A NOVEL APPROACH FOR SIMULATING SOIL AND PIPE …

239
A NOVEL APPROACH FOR SIMULATING SOIL AND PIPE RESPONSE TO SEASONAL, ENVIRONMENTAL AND FIELD CONDITIONS A Thesis Submitted to the Faculty of Graduate Studies and Research In Partial Fulfillment of the Requirements For the Degree of Doctor of Philosophy in Environmental Systems Engineering University of Regina By Ramy M. Saadeldin Regina, Saskatchewan April 2016 Copyright 2016: R. Saadeldin

Transcript of A NOVEL APPROACH FOR SIMULATING SOIL AND PIPE …

A NOVEL APPROACH FOR SIMULATING SOIL AND PIPE RESPONSE TO

SEASONAL, ENVIRONMENTAL AND FIELD CONDITIONS

A Thesis

Submitted to the Faculty of Graduate Studies and Research

In Partial Fulfillment of the Requirements

For the Degree of

Doctor of Philosophy

in

Environmental Systems Engineering

University of Regina

By

Ramy M. Saadeldin

Regina, Saskatchewan

April 2016

Copyright 2016: R. Saadeldin

UNIVERSITY OF REGINA

FACULTY OF GRADUATE STUDIES AND RESEARCH

SUPERVISORY AND EXAMINING COMMITTEE

Ramy M. Saadeldin, candidate for the degree of Doctor of Philosophy in Environmental Systems Engineering, has presented a thesis titled, A Novel Approach for Simulating Soil and Pipe Response to Seasonal, Environmental and Field Conditions, in an oral examination held on April 12, 2016. The following committee members have found the thesis acceptable in form and content, and that the candidate demonstrated satisfactory knowledge of the subject material. External Examiner: *Dr. Mohammed Sakr, Machibroda Engineering

Supervisor: Dr. Amr Henni, Industrial Systems Engineering

Committee Member: **Dr. Yee-Chung Jin, Environmental Systems Engineering

Committee Member: Dr. Ezeddin Shirif, Petroleum Systems Engineering

Committee Member: Dr. Osman Salad Hersi, Department of Geology

Chair of Defense: Dr. Mark Brigham, Department of Biology *Via Skype **Not present at defense

i

ABSTRACT

Climate change related problems are increasing in occurrence and severity leading

to significant economic losses in many places of the world. In semi-arid environments,

like Saskatchewan, the main phenomenon involved in pipe breakages is the volume

change behavior of unsaturated clay deposits. Underground pipelines are typically

buried within the upper zone of soil deposits, and therefore, are highly affected by soil

nature and the different environmental conditions present on the ground surface. To

accurately model field conditions, a mathematical formulation of native soil conditions

was developed based on laboratory and field test results. Volume-mass constitutive

surfaces were established based on a bimodal soil water characteristic curve (SWCC)

and other constitutive relationships.

In order to simulate the response of soil and pipe to various meteorological

conditions, a numerical framework was developed and validated. The strength of the

developed numerical framework lays in the use of bimodal SWCC and modeling the

hydraulic characteristics of a cracked soil structure. This research study also utilized, as

a database, the results of a field instrumentation program conducted in the City of

Regina. A hydro-mechanical analysis was implemented to model the volume change

due to variations in mechanical loading conditions and moisture content. Modeling

scenarios were also studied based on variations in pipe diameter, pipe depth and soil

elasticity.

The developed numerical framework provided insight into the sensitivity of pipe

deformation to possible changes in input parameters of the soil-pipe system. The model

was able to capture the transient water flow through saturated-unsaturated soils. The

results of the modeling of weather conditions applied on the soil-pipe system were in

agreement with the field measurements. Specific relationships between the soil-pipe

ii

interaction and seasonal changes in the local meteorological conditions were

established. The model was also used to provide some insight into the real flux

transferred through the pavement structure to the backfill material surrounding the pipe.

Finally, soil and pipe reactions (i.e. soil and pipe displacements, soil volumetric water

content and soil temperature) to applied surface boundary conditions were predicted

based on the validated numerical approach.

iii

ACKNOWLEDGEMENTS

I would like to express my sincere appreciation to my supervisor, Dr. Amr Henni,

for his precious support and continuous encouragement throughout the course of this

research program. I would like to thank, Dr. Yafei Hu, for his overall guidance and critical

appraisal. I would also like to extend my acknowledgement to my advisory committee

members for their valuable suggestions and input.

The financial support provided by the Faculty of Graduate Studies and Research

(FGSR) at the University of Regina, and the City of Regina (Henry Baker Scholarship

Program) is gratefully acknowledged. In addition, I would like to acknowledge the

contribution of the National Research Council Centre for Sustainable Infrastructure

Research for allowing access to their research facilities.

Many thanks, to my colleagues and staff at the Faculty of Engineering and Applied

Science, for making my stay at the University of Regina an unforgettable one. I would

also like to thank Mr. Gene Froc who gave me invaluable support that helped me to

reach this point.

Last but not the least, I am extremely grateful to my parents, wife, sister and

brothers for their continuous support, and to my children for providing me with the

motivation to complete my thesis.

iv

TABLE OF CONTENTS

ABSTRACT..................................................................................................................... I

ACKNOWLEDGEMENTS ..............................................................................................III

TABLE OF CONTENTS ................................................................................................ IV

LIST OF FIGURES ....................................................................................................... VII

LIST OF TABLES ........................................................................................................ XII

CHAPTER 1: INTRODUCTION ...................................................................................... 1

1.1 Definitions ................................................................................................................. 1

1.2 Problem Recognition ................................................................................................ 5

1.3 Engineering Significance .......................................................................................... 6

1.4 Research Objectives ................................................................................................ 8

1.5 Finite Element Approach .......................................................................................... 9

1.6 Contributions ...........................................................................................................13

1.7 Outline of this Research ..........................................................................................15

CHAPTER 2: LITERATURE REVIEW...........................................................................17

2.1 Surficial Geology .....................................................................................................17

2.2 Climate Conditions ..................................................................................................19

2.3 Expansive Soils .......................................................................................................22

2.4 Clay Mineralogy .......................................................................................................25

2.5 Unsaturated Soil Parameters ...................................................................................30

2.5.1 Stress State Variables .......................................................................................30 2.5.2 Soil Suction .......................................................................................................34

2.6 Desiccation Cracks ..................................................................................................36

2.7 Hydraulic Conductivity Function ...............................................................................41

2.8 Soil Water Characteristic Curve (SWCC) .................................................................42

2.9 Measurements of Soil Moisture-suction Characteristics ...........................................45

2.9.1 Foreword ...........................................................................................................45 2.9.2 Soil Moisture Monitoring (Principle and Techniques) .........................................45 2.9.3 Soil Suction Monitoring (Principle and Techniques) ...........................................47

2.10 Unsaturated Soil-atmosphere Interaction ...............................................................51

2.10.1 Foreword .........................................................................................................51 2.10.2 Soil Evaportranspiration ..................................................................................51 2.10.3 Water Flow ......................................................................................................52 2.10.4 Heat Flow ........................................................................................................55

2.11 Unsaturated Soil-structure Interaction ....................................................................56

2.11.1 Numerical Approaches ....................................................................................56 2.11.2 Volume-mass Constitutive Relationships ........................................................57

2.12 Pipelines Infrastructure ..........................................................................................63

2.13 Applied Loads on Buried Pipes ..............................................................................64

2.13.1 Foreword .........................................................................................................64

v

2.13.2 Soil Load .........................................................................................................65 2.13.3 Live Loads.......................................................................................................67

2.14 Pipe Stresses ........................................................................................................68

2.15 Pipe Deformations .................................................................................................70

2.15.1 Horizontal Deformation....................................................................................70 2.15.2 Bending Displacement ....................................................................................71

2.16 Buried Pipe Damages ............................................................................................73

2.16.1 Failure Mechanisms Associated with Soil Movements.....................................73 2.16.2 Case Studies ...................................................................................................75

2.17 Related Numerical Modeling Studies in the Literature ............................................76

CHAPTER 3: FIELD INVESTIGATION .........................................................................80

3.1 General ...................................................................................................................80

3.2 Field Program Details ..............................................................................................81

3.3 Instrumentation Details ............................................................................................84

3.4 Climate Data ............................................................................................................86

3.5 Backfill and Bedding Soils .......................................................................................90

3.6 Soil Profile and Properties at the Field Site ..............................................................92

3.7 Unsaturated Soil Characteristics ..............................................................................98

3.8 Soil Moisture Field Data ......................................................................................... 102

3.9 Air and Soil Temperatures ..................................................................................... 104

3.10 Measured Pipe Displacements ............................................................................ 104

CHAPTER 4: LOAD-DEFORMATION ANALYSIS UNDER DIFFERENT SOIL AND PIPE CONDITIONS ..................................................................................................... 107

4.1 Problem Statement ................................................................................................ 107

4.2 Governing Equations ............................................................................................. 107

4.3 Modeling Overview, Geometry and Boundary Conditions ...................................... 109

4.4 Numerical Modeling versus Analytical Results ....................................................... 113

4.5 Pipe Deformations ................................................................................................. 115

CHAPTER 5: PIPE RESPONSE TO RELATIVE SATURATION OF SURROUNDING SOIL ............................................................................................................................ 120

5.1 Problem Statement ................................................................................................ 120

5.2 Governing Equations ............................................................................................. 120

5.3 Modeling Overview, Geometry and Boundary Conditions ...................................... 122

5.4 Numerical Modeling versus Analytical Results ....................................................... 127

5.5 Soil Saturation - Pipe Displacement Analysis ........................................................ 127

CHAPTER 6: MATHEMATICAL FORMULATIONS OF SOIL-WATER INTERACTION UNDER UNSATURATED SOIL CONDITIONS ........................................................... 134

6.1 Problem Statement ................................................................................................ 134

6.2 Methodology .......................................................................................................... 135

6.3 Fracture Depth Formulation ................................................................................... 135

6.4 Hydraulic Conductivity Formulation ........................................................................ 142

vi

6.5 Development of a Bimodal SWCC ......................................................................... 143

6.6 Volume-mass Constitutive Relationships ............................................................... 148

6.7 Water Flow Mathematical Formulation ................................................................... 157

6.8 Parametric Study ................................................................................................... 157

6.8.1 General ........................................................................................................... 157 6.8.2 Methodology ................................................................................................... 158 6.8.3 Results and Discussion ................................................................................... 160

CHAPTER 7: SOIL-PIPE-ATMOSPHERE INTERACTION UNDER FIELD CONDITIONS .............................................................................................................. 168

7.1 Problem Statement ................................................................................................ 168

7.2 Methodology .......................................................................................................... 168

7.3 Geometry and Boundary Conditions ...................................................................... 170

7.4 Mathematical Formulation ..................................................................................... 174

7.4.1 Evapotranspiration Process ............................................................................ 174 7.4.2 Water Flow Equations ..................................................................................... 175 7.4.3 Stress-strain Equations ................................................................................... 176

7.5 Climate Data .......................................................................................................... 177

7.6 Pavement Boundary Conditions ............................................................................ 178

7.7 Soil Temperature Analysis Results ........................................................................ 178

7.8 Soil Moisture Analysis Results ............................................................................... 182

7.9 Soil and Pipe Displacements with Time ................................................................. 187

CHAPTER 8: SUMMARY, CONCLUSION AND FUTURE WORK .............................. 189

8.1 Engineering Significance and Applications ............................................................ 189

8.2 Summary of Results .............................................................................................. 189

8.2.1 Field Investigation ........................................................................................... 189 8.2.2 Load-deformation Analysis .............................................................................. 190 8.2.3 Pipe Response to Unsaturated Soil Conditions ............................................... 191 8.2.4 Soil-water Interaction in Highly Plastic Clays ................................................... 192 8.2.5 Soil-pipe-atmosphere Interaction under Field Conditions ................................ 194

8.3 Conclusion ............................................................................................................. 196

8.4 Future Work ........................................................................................................... 197

REFERENCES ............................................................................................................ 198

APPENDIX .................................................................................................................. 208

vii

LIST OF FIGURES

Figure ‎1.1: Research concepts ....................................................................................... 3

