1.0 <1.0
90 (ft) <1.0 <1.0 0.9884* <1.0
100 (ft) <1.0 0.9513* <1.0 <1.0
110 (ft) <1.0 .9709* 1.023* 1.0073*
120 (ft) <1.0 <1.0 <1.0 <1.0 0.974* 1.0223*
130 (ft) <1.0 <1.0 <1.0 <1.0
140 (ft) <1.0 .9365* 0.9425* 1.0385*
150 (ft) <1.0 <1.0 .9770* <1.0
160 (ft) <1.0 <1.0 .9439* <1.0 1.0549*
170 (ft) <1.0 <1.0 .9543* .9808* <1.0
200 (ft) <1.0
Table 6-5 Evaluation of Msf with verification of the length of geogrid (CFU10)

Based on above analysis, stiffer type of CFRP is required to use at different elevations

and locations for the same MSE wall. Therefore, CUF20U with maximum tensile

strength, corresponding strain, modulus of elasticity and thickness of 94 ksi, 0.91, 9000

ksi, and 0.041 in, respectively are used. Based on Table 6-6, CUF20U is placed at seven

different elevations, with thirteen different lengths, and Msf for each position is

calculated.

Elevation 30 40 50 60 70 80 90
Length (ft) (ft) (ft) (ft) (ft) (ft) (ft)
50 (ft) <1.0 .917* <1.0 <1.0 0.984* <1.0
60 (ft) .962* 1.05* 1.023* 0.997*
70 (ft) <1.0 1.029* 1.131* 1.088* 1.094*
80 (ft) <1.0 0.94* 1.125* 1.150* 1.139*
90 (ft) 1.017* 1.164* 1.181
100 (ft) <1.0 1.113* 1.261*
110 (ft) 1.194* 1.212*
120 (ft) 0.929* 1.026* 1.103* 1.17* 1.207
130 (ft) <1.0 <1.0 <1.0
140 (ft) <1.0 1.051* 1.119* 1.214* 1.1878
317

150 (ft) 1.102* 1.276*


160 (ft) <1.0 1.231* 1.2435
170 (ft) <1.0 <1.0 1.034 1.129* 1.080* 1.272*
Table 6-6 Evaluation of Msf with verification of the length of geogrid (CFU20)

Based on Table 6-6, the values of Msf for elevation equal or smaller than 50 ft are less

than 1.0. The values of Msf between elevation of 60 to 90 are higher than 1.0; however,

the strength of CUF20U reached its maximum strength. In this case, CUF20U is not

sufficient for stability of the wall.

Then, for next step, one layer of CFRP made of two layers CUF20U is used. Once more,

the value of Msf is calculated at different elevation with different length as shown in

Table 6-7.

Elevation 30 40 50 60 70 80 90
Length (ft) (ft) (ft) (ft) (ft) (ft) (ft)
50 (ft) 1.062* 1.169* 1.1883 1.1064 1.1319 1.1424 1.0131
60 (ft) 1.2744 1.3273 1.2385 1.2337 1.1017
70 (ft) 1.0313* 1.1837 1.2195 1.3949 1.3683 1.2931 1.2074
80 (ft) 1.3255 1.2798 1.4119 1.3991 1.2788
90 (ft) 1.1607 1.2167 1.4594 1.4462 1.3864
100 (ft) 1.0465* 1.1034 1.2616 1.2486 1.4054 1.4672 1.4885
130 (ft) 1.0529* 1.2264 1.2433 1.22 1.374 1.6541 1.6578
160 (ft) 1.0428 1.1264 1.2283 1.2726 1.4027 1.4932 1.5979
Avg 1.0471 1.1618 1.2377 1.2583 1.3490 1.3911 1.3415
Stdev 0.01145 0.0484 0.0513 0.0849 0.1086 0.1624 0.2320
Table 6-7 Elaluation of Msf with verification of the length of geogrid (2 layers CUF20)

Based on the results obtained in Table 6-7, it can be seen that the value of Msf first

increases as the elevation decreases. Then, the value of Msf decreases as the elevation
318

keeps decreasing. Also, as the length of the reinforcement increases, the value of Msf

increases. Therefore, as the elevation and the length of reinforcement increases, the value

of the Msf increases as shown in Table 6-7. To come up with a conclusion, the

maximum CFRP axial stress needs to check. Therefore, Table 6-8 shows axial tensile

stress on CFRP.

Elevation 30 40 50 60 70 80 90
Length (ft) (ft) (ft) (ft) (ft) (ft) (ft)
50 (ft) 90000* 90000* 80859 71335 79808 78868 58000
60 (ft) 78286 76499 78026 74688 62390
70 (ft) 90000* 85377 73502 75069 73378 68710 63335
80 (ft) 80793 71850 72819 70913 63737
90 (ft) 73119 51566 70361 68174 66420
100 (ft) 90000* 77539 76172 69351 72508 67578 66865
130 (ft) 90000* 87878 74513 70352 72114 73016 67955
160 (ft) 90000* 83460 42434 72331 41430 70012 65179
Avg 90000 84850 72459 69794 70055. 71494 64235
Stdev 0 4778 12507 7732 11994 3845 3156
Table 6-8 Axial force (lb/ft) fot CFRP

Based on Table 6-8, CFRP axial stress increases as the elevation and length of CFRP

decrease. When CFRP reaches 90000 lb/ft, the CFRP fail; then, the MSE wall fails.

From the three types of CFRP that are used for this analysis, two layers of CFU20U are

sufficient for this particular wall. Nevertheless, calculating the tensile strength for

reinforcement using the design guidelines (Section 6-6), that is based on limit equilibrium,

gives almost the same tensile strength as provided in Table 6-8; however, more on the

conservative side. Therefore, it can be concluded that using limit equilibrium for
319

designing the reinforcement is adequate. Table 6-9 shows factor of safety for tensile

strength of CFRP which is applied tensile stress divided by ultimate tensile stress.

Elevation 30 40 50 60 70 80 90
Length (ft) (ft) (ft) (ft) (ft) (ft) (ft)
50 (ft) <1.0 <1.0 1.113 1.262 1.128 1.141 1.552
60 (ft) 1.150 1.176 1.153 1.205 1.443
70 (ft) <1.0 1.054 1.224 1.199 1.227 1.310 1.421
80 (ft) 1.114 1.253 1.236 1.269 1.412
90 (ft) 1.231 1.745 1.279 1.320 1.355
100 (ft) <1.0 1.161 1.182 1.298 1.241 1.332 1.346
130 (ft) <1.0 1.024 1.208 1.279 1.248 1.233 1.324
160 (ft) <1.0 1.078 2.121 1.244 2.172 1.285 1.381
Avg <1.0 1.079 1.293 1.307 1.336 1.262 1.404
Stdev 0.059 0.338 0.182 0.342 0.065 0.072
Table 6-9 Factor of safety for CFRP

Based on Table 6-7 and Table 6-9, it can be concluded that as the elevation decreases

from 90 ft to 30 ft, factor of safety for CFRP decreases. In other words, axial stress

increases in reinforcement as elevation decreases. It can be concluded that at the higher

elevation the failure is the interface between CFRP and Soil which is pullout failure.

Then as the elevation decreases, the failure is due to CFRP failure. Therefore, the most

suitable elevation and length of reinforcement is when CFRP factor of safety and Msf is

higher than 1.3 and the two factors of safeties have a close values. Therefore, the most

appropriate place for reinforcement is at an elevation of 80 ft with a length of 100 ft.

Hence, the factor of safety can be varies depending obligation of design guidelines.

After identifying the most suitable position of the first layer of reinforcement, the most

appropriate place for the second layer is investigated. In this part of the study, the
320

minimum and maximum vertical spacing between the two layers of reinforcement is

studied. The same wall is used and the first layer of reinforcement is placed at an

elevation of 80 ft with a length of 100 ft. However, the strength (Np) and stiffness (EA)

of CFRP are divided by two; in other words, instead of using two layers of CUF20U,

only one layer of CUF20U is used. Then, the second layer, which is a layer of CUF20U,

is placed at a different elevation to classify the maximum and minimum vertical spacing.

Table 6-10 shows the axial stress of reinforcements, Msf, maximum wall deformation,

and factor of safety for CFRP.

Elevation 63 65 67 71 72 73 75 85 90
(ft) (ft) (ft) (ft) (ft) (ft) (ft) (ft) (ft)
Msf 1.581 1.652 1.683 1.571 1.611 1.584 1.347 1.654 1.533
80'
Np lb/ft 45000 44281 43257 40719 41228 39775 35771 36645 31084
cor
Np lb/ft 36247 37759 39378 33513 36078 34507 31092 40610 35382
WallDispft 0.873 0.931 0.943 0.819 0.859 0.812 0.619 0.875 0.72
80'
FSCFRP <1.0 1.016 1.040 1.105 1.091 1.131 1.258 1.228 1.448
cor
FSCFRP 1.241 1.192 1.143 1.343 1.247 1.304 1.447 1.108 1.272
80'
FSMsf/FSCFRP a 0.615 0.618 0.704 0.678 0.714 0.934 0.743 0.945
cor'
FSMsf/FSCFRP b 0.785 0.721 0.679 0.855 0.774 0.823 1.075 0.670 0.830
2 2
(1a) +(1b) 0.226 0.249 0.109 0.155 0.113 0.010 0.175 0.032
Table 6-10 Evaluation of Msf with verification of the length of geogrid (2 layers CUF20)

As mentioned before, the first layer of CUF20U is placed at an elevation of 80 ft, and the

second layer of CUF20U is placed at different elevations to obtain the most appropriate

position. Therefore, the following equation is used to obtain a mathematical procedure to


321

achieve the most suitable vertical spacing by considering the strength of the soil, interface

and geogrids.

(6-23)

As the value of S reaches zero, the design of the wall would be more sufficient due to

vertical spacing. Hence, all the factors of safety need to be checked based on the design

requirements. As the results show above, it can be seen that as the elevation decreases

there is no significant decrease on Msf value; however, there is a significant decrease on

in the CFRP factor of safety. Based on Table 6-10 and using equation 6-23, when the

second layer of CUF20U is at an elevation of 7 ft of the first layer of CFRP, it provides

the most suitable results. When the second layer of CFRP is placed at an elevation of 75

ft, it provides the minimum wall deformation which is 0.619 ft and the lowest value of S.

Based on the investigation, three types of walls are preformed to study the internal failure

of the MSE wall.

The first study, a wall sixteen feet tall was studied. Two layers of CFRP reinforcements

were used for stability of the wall. The stability of the wall was investigated by

increasing the spacing of reinforcements very one foot from center of the wall. Then, for

the second study, the same wall is used; however, the stability of the wall was checked by

increasing the spacing of reinforcements very one foot from center of the forces. Figure

6-27 shows the Msf and maximum wall deformation for the MSE wall in which the
322

spacing of the reinforcement once start from center of the wall (CW) and once from

center of forces (CF).

1.6 8
Msf(CF)
1.4 7

1.2 6 Msf(CW)

1 5

Wall Deformation(ft)
Displacement(CF)
Msf

0.8 4
Displacement(CW)
0.6 3
Poly.(Msf(CF))
0.4 2

0.2 1 Power(Msf(CW))

0 0 Poly.(Displacement
0 2 4 6 8 10 12 14 (CF))
Spacing(ft) Poly.(Displacement
(CW))

Figure 6-27 Comparison results for two cases

According to Figure 6-27, the spacing that the midpoint of layers is the center of forces

(CF) provides less wall deflection and the higher factor of safety. At a spacing of 2 ft, the

maximum CF wall deformation is approximately 3.5 feet and the value of Msf is

approximately 1.25 wall; on the other hand, the CW wall fails. For spacing of 6 ft, the

CF wall deforms approximately 0.6 ft with Msf of 1.44; on the other hand, CW wall with

the same geogrid spacing deforms 5 ft. Most codes do not allow 5 ft wall deformation for

a wall 16 ft tall. At a spacing of 12 ft the CF wall fails; nevertheless, CW wall deforms

approximately 1.2 ft with Msf of 1.4. At higher spacing, both walls fail. Based on this

result, it can be recommended for designing a MSE wall using two layers of CFRP, the
323

midpoint of the vertical spacing between the two layers is placed at the centroid of lateral

loads (earth and surface pressure). The spacing can vary between 5 and 8 ft; this case

provides the most optimum wall. Nevertheless, parameters such as vertical spacing and

length of reinforcement require to check based on design guidelines (Section 6-6). Hence,

the design guideline is according to limit equilibrium method which has some assumption

such as rigid body movement. Therefore, limit equilibrium provides no information

regarding deformation of the MSE wall. For that reason, constitutive models are adopted

to address these limitations. However, there is no a close form solutions for these models;

thus, it requires finite element analysis. By comparing the close form solution (limit

equilibrium method) and finite element (constitutive model), it is acceptable using the

limit equilibrium method plus using the suggestion regarding vertical spacing in this

section, and the design is on a more conservative side.

In the last part of this investigation, a 100 ft tall wall is analyzed. The pin connection was

placed at the bottom of the wall to prevent over turning failure and bearing capacity

failure. Hence, this type of analysis is commonly used for anchored sheet pile. Three

different types of CFRP (CFU10, CFU20 and 2 layers of CFU20) with different lengths

are used. These CFRPs are placed at different elevations, and factor of safety for soil,

interface and CFRP are preformed. Based on 100 models, the most suitable elevation is

suggested at an elevation of 80 ft with a length of 100 ft. Two layers of CFU20 is

required to use for the wall. Hence, limit equilibrium method is required a CFRP layers
324

that has 20% higher tensile strength than two layers of CFU20. Therefore, it can be seen

that limit equilibrium is acceptable for designing the wall.

To determining the most suitable vertical spacing for the wall, the first layer of

reinforcement is placed at an elevation of 80 ft with a length of 100 ft based on previous

analysis; however, only one layer of CFU20 is used. Then the second layer of CFU20 is

placed at different elevations, and the factor of safety for soil, interface, and CFRP is

checked. Based on the finite element analysis, 5 ft vertical spacing between the two

layers of reinforcement provides the most suitable performance.

As mentioned before, soil is not a homogenous material, and it is highly possible that soil

particles within the same area at the sight may be totally different in sense of properties.

They are different models are suggested by researchers; however, engineers try to use a

simple model and also represent most adequate soil behavior. For designing MSE wall,

most geotechnical engineers use limit equilibrium method, and then they usually check

the design by finite element using constitutive model. Based on this section, limit

equilibrium method is acceptable to use to design a CFRP-MSE wall. It is suggested to

have spacing of 4-7 ft between layers. It is also suggested to place the midpoint of the

reinforcement layers at centroid of lateral forces.

6.9. Construction sequences and cost estimation

This section presents a summary of the sequential construction of the Tanque Verde Road

project MSE walls which is taken from Design, construction and performance of tensar
325

geogrid reinforced soil walls at Tanque Verde-Wrightstown-Pantano Road report.

However, the same construction sequences can be use for CFRP instead of geogrid.

Step 1:

Foundation soil is removed to a depth equal to the lesser of the particular wall panel

height or 7 ft and compacted to at least 90% of maximum dry density based on the

modified proctor test (ASTM D1557). Compaction is essential because possibility of

having weak soil or alluvial soil which are potentially collapsible when wetted.

Step 2:

Leveling pads are cast-in-place and stepped to reduce panel height and wall fill earthwork

volumes. At the same time, concrete facing panels are cast on site and stockpiled until

they are needed. The facing panels are cast with lift hooks so that they can be hoisted by

a crane. Panels weigh between 1,350 and 16,500 lbs. The erection braces are pinned to

the reaction blocks and wall panels.

Step3:

Concrete facing panels are lifted by their lift hooks with a crane and set vertically on the

leveling pads. Panel alignment is done by placing the panels such that the leveling bolts

cast into the leveling pads go through the precast recesses in the bottoms of the panels.

The wall panels are held in place with the adjustable erection braces which are set on 10
326

ft centers in front of the vertical wall joints. Initially, the wall panels are battered inward

at 1H:96V. After the erection braces are placed, the leveling bolts are adjusted to

improve panel alignment.

Step4:

At the start of wall construction, no stipulations are made to clamp abutting facing panels

at their tops so as to enforce uniform construction movements of the panels (due to

rotations about the leveling bolts) are measured. At the tops of the panels, dissimilar

movements between adjacent panels on the order of 1 in. are measured. To alleviate this

problem, Sundt Construction Company initiated clamping of adjacent panels at their tops.

The clamps consist of two wood boards, a threaded steel bar and two nuts. The clamps

are positioned at the tops of the vertical joints between adjacent panels and tighten.

Step 5:

Strips of nonwoven geotextile are glued to the facing panels and leveling pads at vertical

wall joints. The geotextile is proposed to function as a filter to keep the wall fill from

washing out between the joints.