Figure ‎1.2: Conceptual presentation of key research elements ...................................... 4

Figure ‎1.3: Summary of the field investigation and numerical modeling programs .........14

Figure ‎2.1: A 100-year average precipitation for the City of Regina (Environment

Canada) ........................................................................................................................21

Figure ‎2.2: A 100-year average temperature for the City of Regina (Environment

Canada) ........................................................................................................................21

Figure ‎2.3: A general presentation of soil mechanics principles showing the role of the

surface flux boundary condition after (Ng and Menzies, 2007) ......................................24

Figure ‎2.4: The relationship between soil plasticity and swelling potential, after (Van Der

Merwe, 1964) ................................................................................................................29

Figure ‎2.5: A theoretical model for cracking mechanical mechanism after (Anderson,

2005) .............................................................................................................................40

Figure ‎2.6: Typical soil water characteristic curve of clay soil ........................................44

Figure ‎2.7: Measurement methods of soil-water characteristics ....................................48

Figure ‎2.8: The unsaturated soil stress state parameters using the combination of (σ –

ua), and matric suction (ua – uw) (Fredlund and Vanapalli, 2002)....................................62

Figure ‎2.9: Constitutive surfaces for an unsaturated, swelling soil .................................62

Figure ‎2.10: Principle stresses of a pipeline (Ng, 1994) .................................................69

Figure ‎2.11: Stresses in a pipeline under longitudinal extension (Ng, 1994) ..................69

Figure ‎2.12: Stresses in a pipeline under longitudinal bending conditions (Ng, 1994) ....69

Figure ‎2.13: Pipe deformation diagram (Ring Theory) ...................................................72

Figure ‎2.14: Pipe displacement due to axial bending .....................................................72

Figure ‎2.15: Common soil movement induced failure modes for pipe networks after

(Cassa, 2008) ................................................................................................................74

Figure ‎3.1: Site location and layout ................................................................................82

Figure ‎3.2: Summary of field instrumentation types .......................................................82

viii

Figure ‎3.3: Schematic of the installed sensors layout at the field site ............................85

Figure ‎3.4: Daily and cumulative precipitation at the field site ........................................87

Figure ‎3.5: Daily air temperature at the field site ............................................................87

Figure ‎3.6: Daily wind speed at the field site ..................................................................88

Figure ‎3.7: Daily net radiation at the field site ................................................................88

Figure ‎3.8: Daily relative humidity at the field site ..........................................................89

Figure ‎3.9: Grain size distribution for backfill materials (sand and mixed concrete) .......91

Figure ‎3.10: Water content and dry unit weight profiles with depth at the field site ........94

Figure ‎3.11: Grain size distribution of Regina clay and clay till ......................................95

Figure ‎3.12: Soil water characteristic curve (SWCC) for Regina Clay ............................99

Figure ‎3.13: Soil water characteristic curve (SWCC) for clay till.....................................99

Figure ‎3.14: Soil water characteristic curves (SWCCs) for the backfill materials .......... 100

Figure ‎3.15: Hydraulic conductivity functions for Regina clay and clay till .................... 100

Figure ‎3.16: Hydraulic conductivity functions for the backfill materials ......................... 101

Figure ‎3.17: Volumetric water content in the clay deposit at various levels .................. 103

Figure ‎3.18: Estimated soil suction in the clay deposit at various levels ....................... 103

Figure ‎3.19: Air and soil temperature observed at the field site ................................... 105

Figure ‎3.20: Pipe displacements and soil pressures at the field site ............................ 106

Figure ‎3.21: Pipe displacements at the field site .......................................................... 106

Figure ‎4.1: Summary of soil-pipe interaction model for parametric study analysis ....... 110

Figure ‎4.2: Model geometry and definition of the problem ........................................... 110

Figure ‎4.3: Generated mesh ........................................................................................ 111

Figure ‎4.4: Effect of the trench width on the maximum pipe deformations ................... 114

Figure ‎4.5: Effect of soil modulus of elasticity on the overall soil surface displacement116

Figure ‎4.6: Effect of soil cover height and loading conditions on pipe deformation ...... 116

ix

Figure ‎4.7: Effect of the backfill modulus of elasticity on pipe deformations ................. 119

Figure ‎4.8: Effect of the soil cover thickness and loading conditions on pipe deformations

.................................................................................................................................... 119

Figure ‎5.1: Schematic diagram of the effect of dry and wet soil conditions on the pipe

performance after (Rajeev et al., 2012) ....................................................................... 124

Figure ‎5.2: Summary of the modeling procedure for displacement analysis ................ 124

Figure ‎5.3: Geometry for the two-dimensional soil-pipeline model ............................... 125

Figure ‎5.4: Theoretical model for the analysis of a buried pipe under unsaturated soil

conditions .................................................................................................................... 125

Figure ‎5.5: Influence of the pipe burial depth on the pipe displacements ..................... 130

Figure ‎5.6: Maximum displacements versus normalized volumetric water content ...... 130

Figure ‎5.7: Pipe displacements due to the variation in soil moisture content under hinged

end restraints ............................................................................................................... 131

Figure ‎5.8: Pipe displacements due to the variation in the soil moisture content under

fixed end restraints ...................................................................................................... 131

Figure ‎5.9: Pipe displacements in case of both hinged and fixed end restraints for a low

elastic modulus magnitude (i.e. PVC pipe) .................................................................. 133

Figure ‎5.10: Pipe displacements in case of hinged and fixed end restraints for a high

elastic modulus magnitude (i.e. steel pipe) .................................................................. 133

Figure ‎6.1: A typical desiccated soil profile and idealized matric suction profile, after

(Lau, 1987) .................................................................................................................. 138

Figure ‎6.2: Estimated cracking depth at different ground water depths (linear elastic

theory) ......................................................................................................................... 140

Figure ‎6.3: Estimated cracking depth at a ground water depth of 9.5m (shear strength

approach) .................................................................................................................... 140

Figure ‎6.4: Estimated cracking depth at a ground water depth of 16 m (shear strength

approach) .................................................................................................................... 141

Figure ‎6.5: Bimodal SWCC and laboratory suction measurements .............................. 146

Figure ‎6.6: SWCC shapes at different values for the fitting parameter (a) ................... 146

Figure ‎6.7: SWCC shapes at different values for the fitting parameter (b) ................... 147

Figure ‎6.8: Normalized volumetric water content versus mean normal stress .............. 151

x

Figure ‎6.9: Parametric analysis for normalized volumetric water content versus mean

normal stress relationship ............................................................................................ 152

Figure ‎6.10: Degree of saturation versus volumetric water content .............................. 153

Figure ‎6.11: Parametric analysis for degree of saturation versus volumetric water

content relationship ..................................................................................................... 154

Figure ‎6.12: Volumetric water content constitutive surfaces ........................................ 155

Figure ‎6.13: Void ratio constitutive surfaces ................................................................ 155

Figure ‎6.14: Degree of saturation constitutive surfaces ............................................... 156

Figure ‎6.15: Soil column configuration for the parametric study ................................... 159

Figure ‎6.16: Predicted suction profiles for Bimodal SWCC versus Unimodal SWCC .. 161

Figure ‎6.17: Predicted volumetric water content profiles for Bimodal SWCC versus

Unimodal SWCC ......................................................................................................... 161

Figure ‎6.18: Hydraulic conductivity factor versus elapsed time .................................... 162

Figure ‎6.19: Hydraulic conductivity versus elapsed time .............................................. 162

Figure ‎6.20: Predicted suction versus elapsed time at cracking factor magnitudes ..... 164

Figure ‎6.21: Predicted volumetric water content versus elapsed time at different

cracking depths ........................................................................................................... 164

Figure ‎6.22: Predicted suction versus elapsed time at different depths....................... 166

Figure ‎6.23: Predicted volumetric water content versus elapsed time at different depths

.................................................................................................................................... 166

Figure ‎6.24: Predicted volumetric water content versus elapsed time at different net

surface flux magnitudes ............................................................................................... 167

Figure ‎6.25: Predicted suction versus elapsed time at different net surface flux

magnitudes .................................................................................................................. 167

Figure ‎7.1: Soil-pipe-atmosphere modeling processes ................................................ 171

Figure ‎7.2: Schematic diagram showing the field site conditions ................................. 171

Figure ‎7.3: The developed mesh for the modeling analysis ......................................... 172

Figure ‎7.4: Soil temperature versus time in the pipe trench ........................................ 180

Figure ‎7.5: Soil temperature versus time at a depth of 2.92 m in the pipe trench ........ 180

xi

Figure ‎7.6: VWC and suction profiles at various levels for the native clay ................... 184

Figure ‎7.7: VWC at a depth of 0.45 m versus time for the native clay ......................... 184

Figure ‎7.8: Volumetric water content change in the pipe trench .................................. 186

Figure ‎7.9: Volumetric water content versus time in the pipe trench .......................... 186

Figure ‎7.10: Pipe displacements versus time.............................................................. 188

Figure ‎7.11: Daily ground displacements (native clay) ................................................ 188

xii

LIST OF TABLES

Table ‎1.1: General description of the FlexPDE script sections (Liu, 2005; PDE Solutions

Inc., 2014) .....................................................................................................................12

Table ‎2.1: Ranges of air temperature, precipitation, rainfall deficit, and freezing index in

Regina area from 1980 to 2004 (Hu and Hubble, 2007) ................................................20

Table ‎2.2: Degree of expansion as estimated from classification test data, after (Holtz

and Kovacs, 1981).........................................................................................................27

Table ‎2.3: Typical values of geotechnical index properties for the different clay minerals

after (Das, 2006; Holtz and Kovacs, 1981; Mitchell, 1993; Yong and Warkentin, 1966; Zhang, 2004) .................................................................................................................27

Table ‎2.4: Regina clay classification and mineralogical tests results (Fredlund, 1967;

Fredlund, 1975; Hu and Vu, 2011) .................................................................................28

Table ‎2.5: Evaluation criterion for volumetric soil water monitoring techniques (Muñoz-

Carpena et al., 2004) .....................................................................................................49

Table ‎2.6: Evaluation criterion for soil suction monitoring techniques (Muñoz-Carpena et

al., 2004) .......................................................................................................................50

Table ‎3.1: Dimensions and properties of the existing AC pipe section ...........................83

Table ‎3.2: Dimensions and properties of the Instrumented PVC pipe ............................83

Table ‎3.3: Summary of the main details of the instruments installed at the field site ......85

Table ‎3.4: Geotechnical index properties of Regina clay ..............................................96

Table ‎3.5: Geotechnical index properties of glacial clay till ...........................................97

Table ‎4.1: Summary of the main input parameters ...................................................... 112

Table ‎5.1: Initial material parameters of the PVC pipe ................................................. 126

Table ‎7.1: Geotechnical index properties for the clay, mixed concrete and sand ......... 173

Table ‎7.2: Summary of the main parameters of the stress-strain modeling analysis .... 173

Table ‎7.3: Summary of the modeling results for the pipe trench for the period from April

2006 to April 2007 ....................................................................................................... 181

Table ‎7.4: VWC change of the native clay for the period from 25th April 2006 to 25th April

2007 ............................................................................................................................ 185

1

CHAPTER 1: INTRODUCTION

1.1 Definitions

Figure 1.1 summarizes the main concepts related to this research. The aim was to

develop an advanced framework rooted in the understanding of the following main

technical areas: unsaturated soil mechanics, seasonal weather changes, soil expansion

and shrinkage criterion, soil-water interaction, and soil-structure interaction. The

development of a soil-structure-atmosphere interaction model requires the use of

mathematical principles drawn from soil mechanics, hydraulics and geophysics and

applies these principles to the main engineering problems including flow, stress, and

deformation phenomena. By definition, expansive refers to the tendency to spread out.

Expansive soils also known as swell-shrink soils are subjected to changes in volume in

response to any changes in moisture content. Highly plastic clays are a good example

of this type of soils.