Step 6:
327

Concrete grout is troweled into the space between the bottom of each facing panel and

the top of the leveling pad. The reason for the grout is to support the wall panels should

the steel leveling bolts rust.

Step 7:

Granular wall fill (clean, moderately well-graded gravelly sand) is positioned up to the

first geogrid (for this study instead of geogrid, Carbon FRP can be used) level by a front

end loader. The granular fill is compacted to at least 95% of the maximum dry density

based on the modified proctor test. Fill is compacted near the wall by hand-operated

lightweight vibratory tampers or hand-operated vibratory steel drum rollers. Further from

the wall, an operator driven, 1-ton, vibratory steel drum roller is used.

Step 8:

At each necessary level of reinforcement, a continuous layer of gergrid/CFRP is placed

on the granular fill and secured to geogrid/CFRP tab embeded in the wall panel using the

PVC pipe connector.

Step 9:

The geogrid/CFRP reinforcement-facing panel connection is pretensioned by placing a

timber wedge between the PVC pipe and wall facing to remove slack between the wall

facing and the connection. Any remaining slack is pulled out of the connection by
328

inserting steel T sorks through the apertures of the geogrid/CFRP and into the ground and

slowly pulling away from the wall. In addition to removing slack from the connection,

this procedure produces a small pretension in the geogrid/CFRP. Once the geogrid/CFRP

is pretensioned, the fork is driven further into the ground to preserve the tension until

adequate wall fill has been placed on top of the geogrid/CFRP to prevent loss of the

tension. The pretensioning procedure works well with the clean, gravelly sand fill.

Step 10:

Granular wall fill is placed on the geogrid/CFRP layer and spread with a front end loader.

The typical thickness of the lifts is 8 in. The compacted lift thicknesses for the

instrumented sections of the walls are thicker to guard the instruments from construction

induced damage. After a lift of fill is placed, the steel T forks used to pretension the

geogrid/CFRP and the timber wedges used at the wall connection are removed. Fill is

compacted as described in step 7.

Step 11:

As the height of the fill increases, the erection braces on the outside of the wall are

loosened, allowing the wall to deflect outward and the load to be transferred to the

geogrid/CFRP reinforcement.

Step 12:
329

Steps 8 through 11 are repeated for each geogrid/CFRP layer until the fill reaches the top

of the wall. In a few cases, the erection braces are removed when the reinforced fill

reaches a level equal to about two thirds of the total wall height. Usually, the braces are

left attached until the fill reached the full wall height.

Step 13:

Retained granular backfill behind the reinforced zone is compacted to at least 95% of the

maximum dry density based on the modified proctor test

Step 14:

A concrete traffic barrier is cast in place at the top of the wall. Prior to casting the traffic

barrier, the connector clamps between panels are removed.

Step 15:

After the completion of wall construction, the wall facings are sandblasted or, in the case

of abutments, lightly jack hammer to produce an exposed aggregate finish.

The following page shows a cost estimate of a 10 ft long by 16 ft tall MSE wall which

was designed in this chapter.

PRELIMINARY PROJECT COST ESTIMATE FOR MSE WALL USING TENSAR REINFORCEMENT
330

SUMMARY OF PROJECT COST ESTIMATE


TOTAL MSE Wall ITEMS $ 27,000
TOTAL PROJECT CAPITAL OUTLAY COSTS $ 27,000
Calculated by Vahab Toufigh 05/26/12
(Date)
Checked by Ed Farnaghi 05/27/12
(Date)

MSE Wall
Section 1 - Earthwork Quantity Unit Unit Price Item Cost

Clearing & Grubbing 1 LS* $1,000 $1,000

Roadway Excavation 62 CY** $25.00 $1,556


1
Imported Borrow 144 CY $12.00 $1,728
2
Compacting Backfill 142 CY $30.00 $4,267
3
Water for compaction 1 LS $2,357.20 $2,357

Subtotal Earthwork: $10,907


Section 2 - Material
Concrete for Panel Class A 3 CY $500.00 $1,481
4
Geotextile material and installation 1 LS $320.00 $320
9
Tensar material and installation 1 LS $1,159.10 $1,159

Install Drainage Inlet Marker 5 EA*** $125.00 $625


6
Concrete for Pavement (Class A) 1 LS $1,500.00 $1,500
7
Concrete for Leveling (Class A) 1.0 LS $400.00 $400

PVC Pipe (schedule 40) 100 FT**** $1.00 $100

Framing for concrete pannel 320 SF $5.00 $1,600

Subtotal Drainage: $7,186

TOTAL SECTIONS 1 & 2 $18,093

Section 3 - Minor Items


$18,093 x 10% = $1,809
(5 to
10%)
TOTAL SECTION 3 MINOR ITEMS: $1,809
8
Section 4 - MSE Wall Mobilization
$19,902 x 10% = $1,990

(10%)
TOTAL SECTION 4 MOBILIZATION ITEMS: $1,990

Section 5 - MSE Wall Additions


Supplemental Work
331

$19,902 x 10% = $1,990


(5 to
10%)
Contingencies
$19,902 x 15% = $2,985

(**%)
TOTAL SECTION 5 MSE Wall ADDITIONS: $4,976

TOTAL MSE Wall ITEMS: $26,868

(Subtotal Sections 1 thru 8)

USE FOR ESTIMATE: $27,000


Estimate Prepared by: Vahab Toufigh

(Print or Type Name)


Ed
Checked Prepared by: Farnaghi

(Print or Type Name)

PRELIMINARY PROJECT COST ESTIMATE FOR MSE WALL USING CFRP REINFORCEMENT

SUMMARY OF PROJECT COST ESTIMATE


TOTAL MSE Wall ITEMS $ 26,000
TOTAL PROJECT CAPITAL OUTLAY COSTS $ 26,000
Calculated by Vahab Toufigh 05/26/12
(Date)

Checked by Ed Farnaghi 05/27/12


(Date)

MSE Wall
Item
Section 1 - Earthwork Quantity Unit Unit Price Cost

Clearing & Grubbing 1 LS* $1,000 $1,000

Roadway Excavation 62 CY** $25.00 $1,556


1
Imported Borrow 144 CY $12.00 $1,728
2
Compacting Backfill 142 CY $30.00 $4,267
3
Water for compaction 1 LS $2,357.20 $2,357

Subtotal Earthwork: $10,907


Section 2 - Material
Concrete for Panel Class A 3 CY $500.00 $1,481
4
Geotextile material and installation 1 LS $320.00 $320
332

CFRP material and installation5 1 LS $410.00 $410

Install Drainage Inlet Marker 5 EA*** $125.00 $625


6
Concrete for Pavement (Class A) 1 LS $1,500.00 $1,500
7
Concrete for Leveling (Class A) 1.0 LS $400.00 $400

PVC Pipe (schedule 40) 100 FT**** $1.00 $100

Framing for concrete pannel 320 SF $5.00 $1,600

Subtotal Drainage: $6,436

TOTAL SECTIONS 1 & 2 $17,344

Section 3 - Minor Items


$17,344 x 10% = $1,734
(5 to
10%)
TOTAL SECTION 3 MINOR ITEMS: $1,734
8
Section 4 - MSE Wall Mobilization
$19,078 x 10% = $1,908

(10%)
TOTAL SECTION 4 MOBILIZATION ITEMS: $1,908

Section 5 - MSE Wall Additions


Supplemental Work
$19,078 x 10% = $1,908
(5 to
10%)
Contingencies
$19,078 x 15% = $2,862

(**%)
TOTAL SECTION 5 MSE Wall ADDITIONS: $4,770

TOTAL MSE Wall ITEMS: $25,756

(Subtotal Sections 1 thru 8)


USE FOR
ESTIMATE: $26,000
Estimate Prepared by: Vahab Toufigh

(Print or Type Name)


Ed
Checked Prepared by: Farnaghi

(Print or Type Name)

*Lump Sum (LS)


**yd3 (CY)
***Each (EA)
****ft (FT)
333

*****yd2 (SY)
1) d=116;=4.38%
2) d=118;=8.41%; Handheld compactor rental @ $150/day & $300/day labor cost
3) (c-l)**V; Water truck rental @ $60/hrs; 2,186 gal of water and water meter rental from local water
agency
4) 18 SY***** of Geotextile for drainage system @ $1.08/SY and $300 labor cost for
installation
5) 11 SY of CFRP for reinforcement (includes two layers) @ $10.00/SY and $300 labor cost
for installation
6) Pavement Concrete: 2 CY @ $500/CY (include concrete and transportation) and $500 labor cost
7) Leveling Concrete: 0.2 CY @ $500/CY (include concrete and transportation) and $300 labor cost.
8) Traffic control and contraction area signs
9) 133 SY of Geogrid for reinforcement @ $2.70/SY and $800 labor cost for
installation
334

Summary

In this chapter a background to MSE walls is presented. A finite element computer

program, Plaxis, is used to calibrate the sensor data obtained from an existing MSE wall

in Tucson, AZ. Then, the same MSE wall was designed using CFRP (instead of

geogrids) based on current design guidelines. Plaxis (HS) model is used to confirm the

design. The results were then compared to the MSE wall designed using geogrids. The

new Geo-Composite which is made of Carbon Fiber Reinforced Polymer (CFRP) is

proposed to overcome some of the short comings of the existing geosynthetics such as

Tensar SR2 geogrid. The new Geo-Composite has all the benefits of the geogrid plus

a relatively higher friction angle, higher tensile strength, higher modulus and no

significant creep. CFRP is about 3-5 times more expensive than geogrids; nevertheless,

CFRP has relatively higher interface friction angle and it is much stiffer. It is highly

possible for designing earth structures such as a MSE wall which is reinforced with

CFRP will use less reinforcement than a MSE wall which is reinforced with

geosynthetics; e.g. for a Tensar reinforced wall (Desai and El Hoseiny (2005)), it took ten

layers of Tensar; however, for CFRP reinforced wall only two layers of CFRP requires to

yield similar performance, and the estimated total cost for CFRP reinforced wall is about

3% cheaper than Tensar reinforced wall.


335

CHAPTER 7. CONCLUSIONS

Soil is weak in tension; however, materials such as geotextiles are used to address this

deficiency. In recent years, more than one million square meters of geotextiles have been

used for reinforcing soil. Despite its wide use, there are a number of significant problems

associated with geotextiles, such as creep, low modulus of elasticity, and susceptibility to

aggressive environments. Carbon fiber reinforced polymer (CFRP) was introduced over

two decades ago to the field of structural engineering and it can also be used in

geotechnical engineering. CFRP has all the benefits associated with geotextiles and it has

a higher strength, higher modulus, no creep and reliability in aggressive environments.

The main purpose of this research is to establish practical analytical models for predicting

the strength and behavior of soil reinforced using CFRP. CFRP is chosen as the

reinforcing element in this research. The laboratory tests include direct shear test, pull

out test, and cyclic multi degree of freedom (CYMDOF) test and they are used to

evaluate the behavior of different types of soils and different types of CFRP interfaces

under different loading conditions.

In this study, hierarchical single surface (HISS) models has been used to model the

behavior of the soil - CFRP interface. This model is able to capture most of the important

characteristics of the soil such as state stress, stress path, initial density, and volume

change. The parameters in the constitutive model are found using the data from interface

direct shear tests. The model is validated with respect to the observed laboratory behavior

of the soil.
336

There was some error with the SSD-2D finite element program using HISS; however,

Plaxis using a hardening model is used to simulate a MSE wall. First, the finite element

model of the geosynthetics reinforced earth structure has been verified against field

measurements of monitored geogrid reinforced earth retaining wall in Tucson, Arizona

USA. The parameters explored were movements of wall face, lateral earth pressure,

vertical stress in the soil mass, strain in the soil mass, and strains along the geogrid. After

assuring observed and calculated values gave a good agreement, CFRP is used instead of

geogrids to look at the advantage and disadvantage of this new product. The results of

Plaxis (HS) shows that only two layers of FRP with length of 7 feet and width of 0.4 feet

per feet wall is needed to have the same wall deformation as ten layers of geogrid with

length of 12 feet and width of one feet for one feet of wall.

Based on the results obtained from laboratory work, field measurements and theoretical

analysis, the following conclusion can be reached:

The literature review demonstrates that the use of geosynthetics as an earth

reinforcement material. However there are some significant problems associated

with geosynthetics, such as creep and low modulus of elasticity.

In this research, a new Geo-Composite which is made of Carbon Fiber Reinforced

Polymer cured CFRP is used to overcome some of the short comings of the

existing geosynthetics.
337

The best results for interface between CFRP and fine sand (Kerman, Iran) and

coarse sand (Tucson, AZ USA) were investigated, and the internal friction angles

are 40 and 45, respectively.

Pull test and direct shear test are used to obtain the friction angle between CFRP

and fine sand, and cyclic multi degree of freedom (CYMDOF) test is used to

obtain the friction angle between CFRP and coarse sand.

Different procedures are used for preparing CFRP; however, the most economical

and optimum CFRP which is made of flexible epoxy resin with no soil attached to

the surface is used for theoretical calculation and modeling.

The model based on hierarchical single surface (HISS) 0 approach successfully

simulated the interface behavior of soil and CFRP under different loading

conditions.

Plaxis software using hardening model is used to analyze a MSE wall.

Plaxis software using a hardening model is calibrated / checked with field data

from a Tucson MSE wall, panel 26-32, which included ten layers of Tensars

SR2.

It is concluded that there is a good correlation between the field measurements

and the predicted behavior.

The same MSE wall redesign using FRP, and then Plaxis software using a

Hardening model is used to analysis the wall.


338

It is concluded two layers of CFRP with length of 7 feet with width of 0.4 feet per

feet achieve the same results as ten layers of Tensars SR2 with a length of 12 feet

and with width of one foot per foot.

CFRP is about 3-5 times more expensive than geogrids; nevertheless, CFRP has

relatively higher interface friction angle and it is much stiffer.

Tensar reinforced wall (Desai and El Hoseiny (2005)), it took ten layers of

Tensar; however, for CFRP reinforced wall only two layers of CFRP requires to

yield similar performance, and the estimated total cost for CFRP reinforced wall

is about 3% cheaper than Tensar reinforced wall.

There are still some questions which need to be answered. Further studies will allow

more practical and accurate analysis and consequently spreading CFRP reinforced

soil techniques. Some of the consequent recommendations for further research can be

suggested as follows:

The connection between CFRP and concrete panels needs to be studied.

The proposed model with finite element analysis should be extended with

pore water pressure as an unknown variable and utilizing soft clay, silt, and

landfill soils. Backfill materials can be analyzed under drained and undrained

loading conditions.

The effect of temperature on the strain behavior should be considered.


339

A full scale or half scale MSE wall with sensors for vertical and horizontal

stress and strain should be constructed to make sure the model and data field

have good correlating results.


340

REFERENCES

AASHTO T 99, (2010), Standard Method of Test for Moisture-Density Relations of


Soils Using a 2.5-kg (5.5-lb) Rammer and a 305-mm (12-in.) Drop.

Abdulgauwad, S.N., Bayomy, F., Al-Shaikh, A.M., and Al-Amoudi, O.S.B., (1993),
Influence of Geotextiles on Performance of Saline Sabkha Soils, Journal of
Geotechnical Engineering, Vol. 120, No. 11, pp. 1939-1960.

Abdullah, W.S., and Al-Abadi, A. M. (2005). Implementation of the Electrokinetic as an


Effective Method for Soil Improvement, International Conference on Problematic Soils
GEOPROB 2005, Famagusta, N. Cyprus, 885-894.

Abg Geosynthetics, 2010, Land Fill, [Online Image] at: http://www.abg-


geosynthetics.com/overview.jpg, 09/18/2010

Abramento, M., (1993), Analysis and Measurement of Stresses in Planar Soil


Reinforcements, Ph.D. thesis, Massachusetts Institute of Technology, USA, pp 288.

Abu-Hejleh, N., Zornberg, J.G., Wang, T., Watcharamonthein, J., (2002), Monitored
Displacement of Unique Geosynthetic-Reinforced Soil Bridge Bbutments,
Geosynthetics International 9 (1), 7195.

Acfwest Inc, 1999, Reinforced embankment. [Online Image] Available


at: http://acfwest.com/wp-
content/themes/rttheme7/timthumb.php?src=http://acfwest.com/wp-
content/uploads/2010/03/Erosion-Control-Blanket1.jpg&w=262&h=103&zc=1,
09/09/2010

Adib, M.E., Mitchell, J.K., and Christopher, B.R., (1990), Finite Element Modeling of
Reinforced Soil Walls, Proc. Design and Performance of Earth Retaining Structures, P.C.
Lamb and L.A Hansen, Eds., ASCE Geotechnical Special Publication No. 25, pp. 209-
423.

Agarwal B., (1990), Analysis and Performance of Fiber Composites, 2nd Edition, John
Wiley and sons, New York, 449 Pages.