Quantitative assessment of the net flux at the soil-atmosphere interface requires

the knowledge of the relevant soil-water interaction properties and the applied

environmental conditions. Seasonal weather changes influence soil moisture storage

characteristics due to variations in precipitation, temperature, rainfall, actual evaporation,

runoff, wind, and groundwater. In order to accurately model soil behavior, the

mechanical stress, pore water pressure, pore air pressure, and temperature have to be

used as stress state variables. The evapotranspiration process at the soil-atmosphere

boundary must be evaluated first. Then, the soil performance can be simulated by a

hydro-mechanical stress analysis. Daily weather data availability is a challenge

constraining not only the features of the modeling process but also its resulting

predictive ability and accuracy. Detailed information including site location, elevation and

daily weather data, such as solar radiation, relative humidity, air temperature and wind

2

speed, are all required. These weather data could normally be obtained from local

weather stations.

The theoretical basis and partial differential equations of ground–atmosphere and

soil-structure interactions have been developed reasonably well (Rajeev et al., 2012). A

key property that is vital for the implementation of unsaturated soil principles is the soil

water characteristic curve (SWCC). SWCC is used in research for the determination of

unsaturated soil-atmosphere interaction (Wilson and Fredlund, 2000) and unsaturated

soil-structure interaction (Zhang, 2004). The main objective of soil-atmosphere transport

models is to determine and evaluate the soil exchange fluxes with the atmosphere. Soil-

atmosphere interaction models involve the estimation of the rates of heat and moisture

flow through the soil structure. The boundary conditions depend on two primary

components, evaporation (or evapotranspiration) and precipitation. Evaporation is

related to free water surfaces and non-vegetated soil surfaces while evapotranspiration

is associated with wet surfaces or plant leaves (Veihmeyer, 1964). The understanding of

the evapotranspiration process has been established since the 1940’s (Gitirana Jr,

2005).

Soil behavior is generally described by constitutive relationships. The constitutive

relationships provide a framework for understanding the soil behavior under different

loading conditions, and can be formatted in finite element and finite difference codes for

use in numerical analyses. Modeling of the volume change of unsaturated soils consists

of stress-deformation analysis under mainly water and heat flow processes. There are

a number of volume-mass constitutive models that have been established for

unsaturated soils (Pham, 2005). The key constitutive relationships for unsaturated soils

were developed as an extension of the saturated constitutive relationships, and

incorporated the use of two independent soil properties, the total normal stress and soil

suction (Hung, 2003). Figure 1.2 summarizes the key research elements and concepts.

3

Figure 1.1: Research concepts

Thesis Technical Aspects

Unsaturated Soil-structure

Interaction

Seasonal Weather Conditions

Unsaturated Soil Mechanics

Problematic Soils

Unsaturated Soil-water Interaction

4

Figure 1.2: Conceptual presentation of key research elements

Key Research Elements

Unsaturated Soil Properties

SWCC

Hydraulic Conductivity

Function

Volume-mass Constitutive

Relationships

Soil-atmosphere Interaction

Soil Moisture Profile

Soil Temprature Profile

Soil-structure Interaction

Displacement Magnitude

SWCC: Soil Water Characteristic Curve

5

1.2 Problem Recognition

The motivation for incorporating the principles of unsaturated soil mechanics to

various geo-environmental engineering problems is increasing (Ng and Menzies, 2007).

The severity of swell-shrink problems is influenced by the current environment, future

changes in loading and environmental conditions (Fredlund, 1975). Factors affecting the

soil swell-shrink potential include stress history, degree of saturation, mineralogical

composition, plasticity characteristics, and loading conditions. The study of the behavior

of expansive soils has been the driving force of unsaturated soil research (Ng and

Menzies, 2007). Field behavior of this type of soil plays a vital role in the performance of

engineered infrastructure, specially, when the foundation soil is exposed to repetitive

variations in moisture content.

The long-term performance of aging infrastructures, such as buried pipelines, is a

great challenge facing municipal engineers. The traditional design of underground

pipelines has typically been to consider the limiting conditions represented by either

completely saturated or entirely dry soil state (Nyman, 1984). The deformations of

underground pipes are in general affected by the pipe itself and the surrounding soils;

and controlled by design factors and construction techniques. The level of soil load on

buried pipes depends on the nature of the soil, its natural degree of saturation, and

seasonal meteorological variations. The core design criteria of any pipeline system are

to provide adequate serviceability, minimize any significant damages, and establish a

structural adequacy for the intended service life. The failure of underground pipelines is

a product of the interaction of three elements, the environment, soil and pipe.

Weather-displacement models for buried pipelines are still not well established,

and therefore, there is an urgent need first to develop these models, and then support

them with field monitoring data. Developing a framework to simulate real case studies

and incorporate more complexities of the boundary conditions as well as advanced

6

unsaturated soil parameters is important for the design and construction of different

infrastructures. The understanding of the response of the soil and pipe to surrounding

environmental and unsaturated soil conditions is also useful for modifying pipe design

and maintenance methods. In addition, predicting soil water content, soil suction, and

temperature profiles is a key element to study the soil-structure-atmosphere interaction.

1.3 Engineering Significance

Damages of under and above ground infrastructures have been well reported in

areas characterized by expansive soils and located in arid and semi-arid climate zones

of the world costing billions of dollars every year (Day, 1999). The impact of these

damages on regional or national scale is exceedingly noticeable. The surficial lacustrine

clay deposit in the City of Regina, in southern Saskatchewan, was characterized as

unsaturated highly plastic clay. This clay deposit experiences high volume changes due

to the variation in its moisture content. Underground pipelines buried in the city would

then incorporate hazard of abnormal displacements due to soil movements. The field

behavior of soil deposits is dependent on the local environmental conditions and

seasonal weather variations. Unsaturated soils may also experience wetting and drying

cycles due to park watering, or any substantial water leakage. These changes produce

variations in soil moisture content, which in turn, result in extensive soil movement

subsequent to installation.

Pipes buried in unsaturated soil deposits may be subject to severe stresses or

even failure as a result of soil movement. Failure of buried is a common problem for

small diameter underground pipelines. Various researchers reported that the swell-

shrink behavior of clay soils is a contributing factor to the failures of shallow buried

pipelines. It was also reported that the majority of small diameter water mains fail in the

circular (circumferential) mode (Rajani et al., 1996). A circumferential break is typically a

7

result of excessive bending stress as a result of differential soil movements. Previous

studies reported that seasonal climate changes are a contributing factor to the failures of

shallow buried pipelines (Clark, 1971; Gould, 2011; Hu et al., 2010; Hudak et al., 1998;

Morris, 1967; Rajeev et al., 2012).

In the City of Regina, It was found that circumferential breaks for the water system

were the predominant failure mode (approximately, 91.45%) among the total number of

breaks (2288) from 1980 to 2004 (Hu and Hubble, 2007). Most of the pipe ruptures

occurred in the 150 mm diameter pipes (approximately, 80.8%), and more than (94%) of

the breaks occurred in the 150 and 200 mm diameter pipes (Hu and Hubble, 2007). The

abnormal annual breakage incidents due to excessive bending stresses and

deformations pose a great demand for advanced research of pipeline modeling under

local conditions. Despite the known influence of unsaturated soils on the performance of

water mains, little work has been completed to model the interaction and quantify clear

relationships for the practice of pipeline engineering.

A review of the historical failure records of pipeline infrastructure worldwide and

in the City of Regina can be found later in this thesis. Generally, a total number of 850

water main breaks occur daily in North America, costing over $ 3 billion for annual

repairs (Uni-Bell PVC Pipe Association, 1991). In the City of Regina, a significant

amount of damage of buried infrastructure (municipal water and wastewater pipelines)

were reported annually (Hu and Hubble, 2007). The effect of the breakage of

underground pipeline networks in most cases causes serious problems to pavements,

buildings, and other surrounding structures. Therefore, proper design and installation of

such systems is directly linked to the sustainable management of water resources and

the environment, and would result in an enhanced standard of living.

8

1.4 Research Objectives

In order to enhance the integrity of hydro-mechanical modeling of soil-structure-

atmosphere interaction, and evaluate the performance of underground buried structures

under the seasonal variations in climate conditions, the core objectives of this research

were to:

Review the theoretical basis and the governing partial differential equations of

ground–atmosphere interaction and soil-structure interaction. Although, there are

several available approaches to estimate the volume-mass constitutive models of

unsaturated soils, however these approaches were still not utilized for studying

soil-pipe interaction problems.

Investigate the unsaturated properties of native soil deposits in the study area.

Analyze the results of a field instrumentation program to depict field conditions,

and employ the analyses of the monitoring data of the first three years after

installation in order to develop theoretical and engineering bases for the structural

response of pipelines under field conditions.

Conduct a sensitivity analysis of a hypothetical buried pipeline under different soil

and loading conditions. The analysis included load-deformation and partial

saturation analysis of the surrounding soil. Develop comparisons between

modeling and analytical approaches. Based on these comparisons, knowledge and

recommendations can be obtained and transferred to model the pipelines under

field conditions.

Develop a mathematical framework for soil-water interaction problems based on a

bimodal SWCC, representative consolidation test results, and a model of cracking

mechanism of the top clay layer. In addition, use this theoretical framework to

9

simulate transient water flow in the soil structure, and determine the changes in the

soil suction and VWC with time.

Determine the net flux for the field simulation using daily weather data which

includes rainfall, solar radiation, air temperature, relative humidity and wind speed.

Then, utilize the developed mathematical approach to model the soil-water

interaction and predict the resulting soil and pipe displacement profile.

1.5 Finite Element Approach

The complexity of the soil-pipe-atmosphere numerical analysis is derived from the

irregularity of boundary conditions, geometries and material properties. The finite

element modelling (FEM) is typically an effective technique for solving various partial

differential equations. The FEM has been widely utilized to solve different engineering

problems. The FEM solves the given partial differential equations over a finite element.

The model elements are typically connected to each other, and the field of elements is

analyzed by yielding the solution from one element to another (Saadawi and Wainer,

2003). Liu (2005) reported that the application of the FEM may consist of the following

main features: (i) selecting the element configuration; (ii) selecting the approximation

function; (iii) defining the governing constitutive relationships; (iv) obtaining the element

relationships; (v) developing global equations and defining boundary conditions; (vi)

solving the main unknowns; (vii) solving the secondary measures; and (viii)

comprehension of results.

In this research, the finite element modelling analysis was implemented using a

commercial Finite Element program named FlexPDE (PDE Solutions Inc., 2014).

FlexPDE is a general partial differential equation solver that utilizes a scripted finite

element model builder for providing numerical solutions of boundary value problems

(PDE Solutions Inc., 2014). The FlexPDE's model script has to be written and formulated

10

by the user, and then, the operations can be performed by FlexPDE to transform the

description form of the partial differential equations into a finite element model, solve the

problem, and produce graphical output of the results (Liu, 2005). Table 1.1 illustrates

general descriptions of the main sections of FlexPDE scripts. The user can employ

FlexPDE scripting language to identify the mathematics of the governing partial

differential equations and the problem geometry (PDE Solutions Inc., 2014). Therefore,

there is no uncertainty concerning the process of solving equations, compared with

fixed-application program applications (PDE Solutions Inc., 2014). Variables, equations

or terms can be effortlessly introduced (PDE Solutions Inc., 2014).

For the purposes of this research, FlexPDE software package was found to

provide the following main capabilities (PDE Solutions Inc., 2014): (i) a finite element

module that chooses a suitable solution format for steady-state or time-dependent

problems (ii) a technique that solves non-linear partial differential equations of second

order or less through separate measures for linear and nonlinear problems, (iii) flexibility

to put in nonlinear functions for material properties (e.g. unsaturated soil properties)

(Pentland et al., 2001); (iv) a script editing module that presents a full text editing tool

and a graphic illustration procedure, (v) an equation analyzer that propagates defined

parameters and relationships, (vi) a mesh generation module that typically builds a

triangular finite element mesh over problem domains, (vii) an error estimation method

that computes the capability of the mesh and apply refinements to the mesh wherever

the error exceeds a user-defined error tolerance (Liu, 2005), (viii) a graphical output

module that accommodates algebraic functions and produces contour result, (ix) a data

export module that generates reports in different forms (Liu, 2005).