Ahmadabadi M., Ghanbari A., (2009), New Procedure for Active Earth Pressure
Calculation in Retaining Walls with Reinforced Cohesive-Frictional Backfill,
Geotextiles and Geomembranes, Volume 27, Issue 6, December 2009, Pages 456-463.
341

Akili, W., and Torrance, J.E., (1981), The Development and Geotechnical Problems of
Sabka with Preliminary Experiments on Static Penetration Resistance of Cement Sands,
Q.J. Engrg. Geol., Vol. 14, pp. 59-73.

Al-Amoudi, O.S.B., Abduljauwad, S.N., El-Naggar, Z.R., and Safar, M.M., (1991),
Geotechnical Considerations on Field and Laboratory Testing of Sabkha, Proceedings,
Symp. on Recent Adv. In Geotech. Engrg. III, Premier Conference, Singapore, pp. 1-6.

Allen T. M., and Bathurst R. J., (2001), PREDICTION OF SOIL REINFORCEMENT


LOADS IN MECHANICALLY STABILIZED EARTH (MSE) WALLS, Prepared for
Washington State Department of Transportation and in cooperation with US Department
of Transportation Federal Highway Administration, pg. 250-257.

American Association of State Highway and Transportation Officials (AASHTO), (2002),


Standard Specifications for Highway Bridges, seventeenth ed.. American Association
of State Highway and Transportation Officials,Washington, DC, USA.

Anagnostopoulos, C. A., and Stavridakis, E.I. (2003). Physical and Engineering


properties of a cement stabilized soft soil treated with acrylic resin additive. Electronic
Journal of Geotechnical Engineering.

Andrejack T.L., and Wartman, J., (2010), Development and interpretation of a multi-
axial tension test for geotextiles, Original Research Article Geotextiles and
Geomembranes, Corrected Proof, Available online 14 May 2010

ASTM D421 - 85(2007) Standard Practice for Dry Preparation of Soil Samples for
Particle-Size Analysis and Determination of Soil Constants, ASTM Book of Standards,
Vol. 04.08, 2007.

ASTM D422 (2007), Standard Test Method for Particle-Size Analysis of Soils, ASTM
Book of Standards, Vol. 04.08, 2007

ASTM D698 (2007) Standard Test Methods for Laboratory Compaction


Characteristics of Soil Using Standard Effort, Vol 04.08, 2007.

ASTM D3039, (2008), Standard Test Method for Tensile Properties of Polymer Matrix
Composite Materials, Vol. 15.03, 2008.

ASTM D4355, (2007), Standard Test Method for Deterioration of Geotextiles by


Exposure to Light, Moisture and Heat in a Xenon Arc Type Apparatus, ASTM Book of
Standards, Vol. 04.13, 2007. ASTM, Philadelphia, PA.
342

ASTM D5321, (2008), Standard Test Method for Determining the Coefficient of Soil
and Geosynthetic or Geosynthetic and Geosynthetic Friction by the Direct Shear Method.
ASTM Book of Standards, Vol. 04.13, 2008.

ASTM D6706, (2007), Standard Test Method for Measuring Geosynthetic Pullout
Resistance in Soil ASTM Book of Standards, Vol. 04.13, 2009.

Bara, J.P., (1997), Research on Wetting Collapsible Foundation Soils, Report No. GR-
14-77, Bureau of Reclamation, Denver, Colorado, Sept.

Barden, L., McGown, A., and Collins, K., (1973), The Collapse Mechanism in Partly
Saturated Soil, Engineering Geology, Amsterdam, The Netherlands, pp. 49-60.

Basudhar, P.K., Dixit, P.M., Gharpure, A., Deb, K., (2008), Finite Element Analysis of
Geotextile-Reinforced Sand-Bed Subjected to Strip Loading, Geotextiles and
Geomembranes 26 (1), 9199.

Bathurst, R.J., Karpurapu, R., and Jaret, P.M., (1992), Finite Element Analysis of A
Geogrid Reinforced Soil Wall, Proc. Grouting, Soil Improvement and Geosynthetics,
R.H. Botden, R.D. Hohz and I. Juran, Eds., ASCE Geotechnical Special Publication No.
30, Vol. 2, pp. 1213-1224.

Berg, R.R., Bonaparte, R., Anderson, R.P., and Chouery, V.E. (1986), Design,
Construction, and Performance of Two Reinforced Soil Retaining Walls, Proceedings of
the 3rd International Conference on Geotextile, Vienna, Austria, Vol. 2, pp. 401-406.

Bergado, D.T., Ramana, G.V., Sia, H.I., and Varun, (2006), Evaluation of Interface
Shear Strength of Composite Liner System and Stability Analysis for a Landfill Lining
System in Thailand, Geotextiles and Geomembranes 24 (6), 371393.

Bergado, D.T., Teerawattanasuk, C., (2008), 2D and 3D Numerical Simulations of


Reinforced Embankments on Soft Ground, Geotextiles and Geomembranes 26 (1), 39
55.

Bohn, H.L., Brian, L.M., and O Connor, G.A., (1979), Chapter 8: Salt-Affected Soil,
Soil Chemistry, John Wiley and Sons, New York, N.Y., pp. 217-245

Brianon, L., Villard, P., (2008). Design of Geosynthetic-Reinforced Platforms


Spanning Localized Sinkholes, Geotextiles and Geomembranes 26 (5), 416428.
343

BS8006, 1995. British Standard, Code of Practice for Strengthened/reinforced Soils and
Other Fills. BSI.

Budhu M., (2008), Foundations and Earth Retaining Structures, Book, Wiley ,ISBN
978-0-471-47012-0.

Castellanos, B.A., (2010), Internal Design of Mechanically Stabilized Earth (MSE)


Retaining Walls Using Crimped Bars. Master Thesis, Utah State University, Dept. Civil
and Environmental Engineering Logan, Utah.

Chattopadhyay B. C., Chakravarty S., (2009), Application of Jute Geotextiles as


Facilitator in Drainage Geotextiles and Geomembranes, Volume 27, Issue 2, Pages 156-
161.

Chen, Q., Abu-Farsakh, M., Sharma, R., Zhang, X., (2007), Laboratory Investigation of
Behavior of Foundations on Geosynthetic-Reinforced Clayey Soil, Transportation
Research Record: Journal of the Transportation Research Board 2004, 2838.

Chen W.F., (2007), Plasticity in Reinforced Concrete, J. Ross Publishing


Chen, W.F., (2008), Limit Analysis and Soil Plasticity, J. Ross Publishing

Chen, Y.-M., Cao, W.-P., Chen, R.-P., (2008), An Experimental Investigation of Soil
Arching Within Basal Reinforced and Unreinforced Piled embankments, Geotextiles
and Geomembranes 26 (2), 164174.

Chew, S.H., Schmertman, G.R., and Mitchell, J.K., (1990), Reinforced Soil Wall
Deformations by Finite Element Method, Proc. Int. Reinforced Soil Conference:
Performance of Reinforced Soil Structures, K.Z. Andrawes, and A. McGown, Eds.,
Thomas Telford, Glasgow, Scottland, pp.34-40.

Christopher B. R., Berg R. R., and Elias V., (2001), Mechanically Stabilized Earth
Walls and Reinforced Soil Slopes. Design and Construction Guidelines, U.S. Dept. of
Transportation, Federal Highway Administration, Publication No. FHWA-NHI-00-044,
p.g. 54.

Christopher B.R., Zornberg J.G., and Mitchell J.K., (1998), Design Guidance for
Reinforced Soil Structures with Marginal Soil Backfills, In: Sixth International
conference onGeosynthetics, Atlanta, GA, USA, 1, 797e804.

Chu, J., Bo, M.W., Choa, V., (2004), Practical Considerations for Using Vertical Drains
in Soil Improvement Projects, Geotextiles and Geomembranes 22, 101117.
344

Chu, J., Bo, M.W., Choa, V., (2006), Improvement of Ultra Soft Soil Using
Prefabricated Vertical Drains, Geotextiles and Geomembranes 24, 339348.

Coyne, M. A., (1927), Murs de Soutenement et Murs de Quai a Echelle, Le Genie Civil,
Tome XCL, no. 16, 29 Oct, Paris.

Cundall, P.A., Strack, O.D.L., 1979. A discrete element model for granular assemblies.
Geotechnique 29 (1), 4765.

Dias, A.C., (2003), Numerical Analyses of SoilGeosynthetic Interaction in Pull-out


Tests, M.Sc. thesis, University of Brasilia, Brasilia, Brazil, 115 pp.

De Gennaro V., and Frank R. (2002), Elasto-plastic analysis of the interface behaviour
between granular media and structure , Computers and Geotechnics, Volume 29, Issue 7,
October 2002, Pages 547-572.

DeMerchant, M.R., Valsangkar, A.J., Schriver, A.B., (2002), Plate Load Tests on
Geogridreinforced Expanded Shale Lightweight Aggregate, Geotextiles and
Geomembranes 20, 173190.

Desai, C.S., (1980) A General Basis for Yield, Failure and Potential Functions in
Plasticity, Int. J. Numer. Analy. Methods Geomech., 4, 361-375.

Desai C.S., (1981), Behavior of Interfaces Between Structural and Geological Media, State
of the Art Paper, Proceeding Int. Conf. on Recent Advances in Geotech. Earthquake Engrg.
and Soil Dynamics, St. Louis, Missouri.

Desai, C.S., (1994), Hierarchical Single Surface and the Disturbed State Constitutive
Models with Emphasis on Geotechnical Applications, Chapter 5 in Geotechnical
Engineering, K.R. Saxena (editor), Oxford and IBH Publ. Co., New Delhi, India.

Desai, C.S., (1995), Constitutive Modelling Using the Disturbed State as Microstructure
Self-Adjustment Concept, Chapter 8 in Continuum Models for Materials with
Microstructures, H.B. Muhlhaus, John Wiley, U.K.

Desai, C. S. (2001), Mechanics of materials and interfaces: The disturbed state concept,
CRC, Boca Raton, Fla.

Desai C.S., (2005a), Prediction of Field Behavior of Reinforced Soil Wall Using Advanced
Constitutive Model, Journal of Geotechnical and Geo-environment Engineering ASCE,
2005.
345

Desai C.S., (2005b), Constitutive Modeling for Geologic Materials: Significance and
Directions International Journal of Geomechanics ASCE, Vol. 5, No. 2, pp. 81-84.

Desai, C.S. and El-Hoseiny, K.E., (2005), Prediction of Field Behavior of Reinforced
Soil Wall Using Advanced Constitutive Model, Journal of Geotechnical and
Geoenvironmental Engineering, 131, no. 6, pp. 729-739.

Desai, C.S. and Faruque, M.O., (1984) Constitutive Model for geologic Materials, J.
Eng. Mech., ASCE, 110, 9, 1391-1408.

Desai, C.S. and Fishman, K.L., (1991) Plasticity Based Constitutive Model with Associated
Testing for Joints, Int. J. Rock Mech. & Min. Sc., 4, 28, 1526.

Desai, C.S., and Hashmi, Q., (1985), Procedure for Finite Element Implementation of a
Non-associative Constitutive Model, Proc., Symp. on Computational Mechanics, Joint
ASCE-ASME Conf., Albuquerque, N.M.

Desai, C.S., and Hashmi, Q., (1989), Analysis, Evaluation and Implementation of
Nonassociative Model for Geologic Materials, Int. J. Plasticity, 5, 1989, 397-420.

Desai, C.S., and Ma Y., (1990), Constitutive Modeling of joint and Interfaces by Using
Disturbed State Concept, International Journal for Numerical and Analytical methods in
Geomechanics, Vol. 16, pp. 623-653.

Desai, C.S., and Nagarj B.K, (1998), Modeling for Cyclic Normal and Shear Behavior of
Interfaces, Journal of Engineering Mechanics, Volume 114, No. 7, pp. 1198-1217.

Desai, C., Pradhan, S., and Cohen, D. (2005). Cyclic Testing and Constitutive Modeling
of Saturated SandConcrete Interfaces Using the Disturbed State Concept. Int. J.
Geomech., Vol 5, No. 4, pp 286294.

Desai, C.S., and Rigby, D.B., (1997), Cyclic Interface and Joint Shear Device Including
Pore Pressure Effects, Journal of Geotechnical and Environmental Engineering, Vol.
123, No. 6, pp. 568-579.

Desai, C.S., Sharma, K.G., Wathugala, G.W., and Rigby, D., Implementation of
Hierarchical Single Surface 0 and 1 Models in Finite Element Procedure, Int. J.
Num.Analyt. Meth. In Geomech., 15, 1991, 649-680.

Desai, C.S., and Siriwardan, H.J., (1984), Constitutive Laws for Engineering Materials,
Prentice Hall, Inc., Englewood Cliffs, New Jersey, U.S.A.
346

Desai, C.S., Somasundaram, S., and Frantzishonis, G., (1986), A Hierarchical Approach
for Constitutive Modelling of Geologic Materials,Int. J. Numer. Anal. Methods in
Geomech., 10, 3, 225-257.

Desai, C.S. and Wathugala, G.N., (1987), Hierarchical and Unified Models for solids
and Discontinuities, Notes for Short Course, Tucson, Arizona, USA.

Desai, C.S., Zaman, M.M., Lightner, J.G., and Siriwardane, H.J., (1984), Thin-Layer
Element for Interfaces and Joints, Int. J. Num. Analyt. Meth. in Geomech., 8, 1, 1943.

Dudley, J.H., (1970), Review of Collapsing Soils, The Journal of the Soil Mechanics
and Foundations Division, May/June., pp. 925-945

Duvall, D. E., (1993), Creep and Stress Rupture Testing of a Polyethylene


Geomembrane Under Equal Biaxial Tensile Stress, Proceedings of the Geosynthetics
'93 Conference, Vol. 2. Industrial Fabrics Association International, MN, USA, 1993, pp.
817-30.

Dyer, M.R., 1985. Observation of the Stress Distribution in Crushed Glass with
Applications to Soil Reinforcement. Ph.D. thesis, University of Oxford, UK.

Elias V., Christopher B.R., and Berg R.R., (2001), Mechanically Stabilized Earth Walls
and Reinforced Soil Slopes-design and Construction Guidelines, FHWA-NHI-00-043,
Federal Highway Administration, Washington, DC, USA.

El-Hoseiny, K.E, (1999), Soil Stabilization Using Geosynthetics PhD Dissertation,


Menoufia University, Shibin El Kom, Egypt.

El-Naggar M. E., Kennedy J.B. (1997), New Design Method for Reinforced Sloped
Embankments, Engineering Structures, Volume 19, Issue 1, January 1997, Pages 28-36.

Ellis, C.I., (1973), Arabian Salt Bearing Soil as an Engineering Material, Transport
and Road Research Laboratory (TRRL), Report LR 523, Crowthorne.

Fishman, K.L., and Desai, C.S., (1991), Response of a Geogrid Earth Reinforced
Retaining Wall with Full Height Precast Concrete Facing, Geosynthetics 91 Conference,
Atlanta, GA, pp. 691-700.
347

Fishman, K.L., Desai, C.S. and Sogge, R.L., (1993), Field Behavior of Instrumented
Geogrid Soil Reinforced Wall, Journal of Geotechnical Engineering, ASCE, Vol. 119,
No. 8, August, pp. 1293-1307.

Fishman, K.L., Berg, R.R., and Desai, C.S., (1991), Geosynthetic-Reinforced Soil Wall:
4-Year History, Transportation Research Record No. 1330, Behavior of Jointed Rock
Masses and Reinforced Soil Structures, Transportation Research Board, Washington
D.C., pp. 30-39.

Fredlund, D.G., Rahardjo, H., 1993. Soil Mechanics for Unsaturated Soils. John Wiley,
USA.

Fujino Y., Sasaki Y., and Hakuno M., (1978), Slip of a Friction Controlled Mass
Excited by Earth Motions, Bulletin of Earthquake Research Institute, University of
Tokyo, Vol. 53, Part2.

Ghazavi, M., Lavasan, A.A., (2008), Interference Effect of Shallow Foundations


Constructed on Sand Reinforced with Geosynthetics, Geotextiles and Geomembranes 26,
404415.

Girard H., Fischer S., Alonso E., (1990) Problems of Friction Posed by the Use of
Geomembranes on Dam SlopesExamples and Measurements, Original Research
Article Geotextiles and Geomembranes, Volume 9, Issue 2, Pages 129-143

Goldfingle, G., 2009. Float On. Ground Engineering, British Geotechnical Association,
UK, October issue, pp. 32e34.

Geosynthetics magazine, 2010, Pipe Structure, [Online Image] at:


http://geosyntheticsmagazine.com/repository/2/2171/large_0608_f5_3.jpg, 09/14/2010

Harrison W. J., and Gerrard C. M., (1972), Elastic Theory Applied to Reinforced Earth.
Journal of the Soil Mechanics and Foundation Division of the American Society of Civil
Engineers, 98, SM12, December.