The FlexPDE's script was written for different models established throughout this

research work based on a mathematical structure. The mesh generating system

associated with FlexPDE automatically created a finite element mesh fitting the problem

11

domain. Cell sizes were typically controlled by the spacing between explicit points in the

domain boundary. The developed initial mesh consisted of triangular finite elements. Cell

density in the initial mesh was managed by the mesh spacing and density parameters

entered in the program. These parameters defined the maximum cell dimension and the

minimum number of cells per unit distance. A consistency check was then applied to the

integrals of the partial differential equations over the mesh cells. The relative uncertainty

of the solution was predicted by the software and compared with the defined accuracy

tolerance of 0.01%. When any mesh cell exceeded the tolerance, the cell was then

automatically refined, and the solution was re-computed until a defined error tolerance of

0.01% was achieved for every cell of the mesh.

12

Table 1.1: General description of the FlexPDE script sections (Liu, 2005; PDE Solutions Inc., 2014)

Section Duties

TITLE Includes a descriptive expression for the modeling output

SELECT Includes the default parameters of FlexPDE

VARIABLES Specifies the dependent variables

DEFINITIONS Defines parameters and relationships

EQUATIONS Defines the governing partial differential equations

INITIAL VALUES Defines the initial values or conditions for nonlinear or time-dependent problems

BOUNDARIES Includes a description of the geometry

PLOTS Lists the required graphical outputs Plots include contour, surface, elevation or vector.

13

1.6 Contributions

The research objectives were accomplished and reported through distinct stages,

include the following:

Detailed evaluation study was first demonstrated to identify the most significant

pipe design factors under load-deformation conditions (Saadeldin and Henni,

2013; Saadeldin et al., 2013a; Saadeldin et al., 2015). The numerical analysis

was performed for various backfill materials and native soils around the pipes, as

well as multiple loading magnitudes.

A plane soil-pipe interaction model was developed to incorporate the effects of

variation in soil moisture on the deformations of buried pipelines (Saadeldin et

al., 2015; Saadeldin et al., 2013b). Soil suction profiles were estimated based on

unsaturated soil characteristics and field test results, and utilized as an input data

for the model.

Mathematical formulation of the soil-water interaction of highly plastic clays was

established (Saadeldin and Henni, 2016). Advanced unsaturated soil parameters

including bimodal SWCC, hydraulic conductivity function, and volume-mass

relationships were developed for the native soils, and then used to model

transient water flow through a soil column.

An advanced climate-ground-pipe interaction model was finally developed to

simulate the field behavior of buried pipes (Saadeldin et al., 2016). The model

incorporated the variation in climate conditions with time. The modeling approach

was validated using field monitoring data.

The diagram in Figure 1.3 illustrates a summary of the main research details and

components.

14

Figure 1.3: Summary of the field investigation and numerical modeling programs

Pipe Installation &

Instrumentation Details

Climate Data &

Unsaturated Soil

Characteristics

Field Measurements

Analysis

Theoretical Evaluation

(Pipe Deformation)

Variations in Loading

and Soil Conditions

Load-deformation-soil-

pipe Analysis

Theoretical Evaluation

(Swelling Magnitude)

Field Soil Moisture &

Suction Variations

Unsaturated Soil Stress-

strain Analysis

Modified Unsaturated

Soil Characteristics

Modified Boundary

Conditions

Heat Flow Analysis

Seepage Analysis

Unsaturated Soil-

stress-strain Analysis

Evaluation of Soil-pipe-atmosphere Interaction under Field Conditions

Evaluation of Research Modifications

Comparisons between Field & Numerical Modeling Results

Field Investigation and Numerical Modeling Programs

Load-deformation Analysis

Atmosphere-soil-pipe Interaction Analysis

Unsaturated Soil Displacement Sensitivity Analysis

Field Investigation

Field Investigation-Numerical Modeling Integration

15

1.7 Outline of this Research

Chapters in this thesis were ordered in accordance with the research objectives. The

main contents of the chapters are as follows:

Chapter one presents a brief introduction of the soil-structure-atmosphere

interaction elements, as well as the problem definition, research objectives and

methodology.

Chapter two reports a literature review on expansive clays, unsaturated soil

parameters, measurements of soil moisture-suction characteristics, soil cracking

mechanisms, unsaturated soil-atmosphere interaction, unsaturated soil-structure

interaction, pipeline infrastructure design criteria and historical failure data, and

finally related previous numerical studies.

Chapter three gives the details of the field investigation program including

material properties, laboratory tests, instrumentations, daily weather conditions,

and soil profiles.

Chapter four provides the results and discussion of load-deformation analysis.

The results of the analysis were compared with analytical solutions obtained using

empirical design equations.

Chapter five provides detailed soil and pipe responses to partial saturation of

surrounding highly plastic clay. The model was based on the field investigation

details and aimed to study the sensitivity of the soil and pipe displacement

magnitudes to unsaturated soil properties. The modeling results were compared

with the results of the empirical equations using laboratory testing results.

Chapter six presents the mathematical formulation of the soil-water interaction

including the development of a bimodal SWCC equation, hydraulic conductivity

function, and the depth of cracking. The mathematical framework was utilized to

16

model the time dependent behavior of a soil column under applied surface water

flux.

Chapter seven presents the modeling of the pipe in the field and reports the

changes in soil and pipe conditions as a response to daily weather conditions. The

modeling results were used to draw conditions for the influence of boundary

conditions on the displacements of underground pipes under field conditions.

Chapter eight contains the conclusions obtained from this research and presents

few research recommendations.

17

CHAPTER 2: LITERATURE REVIEW

2.1 Surficial Geology

The geologic time scale devides the 4.6 billion years of the earth history to a

hierarchy of time periods corresponding to the earth formation history (Natural

Resources Canada, 2010). The Precambrian era began with the development of the

Earth and followed by the Paleozoic, Mesozoic, and Cenozoic eras. Each of these eras

is subdivided into periods, the periods into epochs, and epochs into ages (Natural

Resources Canada, 2010). Geologically, Canada is one of the oldest countries

worldwide, and the Precambrian rocks extend over more than half the land area of

Canada (Wallace, 1948). Three major geological events governed the geological

formation of Canada, namely the shield formation, mountains formation from sediments

accumulated in basins in the region of the margins of the Shield, and the sediments'

deposition within the intervening areas (Stearn et al., 1979). The province of

Saskatchewan is underlain by crystalline Precambrian rocks of the stable North

American Craton (Saskatchewan Geological Society, 2003). According to

Saskatchewan Geological Society (2003), the Precambrian rocks are exposed as parts

of the Canadian Shield in the northern Saskatchewan and its southern part is covered

by un-metamorphosed Phanerozoic sedimentary rocks in two-thirds of the province.

Sediment traps were also developed during the late Cretaceous sedimentation

processes in southern Saskatchewan (Pruett and Murray, 1991).

Various geological processes (i.e. tectonic events, erosion, physical and

mechanical weathering, and glaciation) influenced the behavior of the soil sediments in

the province of Saskatchewan. The soils were glaciated resulting in a considerable

thickness of a glacial drift found between the bedrock and the existing ground surface.

The glacial drift consists of different soil sediments that were transported by the

18

movement of large bodies of glacial ice. The glacial drift also experienced different

chemical and physical weathering processes that took place as a result of frost actions.

The weathering processes included: (i) the contraction of ice-rich frozen soil, (ii)

segregated ice formation, (iii) internal pressures due to the expansion of water upon

freezing (Trenhaile, 2004). The surficial geology of the City of Regina area experienced,

in the past, several glacial ice advances, retreats, and meltdowns. Stratified drift

deposits, more than 200 m in depth, cover the Bearpaw shale below the area (Mollard et

al., 1998). The topography in the Regina area is in general flat to gently undulating

(Dobchuk et al., 2009).The soil profile consists mainly of lacustrine soil sediments

deposited about 14,000 years ago (Christiansen, 1979), specifically, during the last

glaciation phase advance (Christiansen and Sauer, 2002).

The lacustrine drift deposit within Glacial Lake Regina consists mainly of highly

plastic clay, known as Regina clay. The lacustrine drift deposit occasionally encounters

a lower section that is very silty, low plastic, saturated and less stiff than the upper clay.

The lacustrine clays and silts extend to a depth of approximately 12 m (Fredlund, 1975).

A glacial clay till deposit lies underneath the upper lacustrine deposit. Previous studies

identified two different layers of the glacial clay till deposits (Christiansen and Schmid,

2005). The upper clay till was characterized to be relatively thin, weathered (brown)

glacial clay till of the Battleford Formation. However, the lower layer was identified to be

un-oxidized (grey) glacial clay till of the Floral Formation (Christiansen and Sauer, 2002;

Christiansen, 1979). In some areas, the two till layers are directly connected, and in

other areas separated by a layer comprised of sand and gravel or sandy, silty clay with

variable thickness (Christiansen and Sauer, 2002; Christiansen, 1979). Both the glacial

clay till and the lacustrine deposits are a part of a distinct geological unit known as the

Saskatoon Group (Christiansen and Sauer, 2002; Ellis et al., 1965).

19

2.2 Climate Conditions

The climate is a contributing factor to the seasonal variations in moisture content

and temperature of surficial soil deposits. Soils also vary in nature depending on the

contribution of climate conditions. The climate in the City of Regina is identified as a

semi-arid, continental climate characterized by warm summers and cold, dry winters

(Dobchuk et al., 2009). The daily weather conditions for the area were obtained from a

nearby Environment Canada meteorological station (Regina International Airport,

Regina, Saskatchewan) during a 100 year period from 1909 to 2008. Table 2.1 shows

the variation in air temperature, precipitation, rainfall deficit, and freezing index values

during a 25 year period from 1980 to 2004 (Hu and Hubble, 2007).

The total precipitation normally accounts for rainfall and snow throughout the

year. The historical temperature data for the City of Regina area demonstrates that the

mean monthly temperature is below 0 °C for five months (i.e., from November to March)

(Hu and Hubble, 2007). Figures 2.1 and 2.2 show the mean monthly temperature and

precipitation within the City of Regina, during a 100-year period, from 1909 to 2008,

obtained from Environment Canada's database. The difference in mean temperature

between summer and winter months is around 35.7 °C. Likewise, the difference in mean

precipitation is approximately 63.8 mm.

20

Table 2.1: Ranges of air temperature, precipitation, rainfall deficit, and freezing index in Regina area from 1980 to 2004 (Hu and Hubble, 2007)

Measurement (unit) Variation range

Average Minimum Maximum

Average annual air temperature (oC) 3.0 0.1 4.2

Total precipitation (mm) 396 260 593

Annual rainfall deficit (mm) 172 -16 320

Freezing index (oCˑday) 1552 1071 2172

21

Figure 2.1: A 100-year average precipitation for the City of Regina (Environment Canada)

Figure 2.2: A 100-year average temperature for the City of Regina (Environment Canada)

Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov DecMonth

0

20

40

60

80

Pre

cip

itation (

mm

)

Average Precipitation (100-Year Average 1909-2008)

63.8 mm

Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov DecMonth

-20

-10

0

10

20

Tem

pra

ture

(oC

)

Average Temperature (100-Year Average 1909-2008)

35.7 oC

22

2.3 Expansive Soils

Expansive soil deposits are identified in arid and semi-arid areas of the world,

such as throughout Canada, Australia, United States, and many other countries, where

the annual evapotranspiration is equal or more than the precipitation (Chen and Ma,

1987). Expansive clays swell and shrink when subjected to any wetting and drying

processes. The field behavior of these soil deposits plays a vital role in the performance

of engineered infrastructure, with more emphasis, when the foundation soil is exposed

to significant variations in moisture content. The interaction of climate, geology and

hydrological conditions governs the formation and behavior of clays. Fredlund (1975)

and Lu (1969) identified the presence of swelling soils in the City of Regina based on a

series of site investigations, and concluding that Regina clay was found to be

desiccated and active.

The severity of the swell-shrink problems is typically influenced by the past soil

history, the current environment and future changes in loading and environmental

conditions (Fredlund, 1975). In addition, swelling and shrinkage processes were also

found to be not perfectly reversible (Holtz and Kovacs, 1981). When expansive clays

get saturated, water molecules enter between the clay sheets, and in return, the total

soil volume increases significantly (Jones and Jefferson, 2012). For most expansive

clays, volume expansions of 10% are very common (Nelson and Miller, 1997). The

water molecules between the clay sheets also exfiltrate by water evaporation/drying

processes causing the overall volume of the soil to decrease (Jones and Jefferson,

2012). The development of cracks is always associated with the water

evaporation/drying processes, which in return, facilitate the access of water.