Helwany, S.M.B., Wu, J.T.H., Froessl, B., (2003), GRS Bridge AbutmentsAn
Effective Means to Alleviate Bridge Approach Settlement, Geotextiles and
Geomembranes 21, 177196.

Hsuan Y.G. and Li M., (2005), Temperature and Pressure Effects on the Oxidation of
High-Density Polyethylene Geogrids, Geotextiles and Geomembranes, Volume 23,
Issue 1, Pages 55-75.
348

Hufenus R., Regger R., Flum D., and Sterba I. J., (2005), Strength reduction factors
due to installation damage of reinforcing geosynthetics, Geotextiles and Geomembranes,
Volume 23, Issue 5, Pages 401-424

Ibragimov, M., (2005), Soil Stabilization with Cement Grouts, The Journal of the Soil
Mechanics and Foundations, 42, no. 2 (2005): 67-72.

Jirasek, M., and Bazant, Z. P., (2002), Inelastic Analysis of Structures, John Wiley and
Sons.

Juran, I., Guermazi, A., Chen, C. L., and Ider, M. H., (1988), Modeling and Simulation of
Load Transfer in Reinforced Soil, Part 1. International Journal of Numerical Analysis
and Methods in Geomechanics, 12, 141-155.

Katti, D.R., and Desai, C.S., (1998), Modeling of Cohesive Soil Subject to Cyclic
Loading Using Disturbed State Concept, ASCE- Geotechnical Earthquake Engineering
and Soil Dynamics conference, Seattle.

Kazimierowicz-Frankowska, K., (2007), Influence of Geosynthetic Reinforcement on


The Load-Settlement Characteristics of Two-Layer Subgrade, Geotextiles and
Geomembranes 25 (6), 366376.

Khedkar M. S., Mandal J.N., (2009), Pullout Behaviour of Cellular Reinforcements,


Original Research Article Geotextiles and Geomembranes, Volume 27, Issue 4, Pages
262-271.

Koerner, R.M., 2005. Designing with Geosynthetics, fifth ed.. Pearson Prentice Hall,
Upper Saddle River, NJ, USA.

Knight, K., (1963), The Origin and Occurance of Collapsing Soils, Proceedings, 3rd
Regional Conference for Africa on Soil Mechanics and Foundation Engineering, Vol. 1,
pp. 127-130.

Lawson C., (2005), Geosynthetic Reinforced MSE Walls and Slopes with Fine Grained
Fills: International Perspectives, NAGS 2005/GRI-19 Conference, Las Vegas, Nevada,
USA.

Lennoz-Gratin, C. H., (1987), The Use of Geotextiles and Drain Envelopes in France in
Connection with Mineral Clogging, Geotextiles and Geomembranes, 5(2), 71-90.

Leveque, J.L., and Agache, P., (1993), Aging Skin:Properties and Functional Changes,
Marcel Dekker.
349

Li, A.L., Rowe, R.K., (2008), Effects of Viscous Behaviour of Geosynthetic


Reinforcement and Foundation Soils on Embankment Performance, Geotextiles and
Geomembranes 26 (4), 317334.

Liu, S.Y., Han, J., Zhang, D.W., Hong, Z.S., (2008), A Combined DJM-PVD Method
for Soft Ground Improvement, Geosynthetics International 15 (1), 4354.

Lo, K. Y., Micic, S., Shang, J. Q., Lee, Y. N., and Lee, S. W. (2000) Electrokinetic
Strengthening of a Soft Marine Sediment, International Journal of Offshore and Polar
Engineering, 10(2).

Lorenzo, G.A., Bergade, D.T., Burthai, W., Hormdee, D., Phothiraksanon, P., (2004),
Innovations and performance of prefabricated vertical drain and dual function
geosynthetic applications, Geotextiles and Geomembranes 22, 7599.

Lu N., and Likos W.J., (2004), Unsaturated Soil Mechanics. Wiley, New Jersey, USA.

Lubliner, J., (2008), Plasticity Theory, Dover, ISBN 0486462900, 9780486462905.


McCartney J., Berends R. E., (2010), Measurement of Filtration Effects on the
Transmissivity of Geocomposite Drains for Phosphogypsum, Original Research Article
Geotextiles and Geomembranes, Volume 28, Issue 2, Pages 226-235.

Markham Culverts, 2010, Sea Wall, [Online Image] at:


http://www.markham.com.pg/images/BidimPic2.jpg, 09/14/2010.

Mitchell, J.K., Zornberg, J.G., 1995. Reinforced soil structures with poorly draining
backfills. Part II: case histories and applications. Geosynthetics International 2 (1),
265e307.

Musser S.W., and Denning, C., (2005), Deep Patch Road Embankment Repair
Application Guide, United States Department of Agriculture Forest Service. Technology
& Development Program 7700. Transportation Management, October 2005, 0577 1204-
SDTDC.

Nakai T., (1985), Analysis of Earth Pressure Problems Considering the Influence of
Wall Friction and Wall Deflection, Proc. 5th International Conference on Numerical
Methods in Geomechanics, pp.765-772.

Narejo D. B., (2003), A Simple Tilt Table Device to Measure Index Friction Angle of
Geosynthetics, Geotextiles and Geomembranes, Volume 21, Issue 1, February 2003,
Pages 49-57.
350

Nataraj, M. S., Maganti, R. S., McManis, K. L., (1995), Interface Frictional


Characteristics of Geosynthetics, In Proceedings of the Geosynthetics, 95 Conference,
IFAI, Nashville, TN, Vol. 3, pp. 1057-1069.

NCMA, (2002), Design Manual for Segmental Retaining Walls, second ed. National
Concrete Masonry Association, Herndon, VA, USA.

Noorzad R., Mirmoradi S.H., (2010), Laboratory Evaluation of the Behavior of a


Geotextile Reinforced Clay, Geotextiles and Geomembranes, Volume 28, Issue 4, Pages
386-392.

Pal, S., (1997), Numerical Simulation of Geosynthetic Reinforced Earth Structures Using
Finite Element Method, PhD Dissertation, Louisiana State University, Louisiana.

Palmeira, E.M., (1987), The Study of SoilReinforcement Interaction by Means of


Large Scale Laboratory Tests, Ph.D. thesis, University of Oxford, UK, pp 238.

Palmeira, E., (2009), Soil-Geosynthetic Interaction: Modelling and Analysis, Original


Research Article Geotextiles and Geomembranes, Volume 27, Issue 5, Pages 368-390.

Palmeira, E.M., Lima Jr., N.R., Mello, L.G.R., (2002), Interaction between soils and
geosynthetic layers in large-scale ramp tests, Geosynthetic s International 9(2) (IFAI,
USA), 149187.

Palmeira, E.M., (2009), SoilGeosynthetic Interaction: Modelling and Analysis,


Geotextiles and Geomembranes 27, 368390.

Phanikumar B.R., Prasad R., Singh A., (2009), Compressive Load Response of Geogrid-
Reinforced Fine, Medium and Coarse Sands, Original Research Article Geotextiles and
Geomembranes, Volume 27, Issue 3, Pages 183-186.

PLAXIAS b. v. (2006), Plaxis V8 manual, PLAXIAS b. v, Delft, Netherlands.

Powel W., Keller G.R., and Brunette B., (1999), Applications for Geosynthetics on
Forest Service Low-Volume Roads, Transportation Research Record 1652, 113e120.

Sandri D., (2005), Drainage Recommendations for MSE Walls Constructed with
Marginal Fills, NAGS 2005/GRI-19 Conference, Las Vegas, Nevada, USA.
351

Shehata, W.M., Al-Saafin, A.K., Harari, Z.Y. and Bader, T.A., (1990), Potential Sabkha
Hazards in the Saudi Arabia, Proc. 6th Int. EAEG Congr., D.G. Price, Edt., A.A.
Balkema, Amsterdam, the Netherlands, pp. 2003-2010.

Skinner, H., 2010, Pipe Structure, [Online Image] at:


http://www3.imperial.ac.uk/portal/pls/portallive/docs/1/6981920.JPG, 09/14/2010

Smith, M.E., (2008), Emerging Issues in Heap Leaching Technology, In: Proceedings
4th European Geosynthetics Conference, Edinburgh, U.K (CD-Rom).

Stulgis R.P., (2005), Full-scale MSE Test Walls, NAGS 2005/GRI-19 Conference, Las
Vegas, Nevada, USA.

Sugimoto, M., Alagiyawanna, A.N.M., and Kadoguchi, K., (2001), Influence of rigid
and flexible face on geogrid pullout tests, Geotextiles and Geomembranes 19 (5), 257
277.