The field (natural) degree of saturation is a fundamental factor affecting the

swelling magnitude. The amount of swelling-shrinkage induced soil displacements is

controlled by the initial moisture condition of the near-surface zone, the zone of

23

seasonal fluctuations, or the active zone, which extends to a depth of 3 m below the

ground level or more when tree roots exist (Driscoll, 1983). According to Nelson et al.

(2001), the active soil zone may be defined as follows: (i) the area of soil that

significantly contributes to soil expansion at any time, (ii) the depth of seasonal

fluctuation in soil moisture content caused by climatic changes near the ground surface,

(iii) the depth of wetting in which soil moisture contents have changed as a result of the

water supply from external sources, and (iv) the depth of potential heave in which the

overburden vertical soil stress equals or exceeds the soil swelling pressure. The active

soil zone is of particular importance to estimate the total soil heave by integrating the

displacement produced over the zone in which the soil moisture content was changed

(Jones and Jefferson, 2012; Walsh et al., 2009).

Negative pore water pressures are typically developed in the active soil zone due

to water infiltration/exfiltration processes as presented in Figure 2.3 (Nelson and Miller,

1997). The extent of the active zone is influenced by several factors, mainly the air

temperature, soil type, site topography, and the depth of ground water table (Chao et al.,

2015). The depth of the active zone can be verified by collecting soil suction

measurements in the field (Chen, 1988). The development of the ground water table is

mainly controlled by the net surface flux (the difference between the downward flux (i.e.

precipitation) and the upward flux (i.e. evapotranspiration) (Ng and Chen, 2008). For arid

or semi-arid regions, the groundwater level is typically lowered with time (Ng and Chen,

2008). However, it most likely remains relatively close to the ground surface under

temperate or humid conditions (Ng and Menzies, 2007). The hydrostatic pore water

pressure condition exists when the net flux applied on the ground surface equals zero

(Ng and Menzies, 2007).

24

Figure 2.3: A general presentation of soil mechanics principles showing the role of the

surface flux boundary condition after (Ng and Menzies, 2007)

25

2.4 Clay Mineralogy

Clay mineralogy is a primary factor controlling the swelling /shrinkage behavior of

clays. Internal surface area, charge distribution, and cation species determine the

swell/shrink potential of a clay mineral (Thomas, 1998). The most common clay minerals

consist of Kaolinite, Montmorillonite, and Illite. Clays composed of Montmorillonite

minerals are characterized by high swelling potential, high Liquid Limit, high Plasticity

Index, and less particle size compared to other clay minerals (Illite and Kaolinite). The

changes in the clay mineral composition are the primary factors for the swelling to occur

(Mitchell, 1993). The swelling action is controlled by the balancing of forces of

interaction between water, clay surface, and ions. The clay particles typically hold a net

negative charge and the larger negative charges are originated from larger specific

surfaces (Das, 2006).

Clay minerals adsorb external cations and maintain them in an exchangeable

state (Ma and Eggleton, 1999). In addition, adsorbed cations may also be replaced by

other cations which is known as the cation exchange process. The cation exchange

capacity (CEC) of soils is defined as the total number of adsorbed exchangeable cations

per 100 gms of dry soil (milliequivalent/100 gms) (Robertson et al., 1999). Swell

potential typically increases as CEC increases (Manosuthikij, 2008). Norrish (1954)

found that a montmorillonite could experience a swelling magnitude of 100% in volume

when it gets saturated with a divalent cation, such as Ca2+. The swelling potential is

influenced by the soil index properties, such as plasticity index, shrinkage limit and clay

content. Various empirical methods were established to correlate the swelling potential

to soil index properties such as Table 2.2 and Figure 2.4.

The colloid size for soils is defined as the fraction finer than 0.002 mm determined

by hydrometer analysis. Clay particles are mostly in the colloidal size range of 1mm,

where 2 mm is considered to be the upper limit (Das, 2006). The magnitude of swell

26

increases with the increase of colloid content. The plasticity index of clays increases

linearly with the percentage of clay-size fraction (% finer than 2 mm by weight)

(Skempton, 1953). Typical values of the liquid and plastic limits and specific gravity for

some clay minerals are summarized in Table 2.3. The activity of clays is defined as the

slope of the line correlating plasticity index and percentage finer than 2 µm (Skempton,

1953). Skempton (1953) identified three classes of the clay activity based on the

magnitude of the soil activity: (i) inactive (soil activity is less than 0.75); (ii) normal

(soil activity is between 0.75 and 1.25); and (iii) active (soil activity is greater than

1.25). Typical values of surface charge density, activity, and CEC for some clay minerals

are also summarized in Table 2.3.

Surficial lacustrine soil deposits in Saskatchewan were characterized by physically

powerful montmorillonitic pattern (Rice et al., 1959). Regina clay classification and

previous mineralogical tests results are illustrated in Table 2.4. The test results of

Regina clay resulted in the following mineralogical composition, approximately 53% of

montimorillonite, 35% of Illite and 12% of Kaolinite (Fredlund, 1975). Regina clay is

generally identified as montmorillonitic clay due to the high Montmorillonite content

which controls its behavior. In addition, previous test results confirmed that Regina clay

is in particular a calcium Montmorillonitic clay (Fredlund, 1975). The geotechnical

classification and mineralogical test records also confirmed the high swelling and

shrinkage potential of Regina clay. This high shrink-swell potential is resulting from the

high clay content (approximately more than 60%), and the dominance of Smectite or

Montmorillonite minerals (Anderson, 2010).

27

Table 2.2: Degree of expansion as estimated from classification test data, after (Holtz and Kovacs, 1981)

Degree of Expansion ( % - 1 µm) Plasticity

Index Shrinkage

Limit

Very High 28 > 35 < 11

High 20 - 31 25 - 41 7 - 12

Medium 13 - 23 15 - 28 10 - 16

Low 15 < 18 > 15

Table 2.3: Typical values of geotechnical index properties for the different clay minerals after (Das, 2006; Holtz and Kovacs, 1981; Mitchell, 1993; Yong and Warkentin, 1966; Zhang, 2004)

Property

Clay Minerals

Kaolinite Illite Montmorillonite

Specific Gravity, Gs 2.6 2.8 2.65 - 2.80

Plastic Limit, LL (%) 20 - 40 35 - 60 50 - 100

Liquid Limit, PL (%) 35 - 100 60 - 120 100 - 900

Cation Exchange Capacity, CEC (me/100gm)

3 - 15 10 - 40 80 - 150

Activity 0.3 - 0.5 0.5 - 1.2 1.5 - 7.0

Length and Width (mm) 0.3 - 3.0 0.1 - 2.0 0.1 - 1.0

Specific Area (m2/g) 20 - 80 80 - 100 800

Reciprocal of average surface charge density (Å2/electronic charge)

25 - 100

Swell-shrink Potential Low Medium High

28

Table 2.4: Regina clay classification and mineralogical tests results (Fredlund, 1967; Fredlund, 1975; Hu and Vu, 2011)

Property Value

Specific Gravity 2.7 – 2.8 A

tte

rberg

Lim

its

Shrinkage limit, SL (%) 13.1

Plastic limit, PL (%) 24.9 – 34

Liquid limit, LL (%) 50.6 – 94

Plasticity index, PI (%) 25.7 – 60

Gra

in S

ize

Dis

trib

utio

n Gravel sizes (≤62 mm) (%) 0 – 0.4

Sand sizes (≤5 mm) (%) 0.1 – 8

Silt sizes (%) 40.6 – 41

Clay sizes (%) 51 – 58

Cla

y

Min

era

ls

(fo

r le

ss 2

mic

ron) Kaolinite (%) 8

Illite (%) 15

Montmorillonite (%) 77

Exch

an

ge

Ca

tion

s

Magnesium (me/100 gm) 15.3

Calcium (me/100 gm) 54.4

Potassium (me/100 gm) 0.59

Sodium (me/100 gm) 1.77

CEC (me/100 gm) 31.7

Activity 7.0

29

Figure 2.4: The relationship between soil plasticity and swelling potential, after (Van Der Merwe, 1964)

0 10 20 30 40 50 60 70

Clay Fraction of Whole Sample (%2)

0

10

20

30

40

50

60

70

Pla

sticity Ind

ex (

%)

Very High

High

Medium

30

2.5 Unsaturated Soil Parameters

2.5.1 Stress State Variables

Several equations and definitions were developed to determine the effective

stresses for unsaturated soils. Previous research studies indicated that there are

significant differences between the field behavior of saturated and unsaturated soil

deposits due to the existence of matric suction (Wulfsohn et al., 1996). Therefore, the

effective stress theory of saturated soils cannot be applied directly to unsaturated soils.

Unsaturated soils were represented by three-independent phases, namely air, water and

solid (Lambe and Whitman, 2008). The contractile skin (air-water interphase) can be

considered as an independent fourth phase, which develops normal stresses due to the

pore-water pressure (Davies and Rideal, 1961). However, the contractile skin phase is

usually considered in the water phase when establishing volume-mass relationships for

unsaturated soils (Fredlund and Rahardjo, 1993b).

The stress state influences the mechanical behavior of soils. Terzaghi (1936)

established the effective stress concept for saturated soils as indicated in Equation (2.1).

(2.1)

where;

σ’ is the effective normal stress,

σ is the total normal stress, and

uw is the pore-water pressure.

Skempton (1961) modified the effective stress equation of Terzaghi to count for

the effect of soil compressibility as shown in Equation (2.2). The effective stress

principle has then been extensively used in engineering modelling and analysis to

correlate between elastic deformations and stresses of a soil structure.

31

(2.2)

where;

cg refers to the compressibility characteristic of the solid grains, and

cs refers to the compressibility characteristic of the solid skeleton.

The effective stress equation of Terzaghi did not account for some factors,

including the capillary forces, particle and pore size, pore air and pore-water pressure or

matric suction, air–water surface tension, degree of saturation and contact angle

(Fredlund and Rahardjo, 1993b). The effective stress equations for unsaturated soils

were derived from the definition of the normal inter-particle forces. The principle of

effective stress, applied in saturated soils, was found to be inadequate to express the

volumetric behavior of unsaturated soils (Matyas and Radhakrishna, 1968). The

effective stress condition in unsaturated soils, therefore, has been directly demonstrated

as a function of the total stress and the pore water pressure. The main theoretical

concept is to transfer a multi-stress medium to a single stress state continuum (Nuth and

Laloui, 2008).

Bishop (1959) also developed an effective stress equation as presented in

Equation (2.3).

(2.3)

where;

ua is the pore-air pressure,

The difference (σ-ua) is the net normal stress,

The difference (uw−ua) the matric suction ( ), and

is a parameter that is a function of the degree of saturation (S %).

= 1 for fully saturated soils and = 0 for dry soils.

32

The relationship between x and S was experimentally determined and

documented.

Another mechanistic approach was developed by (Lambe, 1960) who defined the

effective stress of unsaturated soils in terms of the inter-particle forces, as well as the

applied forces as shown in Equation (2.4).

(2.4)

where;

aa is the fraction of the total area that is air,

am is the fraction of the total area that is mineral,

aw is the fraction of the total area that is water, and

R and A are repulsive and attractive electrical forces.

Aitchison (1965) proposed a form of the effective stress equation in order to show

the relationship of matric suction and effective stress Equation (2.5).

(2.5)

where;

p" is the pore water pressure deficiency parameter, and

is a parameter that is a function of the degree of saturation ( varies from zero

to one).

Jennings (1961) modified the form of effective stress equation based on the soil suction

and a parameter β as illustrated in Equation (2.6).

(2.6)

where;

p’’ is the pore water pressure deficiency parameter, and

β is a statistical factor (β can generally be measured experimentally).

33

Richards (1967) included a new component in order to represent osmotic suction in the

effective stress equation, as illustrated in Equation (2.7).

(2.7)

where;

is an effective stress parameter,

is an effective stress parameter intended for solute suction,

hm is the matric suction element, and

hs is the solute suction element.

Aitchison (1965) published an effective stress equation based on Richards (1967)

as presented in Equation (2.8).

(2.8)

where;

P”s is the solute suction,

P”m is the matric suction, and

and are soil parameters ( and typically varies between zero and one).

Khalili and Khabbaz (1998) introduced the value of effective stress parameters through

experimental data as indicated in Equation (2.9).

(2.9)

where;

is a parameter that is a function of the degree of saturation (Khalili and Khabbaz,

1998).

34

Houlsby (1997) developed an effective stress concept to incorporate the stress and

strain variables. The equation was developed using the same form of Bishop (1959), but

through replacing the parameter ( ) with the degree of saturation as shown in Equation

(2.10).

(2.10)

where;

S is the degree of saturation, and

is a substitution factor (Houlsby, 1997).

As shown in the previous equations, the unsaturated soil behavior was found to be

better expressed using two independent stress components, normal stress and soil

matric suction (Bishop and Blight, 1963). Fredlund and Morgenstern (1976) also

reported that the volume change of unsaturated soils can be defined by any two of the

three possible stress variables, normal stress (σ – ua), matric suction, (ua – uw), and

effective stress (σ – uw). Fredlund and Rahardjo (1993b) noted that the pore-air

pressure is a significant parameter for most engineering real problems. The combination

of stress state variables, (σ – ua) and (ua – uw), has been the most extensively used

variables because the changes in total normal stress can be determined independently

of the changes in the pore-water pressure.

2.5.2 Soil Suction

Soil suction is characterized as the free energy state of soil water and is determined as

a function of the partial vapor pressure. Equation (2.11) shows the basis for measuring

the soil suction (Richards, 1965).

35

(2.11)

where;

is the total suction,

is the temperature,

is the universal gas constant,

is the specific volume of water,

is the molecular mass of water, and

is the relative humidity (presented as a decimal).

Soil suction is a vital stress component for evaluating the behavior of unsaturated

soils. Soil suction may present a significant enhancement for geotechnical engineering

analysis of earth structures (Oh et al., 2008). Croney and Coleman (1948) showed the

significant role of soil suction in identifying the mechanical behavior of unsaturated soils.

Aitchison (1965) established detailed definitions of the soil suction mechanisms. The

total soil suction of a soil can be subdivided into two main mechanisms (i.e. matric

suction and osmotic suction) (Marshall, 1958).

The matric suction is expressed as the difference between the pore-air and pore-

water pressures, ( = ua – uw), and is a function of the radius of the spherical surface

and the surface tension as shown in Equation (2.12) (Fredlund and Rahardjo, 1993b).

Matric suction is typically derived from the capillary actions associated with the surface

tension forces of the pore water within the soil structure. The surface tension is

associated with the intermolecular forces applied to the water molecules at the air-water

interface (Fredlund, 1979).

(2.12)

where;

is the matric suction,

36

is the radius of the spherical surface, and

is the surface tension.

Osmotic suction is originated from the salt concentration in the pore-water, and it

occurs in saturated and unsaturated soils (Fredlund and Rahardjo, 1993b). The

concentration of the salt solution drives an attraction to water molecules in the soil

structure. Osmotic suction reduces the energy state of water in the soil and is

associated with physico-chemical interactions between pore water and soil minerals

(Fredlund and Rahardjo, 1993b). Osmotic suction is typically considered significantly

less than matric suction and less sensitive to variations in soil moisture content

(Fredlund and Rahardjo, 1993b). Therefore, the changes in matric suction are

considered practically representative of changes in total suction for unsaturated soil.

Equation (2.13) shows the relationship between the mechanisms of soil suction

(Fredlund and Rahardjo, 1993a).

(2.13)

where;

is the total suction,

is the matric suction component, and

is the osmotic suction component.

2.6 Desiccation Cracks

Unsaturated clay soil deposits are characterized by the presence of naturally

occurring desiccation cracks. The topic of desiccation cracks phenomenon has been

covered in the past in a number of research studies (Kodikara et al., 2002; Miller et al.,

1998; Morris et al., 1992; Nahlawi and Kodikara, 2006; Yesiller et al., 2000). The

development of soil cracks influences the behavior of the near surface clay layer, and

consequently, the moisture content changes with time. The effect of desiccation cracks

37

occurring in highly plastic clay soils has been investigated and studied extensively using

different experimental methods. Albrecht and Benson (2001) found that the development

of cracks resulted in a noticeable increase in the hydraulic conductivity of clays, at times

as high as three orders of magnitude. Generally, desiccation cracks are formed because

of the soil shrinkage during the water loss process. The extent of desiccation cracks

below the ground surface has to be estimated so that it can be used an input in the soil-

weather models. The depth of cracking is often required for equilibrium analyses in

geotechnical engineering (Lau, 1987). Researchers have developed various methods to

compute the likely depth of desiccation cracking using various theoretical assumptions.

Fracture mechanics is the technical approach defining the crack propagation

criteria in materials mainly by evaluating the stress and strain fields near the crack tip

(Hanson et al., 1994). In the area of fracture mechanics, the cracking process is treated

as a mechanical process, and the cracking criterion is based on the critical stress field

(Anderson, 2005). The basic concept is that the crack propagation occurs if the energy

released upon crack growth is adequate to exceed the material resistance (Griffith,

1921). The material resistance involves the plastic work, surface energy, or other types

of energy dissipation caused by a crack (Anderson, 2005).

From a mechanical point of view, the stress distribution at the crack tip can be

assumed as presented in Figure 2.5, and therefore can be computed as a function of the

material stress intensity factor (fracture toughness) and applied tensile stress as shown

in Equations (2.14) to (2.16).

(2.14)

(2.15)

(2.16)

38

where;

r is the distance from the crack tip at inclination from the horizontal axis, and

is the fracture toughness.

The stress intensity factor (fracture toughness) was found in the literature for

clayey soils, 12.7 kPam1/2 (Ayad et al., 1997; Lee et al., 1988). The fracture

toughness can be defined as presented in Equation (2.17) (Lawn and Wilshaw,

1975a; Lawn and Wilshaw, 1975b).

(2.17)

where;

is the compression modulus,

is the Poisson's ratio, and

ζ is the specific surface energy of the soil, range, from 0.1 to 1.0 J/m2 (Lee and

Ingles, 1968).

In addition, the critical crack size is a material property specifying the size of a

crack that causes failure at a certain stress level. Consequently, the critical crack size

can vary significantly with the material type and can be defined as presented in Equation

(2.18) (Anderson, 2005).

(2.18)

where;

is the critical crack length in an infinite plate,

is a parameter dependent on the material properties and crack geometry, and

is the tensile stress.

Several studies (Blight and Williams, 1971; Briones and Uehara, 1977; Raats,

1984) applied the Griffith's brittle fracture theory on the desiccation cracking of soils. The

39

fracture toughness was defined as a mechanical parameter indicating the ability to

withstand a crack failure. This parameter was used to estimate the extent of desiccation

cracks. The relationship between fracture toughness and other mechanical parameters,

such as tensile strength, were previously examined (Wang et al., 2007). Corte and

Higashi (1964) reported that cracking by desiccation is somehow different from

mechanical cracking mainly because of the material loses mass during the process. Lau

(1987) presented a mathematical model for the determination of the cracking depth as a

function of the soil shear strength represented by the Rankine theory of lateral earth

pressure. Morris et al. (1992) proposed an analytical approach based on the linear

elastic theory to predict the cracking depth for various soil parameters and suction

values. These two methods were used to estimate the theoretical depth of cracking for

the native soils in this research study.

40

Figure 2.5: A theoretical model for cracking mechanical mechanism after (Anderson, 2005)

41

2.7 Hydraulic Conductivity Function

The hydraulic conductivity is a main soil property controlling the soil-water

interaction. The hydraulic conductivity for saturated soils can be assumed as a constant

value. In principle, water flows through the pore spaces containing water. Therefore, the

soil hydraulic conductivity is dependent on the water content for unsaturated soils

(Childs and Collis-George, 1950). The lower moisture content is, the lower the hydraulic

conductivity value is for unsaturated soils because of the lower availability of saturated

pores. Therefore, the hydraulic conductivity of unsaturated soil should be considered as

a function of matric suction (Gallage et al., 2013). As a general trend, the hydraulic

conductivity of soil decreases noticeably with the decrease in the degree of saturation

(Ng and Menzies, 2007).

LIoret and Alonso (1980) reported that the soil hydraulic conductivity is not

constant and is a function of volume-mass soil properties. The hydraulic conductivity

function was, therefore, expressed as the relationship between the hydraulic

conductivity and soil suction. Due to the difficulty of measuring the hydraulic conductivity

function, methodologies to estimate the physical properties of the pore space provided

insight in estimating the saturated-unsaturated hydraulic conductivity. Therefore,

prediction methods are frequently used and mostly based on the soil-water retention

characteristics (Durner, 1994). Depending on the severity of the structural changes

associated with the formation of desiccation cracks, the hydraulic conductivity may get

increased by several orders of magnitude (Johnston and Haug, 1992).

42

2.8 Soil Water Characteristic Curve (SWCC)

The soil water characteristic curve (SWCC) is a relationship between matric

suction and soil moisture condition under a normal stress of zero or (small value). The

SWCC is a critical soil property demonstrating the transient water flow through

unsaturated soils. The soil moisture characteristics may be presented in terms of

volumetric water content, gravimetric water content, or degree of saturation. SWCC

function describes the water storage capacity of a soil at various soil suctions. A review

of the literature indicated the presence of different graphical presentations for the

SWCC. Soil suction and water content (volumetric or gravimetric) can be plotted on the

abscissa, ordinate, or on a logarithmic scale. Fredlund et al. (2001) recommended that

soil suction be plotted on the abscissa and water content be plotted on the ordinate. A

standard unimodal SWCC for a drying cycle is shown in Figure 2.6.

The main parameters always identified on SWCCs are the air entry value and the

residual water content. The air-entry value is identified as the matric suction at which air

begins to go through the pores in the soil structure. The residual water content is the

water content at which a considerable amount of suction change is required to take out

any additional water from the soil structure (Fredlund and Xing, 1994). The residual

water content may be determined as the intersection of two lines, a tangent line drawn

from the inflection point and a tangent line drawn from the high-suction range as shown

in Figure 2.6 (Fredlund and Rahardjo, 1993b). The soil suction at a water content of zero

value was found to be approximately 106 kPa (Fredlund and Xing, 1994). White et al.

(1970) identified three main de-saturation stages along the drying curve, namely the

boundary effect phase, the transition state, and the residual stage. Similar stages of

saturation were found to be applicable to the wetting curve (Fredlund, 2000).

43

The most commonly known shape of the SWCC is the unimodal SWCC which

contains two bending curves only. The capacity of fine-grained soils to retain water

depends on the soil fabric including macropores, mesopores, and micropores (Elkady,

2014; Mitchell, 1993). The soil native structure and grain size distribution are

contributing factors affecting the nature of the SWCC (Fredlund and Rahardjo, 1993b;

Vanapalli et al., 1998). Recent research studies have shown bimodal SWCCs with two

distinct structures of the pore size distribution (de FN Gitirana Jr and Fredlund, 2004;

Satyanaga et al., 2013; Zhang and Chen, 2005). Bimodal SWCCs are typically observed

in the case of soils with cracks, and it can be utilized to simulate the changes in soil

water content in the field. The bimodal curve defines two air entry values and two

residual volumetric water contents including near horizontal or horizontal intermediate

segment (Elkady, 2014).