Raisinghani D.V., and Viswanadham B.V.S, (2010), Evaluation of Permeability


Characteristics of a Geosynthetic-Reinforced Soil Through Laboratory Tests,
Geotextiles and Geomembranes, In Press, Corrected Proof, Available online 22 January
2010.

Retaining Wall. (n.d.). In Wikipedia. Retrieved March 21, 2012 3:06:10 PM from
http://en.wikipedia.org/wiki/Retaining_wall

Richaud E., Farcas F, Divet L, and Benneton J. P., (2008), Accelerated ageing of
polypropylene geotextiles, the effect of temperature, oxygen pressure and aqueous media
on fibersMethodological aspects, Geotextiles and Geomembranes, Volume 26, Issue
1, Pages 71-81.

Roering, J., Schmidt, K., Stock J., Dietrich, W., Montgomery, D., (2003), Shallow
landsliding, root reinforcement, and the spatial distribution of trees in the Oregon Coast
Range, Canadian Geotechnical Journal, Vol 40, Pages 237-253.
Rollin, A. L., (1987), Guest editorial, Geotextiles and Geomembranes, 5(2) 69.

Roscoe, K. H.; Schofield, A. N.; Wroth, C. P. (1958), "On the Yielding of


Soils", Geotechnique 8: 2253

Rowe, R.K., Taechakumthorn, C., (2008), Combined Effect of PVDs and Reinforcement
on Embankments Over Rate-Sensitive Soils. Geotextiles and Geomembranes 26 (3), 239
249.
352

Rowe, R.K., and Ho, S.K., (1992), A Review of the Behavior of Reinforced Soil Walls,
Preprint, Special and Keynote Lecture, Int. Symp. On Earth Reinforcement Practice, A.A.
Balkema, Rotterdam, pp. 47-76.

Saadatmanesh, H., Tavakkolizadeh, M., Mostofinejad, D., (2010), Environmental effects


on mechanical properties of wet lay-up fiber-reinforced polymer, ACI Materials Journal,
Vol. 107, No. 3, pp. 267-274.

Sarsby, R.S., (2007), Use of Limited Life Geotextiles (LLGs) for Basal Reinforcement
of Embankments Built on Soft Clay, Geotextiles and eomembranes 25 (45), 302310.

Schofield, A. N.; Wroth, C. P. (1968), Critical State Soil Mechanics, McGraw-Hill,


pp. 310, ISBN 978-0641940484

Schofield, A. N. (2006), Disturbed Soil Properties and Geotechnical Design, Thomas


Telford, pp. 216, ISBN 978-0727729828

Seed, R.B., Collin, J.G., and Mitchell, J.K., (1986), FEM Analysis of Compacted
Reinforced Soil Walls, Proc. 2nd Int. Conf. on Numerical Models in Geomech., G.N.
Pande and W.F. Van Impe, Eds., M. Jackson and Son Publisher Redruth, Cornwall,
England, pp. 553-562.

Sharma, K.G. and Desai, C.S., (1992), An Analysis and Implementation of Thin-Layer
Element for Interfaces and Joints, J. of Eng. Mech., ASCE, 118, 545569.

Sherif W. A., and Colin J. F. P., (1996), Design of reinforced fill systems to support
footings overlying cavities,Geotextiles and Geomembranes, Volume 14, Issue 1, Pages
57-72.

Shivashankar, R., Dheerendra Babu, M.R., Sitaram Nayak and Manjunath, R., (2010),
Stone Columns with Vertical Circumferential Nails: Laboratory Model Study,
Geotechnical and Geological Engineering, Vol 28, No. 5, Pages 695-706

Sokolovich, V.E., (1987), Chemical Soil Stabilization and the Environment, The
Journal of the Soil Mechanics and Foundations, Vol 24, No. 6, Pages 233-236.

Stuyt, L. C. P. M. & Oosten, A. J., Mineral and ochre clogging of subsurfacel and
drainage systems in The Netherlands. Geotextiles and Geomembranes, 5(2) (1987) 123-
40.

Sychev, A.K., (1954), Laboratory study of stabilization of loess by the thermal method,"
Vestn., VIA, No. 78, VIA, Moscow.
353

Tannous, F.E. (1997). Durability of Non-metallic Reinforcing Poars and Prestressing


Tendons, PhD Dissertation, Department of Civil Engineering and Engineering
Mechanics, University of Arizona, Tucson, Arizona, June, 287 pages.

Tavakolizadeh, M.R., and Saadatmanesh, H.(2004). Environmental Effects on Tensile


Properties of FRP Laminates Made Using Wet-Lay-Up Method, Proceedings of The
Advanced Polymer Composites for Structural Applications in Construction (ACIC),
University of Surrey, England, pp. 617-632.

Tensar SR2 Reinforced Retaining Walls for Tanque Verde Road, Tucson, Arizona.
(1984), Value Engineering Study, Jon No. 13780-002-22, Pima County Department of
Transportation and Flood Control District, Tucson, AZ.

Totalpetrochemicals Inc, (2008), Reinforced embankments for highways. [Online Image]


at http://www.totalpetrochemicals.biz/content/documents/212_dscn0629.jpg, 09/09/10.

Toth, J., and Desai, C.S., (1994) Development of lunar ceramic composites testing and
constitutive modeling including cemented sand, Report, Dept. of Civil Eng. And Engng.
Mechanics, University of Arizona, Tucson, AZ, USA.

Undergroundozarks Inc, 1999, Reinforced Pack for Roof Supporting Underground


Mining, [Online Image] at: http://undergroundozarks.com/blog/media/campout03.jpg,
09/14/2010.

Uwabe T. and Moriya H., (1987), Sliding Behavior of Gravity Retaining Walls During
Earthquake, Proc. 22nd Japan National Conference on SMFE, Niigata, pp. 713-714.

Vidal, H., (1969), The Principle of Reinforced Earth, Highway Research Record, No.
282, USA

Villemus B., Morel J.C, and Boutin, C., (2007), Experimental assessment of dry stone
retaining wall stability on a rigid foundation, Engineering Structures, Volume 29, Issue
9, September 2007, Pages 2124-2132

Watanabe H., and Tochigi H., (1986), Model Vibration Tests Concerning the Dynamic
Interaction of Soil Structure System Followed by Sliding and Separation and their
Numerical Simulation, Proc. Of JSCE, No. 368, pp. 319-327.
354

Wathugala, G.W., and Desai, C.S., (1989), Damage Based Constitutive Model for Soil,
Proceeding 12th Canadian Congress of Application Mech., Ottawa, Canada, pp. 530-531.

Wesseloo J., Visser A. T., and Rust E., (2004), A Mathematical Model for the Strain-
Rate Dependent StressStrain Response of HDPE Geomembranes, Original Research
Article Geotextiles and Geomembranes Volume 22, Issue 4, Pages 273-295.

Westergaard, H. M., (1938), A Problem of Elasticity Suggested by a Problem in Soil


Mechanics: Soft Material Reinforced by Numerous Strong Horizontal Sheet,
Contributions to the mechanics of solids dedicated to Stephen Timoshenko by his friends
on the occasion of the sixtieth birthday anniversary, the Macmillan Company, New York.

Wu, J.T.H., (1992), Predicting Performance of the Denver Walls: General Report, in
Geosynthetic Reinforced Soil Retaining Walls, A.A. Balkema, pp. 3-20.

Wu, W., Wick, H., and Ferstl, F., Aschauer, F., (2008), A Tilt Table Device for Testing
Geosynthetic Interfaces in Centrifuge, Geotextiles and Geomembranes 26 (1), 3138.

Yang, Y., (2007), The Other Soil Parameters in Stability Limit Analysis of Soil-Nailed
Walls in Soft Soil, Soft Soil Engineering, 573-577.

Yasuhara, K., Recio-Molina, J., (2007), Geosynthetic-Wrap Around Revetments for


Shore Protection, Geotextiles and Geomembranes, Volume 25, Issues 4-5, August-
October 2007, Pages 221-232.

Youssef, A.A., (1998), Suggested Research No. 154: General Studies for Soft Clay and
Sabka Soils in Egypt, Submitted to Egyptian Academic Research for Science and
Technology, University of Menoufia, Shebin El-Kom, Egypt.

Zhuan-zhi H., (2003), Sliding resistance of cushion reinforced with geosynthetic and its
analysis method of stability, Post proc. of Chinese symposium on reinforced soil
engineering, Modern Knowledge Press, Hong Kong 2003, 152-165.

Potrebbero piacerti anche