44

Figure 2.6: Typical soil water characteristic curve of clay soil

Matric Suction, (kPa)

Volu

metr

ic W

ate

r C

onte

nt,

(a,s)

(r,r)

Air-Entry Value (AEV)

Residual Volumetric Water Content

Saturated Volumetric Water Content

45

2.9 Measurements of Soil Moisture-suction Characteristics

2.9.1 Foreword

Measuring the soil-water characteristics using field instrumentation is essential in

understanding the behavior of underground structures and the soil-water interaction

(Fredlund, 2006). Monitoring the soil-water characteristic in the field is an efficient tool to

validate the results of the numerical studies. Various types of instruments are employed

in measuring the soil water characteristics. These methods determine the soil moisture

using calibrated relationships derived from other measurable parameters (Muñoz-

Carpena et al., 2004). The suitability of each method relies on a number of aspects (i.e.

cost, installation, accuracy, purpose, response time, management and durability) (SAI

Platform, 2010). Primary methods and techniques normally utilized for measuring the

soil-water characteristics are classified into volumetric and tensiometric techniques as

shown in Figure 2.7, and discussed in the following subsections.

2.9.2 Soil Moisture Monitoring (Principle and Techniques)

There is a range of readily available techniques for the determination of soil

moisture conditions. The moisture content of a given soil can be measured through

direct and indirect techniques (Muñoz-Carpena et al., 2004). The direct measuring

technique can be achieved through a thermo-gravimetric technique (i.e. weight of water

over the weight of dry soil) or a thermo-volumetric technique (volume of water within a

volume of undisturbed soil) (Muñoz-Carpena et al., 2004). These direct techniques are

laboratory based methods and, therefore, are accurate and inexpensive. However, they

are destructive, time-consuming and cannot be used to monitor the moisture content in

the field (Agricultural Technology Centre, 2004).

46

Various field monitoring techniques (i.e. indirect techniques) were developed for

measuring the soil volumetric water content in the field (SAI Platform, 2010). It is

essential to understand the differences between these available techniques in order to

choose the most suitable instrument for the proposed function. All the volumetric

measuring methods measure the volume of water within a volume of undisturbed soil

(SAI Platform, 2010). The available techniques include neutron moderation technique,

dielectric techniques, and other techniques (such as Ground Penetrating Radar (GPR),

Gamma Attenuation, Nuclear Magnetic Resonance, Capacitive Sensor, and Optical

Methods) (Muñoz-Carpena et al., 2004). Table 2.5 provides an evaluation of the most

common water measurements techniques. Amongst these available methods, the

Neutron moderation, Time Domain Reflectometry (TDR), Water Content Reflectometry

(WCR), Frequency Domain Reflectometry (FDR), Amplitude Domain Reflectometry

(ADR) methods can be used for research purposes (Muñoz-Carpena et al., 2004). The

neutron moderation apparatus estimates the volumetric water content by measuring the

thermal density or the slowed neutron density (Agricultural Technology Centre, 2004) .

The neutron moderation is derived from a linear calibration between a probe reading ,

and the measured soil moisture content using field samples (Muñoz-Carpena et al.,

2004). This technique may be used to assess the soil moisture at various depths below

the ground surface (SAI Platform, 2010).

Dielectric methods are widely used to monitor soil volumetric water content in

coarse-grained or fine-grained soils with low to medium plasticity (Hu et al., 2008).

These methods consist of the time domain reflectometry (TDR), the water content

reflectometry (WCR), the amplitude domain reflectometry (ADR), and the frequency

domain reflectometry (FDR) (Muñoz-Carpena et al., 2004). The dielectric techniques

utilize empirically verified relationships between the volumetric water content and the

sensor output magnitude (mainly time, frequency, impedance and wave phase) (Muñoz-

47

Carpena et al., 2004). These generic calibration relationships were developed for typical

soils and are sensitive to the soil mineralogy, bulk density and temperature. The use of

dielectric techniques in highly plastic, unsaturated, expansive clay soil is limited, and

therefore, soil-specific calibrations are considered necessary (Hu et al., 2008).

2.9.3 Soil Suction Monitoring (Principle and Techniques)

Tensiometer systems are methods for measuring the energy holding water in soils

(Muñoz-Carpena et al., 2004). Tensiometers are devices for determining the energy of

the soil solution (known as, the soil water matric potential) (Soil Science Society of

America, 2008). They are typically able to monitor the soil tension variations occurring

during different seasonal event such as infiltration, irrigation, groundwater recharge and

evapotranspiration processes (Agricultural Technology Centre, 2004). Tensiometers

consist of three essential elements, a porous cup, a water reservoir, and a measurement

gauge (Young and Sisson, 2002). Tensiometers generally require the gauge to reach

equilibration with the soil moisture. Table 2.6 provides an evaluation of the most

common tensiometric techniques. Amongst the available methods, tensiomneters, heat

dissipation, and soil pyschrometers are the ones which can be used for research

purposes (Muñoz-Carpena et al., 2004). Tensiometers may or may not get affected by

osmotic or gravitational potentials depending on the manufacture (Young and Sisson,

2002). The soil water potential measured by a tensiometer always captures the matric,

pneumatic and overburden potentials (Young and Sisson, 2002). Generally, tensiometer

systems may not require a soil specific calibration/validation, however, they may have to

be installed permanently, or at least a sufficient time period has be provided in order to

develop equilibration between the tensiometer and soil before retrieving the readings

(Agricultural Technology Centre, 2004).

48

Figure 2.7: Measurement methods of soil-water characteristics

49

Table 2.5: Evaluation criterion for volumetric soil water monitoring techniques (Muñoz-Carpena et al., 2004)

Specification Neutron TDR/WCR FDR ADR Phase

Transmission

TDT

Reading range

0 - 0.60

0.05 - 0.50 or

0.05 - Saturation (with soil specific

calibration)

0 - Saturation 0-Saturation 0.05-0.50

0.05 - 0.50 or

0 - 0.70 Depending

on instrument

Accuracy (considering soil specific calibration)

±0.005 ±0.01 ±0.01-0.05 ±0.01 ±0.05

Installation technique

Access Tube

Permanently

buried in the field

or inserted for

manual readings

Permanently buried in the field or PVC access

tube

Permanently buried in the

field or inserted for

manual readings

Permanently buried in the field

Logging capacity

No Depending on

instrument Yes

Influenced by salinity

No Yes Minimal No >3 dS/m Yes

Not commonly used for

None

Organic,

dense, salty or

high clay content

soils (depending

on instrument &

calibration)

None None None

Organic,

dense, sallty or high clay content soils (depending

on instrument & calibration)

Field maintenance

No

Safety hazard Yes No

Applications Irrigation, Research, Consultants Irrigation, Research Irrigation

50

Table 2.6: Evaluation criterion for soil suction monitoring techniques (Muñoz-Carpena et al., 2004)

Specification Tensiometer Gypsum Block GMS Heat

Dissipation Soil

Psychromotor

Reading range

0 - 0.80 bar 0 - Saturation 0-Saturation 0.05-0.50

0.05 - 0.50 or

0 - 0.70 Depending on the

instrument

Accuracy, bar (considering soil

specific calibration) ±0.01 ±0.01 ±0.01-0.05 ±0.01 ±0.05

Installation method

Get inserted into hole permanently

Logging capacity Only when

using transducers

Yes

Influenced by salinity

No >6 dS/m >6 dS/m No Yes for

(ceramic type)

Not commonly used for

Sandy or coarse soils

Sandy or coarse grained soils

swelling clay

Coarse soils

Sandy or coarse grained

soils

swelling clay

Maintenance Yes No Medium No No

Safety hazard No

Applications Irrigation, Research

Irrigation Irrigation, Research

Research

51

2.10 Unsaturated Soil-atmosphere Interaction

2.10.1 Foreword

The main objective of soil-atmosphere transport models is to determine and

evaluate the soil exchange fluxes between soil and atmosphere. Soil-atmosphere

interaction models involve the estimation of the rates of heat and moisture flow through

the soil structure. The boundary conditions depend on two principal components,

evaporation (or evapotranspiration) and precipitation. Evaporation is related to free

water surfaces while evapotranspiration is associated with wet soil surfaces (Veihmeyer,

1964). The following subsections discuss the soil evaportranspiration process, and

water and heat flow through soil structures.

2.10.2 Soil Evaportranspiration

The understanding of the evapotranspiration process was established in the

1940’s (Gitirana Jr, 2005). Early studies reported that the evapotranspiration process

occurs under three main conditions: (i) a continuous supply of energy should be existing

for the latent heat of vaporization, (ii) the vapour pressure of the atmosphere above the

surface should be less than the vapour pressure at the soil surface, and iii) the vailability

of a continual supply of water to the evaporating surface (Baver et al., 1972; Hillel, 1982;

Penman, 1948; Veihmeyer, 1964; Wilson, 1990). For soils, the provision of water is

influenced by soil conditions (i.e. soil properties, moisture content, and groundwater

table). Generally, the shallower the groundwater level is, the higher the

evapotranspiration rate (Gardner, 1958).

A detailed review of the equations for calculating the potential evapotranspiration

(PE) can be found in (Wilson, 1990). Penman (1948) developed an equation that is a

very popular one for indirectly estimating potential evapotranspiration. This equation

takes into consideration the effect of a number of essential meteorological factors (i.e.

52

air temperature, relative humidity, net solar radiation and wind speed). Modified forms of

this equation were also developed by different scholars in subsequent publications

(Black, 2007). Wilson (1990) and Wilson et al. (1997) developed partial differential

equations for calculating evapotranspiration fluxes from saturated/unsaturated soils. The

approach was experimental in nature, but had a sound theoretical basis, and was

therefore adopted in this research work.

2.10.3 Water Flow

The main concept of modeling of the in/exfiltration flux into soils was developed

based on the law of conservation of mass. The total mass of water per unit volume of

the soil medium can be presented as shown in Equation (2.19).

(2.19)

where;

is the water density (998.2 kg/m³);

θ is the volumetric moisture content (vol/vol),

ρv is the vapour density (kg/m³), and

θa is the volumetric air content (vol/vol).

Mass conservation would then result in Equation (2.20).

(2.20)

where;

t is the time, and

qm is the total moisture flow (kg/m²s).

The hydraulic head gradient can be expressed as the driving force of water to flow

through the soil structure from points of high head to points of low head. Other

driving forces include moisture content and matric suction gradients. However, these

gradients do not fundamentally govern the water movement through the soil structure

53

(Fredlund and Rahardjo, 1993b). The flow net technique was initially used for predicting

seepage through earth assuming that water only flows below the ground water table

(Casagrande, 1937). Finite element methods have alternated the flow net technique due

to their powerful nature. Due to the complexity of modeling seepage in

saturated/unsaturated soils, It is viable to employ general partial differential equation

solvers (Nguyen, 1999). The water flow rate in saturated/unsaturated soils can be

identified using a generalisation of Darcy’s Law, where the total head gradient and the

hydraulic conductivity are functions of matric suction (Bear, 1972; Fredlund and

Rahardjo, 1993b). The generalised Darcy’s law may be expressed as shown in Equation

(2.21).

(2.21)

where;

and are liquid pore-water flow rates in the x- and y- direction across a unit

area of the soil due to the hydraulic head gradients (m3/m2s),

, and are hydraulic conductivity (m/s) in the x-, and y- direction, and

is the total hydraulic head (m).

In the field, piezometers are used to provide a measurement of the water head in

saturated soils. The hydraulic conductivity is approximately constant for saturated soils.

Above the water table, the pressure heads are negative, and the hydraulic head can be

determined indirectly through the field measurement of the suction head or negative

pore-water pressure. Childs and Collis-George (1950) reported that water flows only

through the pore spaces filled with water and the soil hydraulic conductivity is a function

of the soil suction for unsaturated soils. Under unsaturated soil conditions, few saturated

54

pores are available for water to flow, and consequently, the lower moisture content is the

lower the hydraulic conductivity value.

LIoret and Alonso (1980) reported that the soil hydraulic conductivity is not

constant and is dependent on volume-mass soil properties. The hydraulic conductivity

function was, therefore, defined as the relationship between the hydraulic conductivity

and soil suction. During the de-saturation process of soil, air begins to enter the large

pore spaces first, and water flow is forced to move along the more tortuous path

in the smaller soil pores. The hydraulic conductivity decreases rapidly as the volume of

pore space occupied by water decreases. Although Darcy’s law was developed for

predicting the flow in saturated soils, it was then applied to the flow of water through

unsaturated soil with the consideration of the change in the hydraulic conductivity

(Richards, 1931). There are different versions of Richards’ equation for the

determination of moisture movement through saturated-unsaturated soils. All these

versions were mainly based on calculating water continuity in terms of matric suction.

Equation (2.22) illustrates the two-dimensional partial differential form of Richards’

equation.

(2.22)

where;

is the total hydraulic head (m),

kx is the hydraulic conductivity (m/s) in x-direction,

ky is the hydraulic conductivity (m/s) in y-direction, and

θ is the volumetric water content (VWC) (vol/vol).

Fredlund and Rahardjo (1993a) presented a general partial differential equation derived

from Richards’ equation for the transient moisture flow through saturated/unsaturated

soils as shown in Equation (2.23).

55

(2.23)

where;

h is the total head (m),

is the slope of the soil water characteristic curve, and

is the unit weight of water (kN/m3).

2.10.4 Heat Flow

Soil temperature is a critical factor controlling many physical processes or changes

that take place in soils. Soil temperature changes are attributable to the changes in

radiant, thermal and latent energy exchange processes (Hillel, 1982). These processes

are controlled by a complex series of heat transfer actions (i.e. radiation, conduction,

convection, advection, and phase change) known as latent heat transfer or evaporation

(Rutten et al., 2010). The heat flow element of greatest concern is the conductive heat

flow. In soils, convection heat transfer of the pore-fluid is noticeably less than conductive

heat transfer (Milly, 1984) and, therefore, can generally be neglected.

The differential form of Fourier's Law has been widely used to compute the vertical

heat flux in materials (Fourier, 1878). Differential equations for the conservation of heat

in soils were developed, taking into consideration the flow rates in and out of the soil

elemental volume, and corresponding to the difference in the rate of change in heat

stored with time (Fredlund and Gitirana Jr, 2005). The main factor in the heat flow

equation that describes the soil’s ability to transfer heat is the thermal conductivity. The

thermal conductivity of soils was found to be significantly influenced by the degree of

saturation and dry density (Fricke, 1997). As a general trend, an increase in the degree

of saturation or dry density results in an increase in thermal conductivity (Fricke, 1997).

Other factors that have a secondary effect upon soil thermal conductivity include frozen

56

versus unfrozen soil conditions, mineral composition and texture (Becker et al., 1992;

Fricke, 1997; Mitchell, 1991).

Different correlations for estimating the soil thermal conductivity were developed

for both coarse- and fine-grained soils (De Vries, 1952; Johansen, 1975). Al

Nakshabandi and Kohnke (1965) reported that the thermal conductivity of soils at the

same moisture condition is highest in granular soils (gravel and sand), intermediate in

loam, and lowest in clay soils. Kersten (1949) tested various soil types and developed

equations for frozen and unfrozen silt-clay soils and sandy soils. Canadian Geotechnical

Society (1978) provided a generalized basis for estimating the thermal conductivity of

frozen soils based on Kersten's research and was adopted in this research to estimate

the initial thermal conductivity for the top clay soil layer.

2.11 Unsaturated Soil-structure Interaction

2.11.1 Numerical Approaches

The main technical aspects of modeling the soil-structure interaction under

unsaturated soil conditions include the degree of simplicity, practicality to buried

structures, insufficiency of site local specific parameters and field verifications, and lack

of knowledge of some features. Soil behavior can generally be described by constitutive

relationships. Soil behavior can be described by constitutive relationships as a function

of stress state variables. The constitutive relationships provide a framework for

understanding how soil behaves under different loading conditions, and can be

formatted in finite element and finite difference codes for use in numerical analysis.

Modeling of the volume change of unsaturated soils consists of stress-deformation

analysis under mainly water and heat flow processes. The volume change behavior of

unsaturated soil can be modeled using coupled or uncoupled approaches.

57

In the uncoupled approach, the water seepage equation (adsorption-drainage

processes) is solved separately from the stress-deformation equations (mechanical

behavior) (Corapcioglu, 1984). For the seepage analysis, the dependent variable is the

hydraulic head. However, for the stress-deformation analysis, the dependent variables

are the horizontal and vertical displacements. In the coupled approach, however, the

water seepage equation and the equilibrium equations are solved simultaneously

(Pham, 2005). Previous research results showed that the outcome of uncoupled

solutions compared well with those from the coupled solutions, and uncoupled solutions

are adequate for the analysis of most volume change problems related to unsaturated

expansive soils (Vu and Fredlund, 2004).

2.11.2 Volume-mass Constitutive Relationships

A number of volume-mass constitutive models have been developed for

unsaturated soils (Pham, 2005). The typical constitutive relationships for unsaturated

soils were expressed as an extension of the saturated soil constitutive equations and

incorporated the use of two independent soil properties (total normal stress and soil

suction). The development of constitutive models for unsaturated soils is directly related

to the unsaturated soil stress variables. The stress state variables for an unsaturated

soil are generally interpreted from possible combinations of the total stress, σ, the pore-

air pressure, ua, and the pore-water pressure, uw. The combination of net stress (σ – ua),

and matric suction (ua – uw) presented in Figure 2.8 are the most widely one (Fredlund

and Vanapalli, 2002; Fredlund and Rahardjo, 1993a). Tensors for these two

independent stress variables can be formulated as shown in equations (2.24) and (2.25)

(Fredlund and Rahardjo, 1993b).

58

(2.24)

(2.25)

where;

The three normal stress parameters ( , , and ) are mutually

orthogonal with respect to x-, y- and z- direction, and

The six shear-stress parameters ( act on the i plane and on the j direction.

Volume change of unsaturated soils can occur as a result of changes in the total

stress, the matric suction, or a combination of the two cases. Fredlund and Rahardjo

(1993a) presented the theoretical basis for developing the volume-mass constitutive

surfaces for saturated-unsaturated soils. The change in void ratio as a deformation

(volume change) state variable for saturated soils was defined as a function of the net

normal stress as presented in Equation (2.26). Both void ratio and gravimetric water

content for unsaturated soils were defined as functions of the net normal stress and soil

suction as presented in Equations (2.27) and (2.28) (Fredlund and Rahardjo, 1993a).

Using these definitions, the constitutive relationships (equations) can be plotted as

surfaces on a three-dimensional graph, with each abscissa representing a stress state

variable, and the ordinate representing the soil volume-mass property as presented in

Figure 2.9 (Fredlund and Rahardjo, 1993a).

σ (2.26)

σ (2.27)

σ (2.28)

59

where;

is the change in void ratio,

is the change in water content,

is the coefficient of compressibility,

is the coefficient of compressibility as a function of the change in net normal

stress, σ ,

is the coefficient of compressibility as a function of the change in matric

suction, ,

is the coefficient of water content change as a function of the change in net

normal stress, σ , and

is the coefficient of water content change as a funtion of the change in matric

suction, .

The development methods of the volume change principle for unsaturated soils

were presented in (Fredlund and Rahardjo, 1993b). Alonso et al. (1990) established the

volume change constitutive relationships for unsaturated soils based on the net stress

and soil suction. The prediction of volume change is based on the relationship between

vertical strain (i.e. void ratio) and the logarithms of soil suction or net normal stress (Vu

and Fredlund, 2004). The fundamental concept is that the total volume change of

unsaturated soils is equivalent to the summation of volume change magnitudes

associated with the air and water phases (Fredlund and Morgenstern, 1976). The overall

volume change of unsaturated soil components can be described as shown in Equation

(2.29) (Fredlund and Rahardjo, 1993b).

(2.29)

where;

V is the total volume,

60

Vw is the volume of water, and

Va is the volume of air.

Due to the difficulty in the determination of the air volume changes in the soil

structure, the practical way is to predict the change in overall and water volumes. The

displacements associated with the change in total soil volume can then be computed as

the sum of the normal strains as presented in Equation (2.30) (Fredlund and Rahardjo,

1993a). The normal strain in a given direction, ε, is computed as the change in length

per unit length of a line. Shear strain, ζ, is, however, formulated as the change in the

right angle between referenced axes (Chou and Pagano, 1992). The relationships

between the normal and shear strains and displacements in x, y, and z directions are as

illustrated in equations (2.31) and (2.32) (Fredlund and Rahardjo, 1993b).

(2.30)

(2.31)

(2.32)

where;

is the volumetric strain,

are the normal strain components in x-, y-, and z- direction, respectively,

is the shear strain, and

are the displacements in x-, y-, z- direction, respectively.

The main principles for establishing the volume change models were based on

either the elasticity theory or empirical models. The empirical models are derived from

mathematical models fitting the relationship between stress state variables obtained

from laboratory tests measurements. The elastic models, however, are based on

relevant parameters (i.e., modulus of elasticity and Passion's ratio) and are relatively

61

simple to be used in the numerical analysis (Wheeler and Karube, 1995). Fredlund

and Rahardjo (1993b) proposed an equation based on a semi-empirical approach

derived from the linear relationship between the volumetric water content and stress

variables. Fredlund and Rahardjo (1993b) presented the volume-mass constitutive

relationship for a soil structure in an incremental elasticity form as shown in Equation

(2.33). The coefficients of volume change, and

, were defined as the slopes of the

soil structure constitutive surfaces and can be computed by differentiating the surfaces

of net normal stress and matric suction (Fredlund and Rahardjo, 1993b). The

constitutive equations were based on the following main assumptions: (i) the air phase is

continuous and remains at atmospheric pressure; (ii) the soil is elastic, nonlinear and

isotropic; (iii) the pore water is incompressible; and (iv) the effects of the air diffusing

through water, air dissolving in water and water vapor movement can be neglected (Vu

and Fredlund, 2004).

(2.33)

where;

is the mean net total stress,

is the coefficient of volume change as a function of a change in net normal

stress, and

is the coefficient of volume change as a function of a change in matric suction.

62

Figure 2.8: The unsaturated soil stress state parameters using the combination of (σ – ua), and matric suction (ua – uw) (Fredlund and Vanapalli, 2002)

Figure 2.9: Constitutive surfaces for an unsaturated, swelling soil (Fredlund and Rahardjo, 1993b)

Normal S

tress

Soil Suction

Volume Change

Swell-Shrink Relationship

Consolidation Relationship

63

2.12 Pipelines Infrastructure

Buried pipes are vital infrastructures and are typically used to transport energy

and other essential commodities. Pipeline systems have improved the standard of living

and have rapidly grown in use over the last 60 years. The failure of underground

pipelines occurs when the applied stresses exceed its structural resiliency. A

comprehensive analysis of buried pipes should consider pipe characteristics, internal

and external loads, and surrounding conditions such as backfill and side fill materials,

installation depth, compaction quality, and road superstructure loads. Pipes can be

classified as either flexible or rigid, depending on whether they can deform up to 2%

without incurring damage (Suleiman, 2002).

Rigid pipes, such as reinforced/non-reinforced concrete, and clay pipes, may

experience significant structural cracks if they deflect more than 2% (Zhao et al., 1998).

Flexible pipes have been defined as conduits that can deflect at least 2% without

exhibiting any sign of structural distress such as cracking (Uni-Bell PVC Pipe

Association, 1991). Flexible pipes include thermoplastics [i.e., Polyvinyl Chloride (PVC)

and High Density Polyethylene (HDPE)], thermosetting, and corrugated steel pipes. The

relatively low cost and high processability of PVC make it the material of choice for

various industries.

Soil-pipe interaction differs between flexible and rigid pipes. A rigid pipe is

responsible for transferring the applied loads to the bedding material. Rigid pipes are

generally stiffer than the surrounding soil (Zhao et al., 1998). Flexible pipes support the

applied vertical loads through passive pressures induced by the pipe deformation

against the surrounding soils (Moser, 1990). Flexible pipes have less inherent stiffness

when compared to rigid pipes. Therefore, flexible pipes usually require efficient

compaction of the backfill soils during installation. The design of buried pipes in North

America began in the early 1900s, initiated mainly by Marston and Anderson (Marston

64

and Anderson, 1913). The design criteria of buried PVC pipes incorporate the

determination of sufficient pipe stiffness to resist buckling. Pipe deformation must be

limited to eliminate any disruption in the flow or joint leakage (American Water Works

Association, 2002). General design guidelines for underground PVC pipes can be

summarized as follows: (i) limiting the pipe deformation to a maximum of 3% [i.e.

thermoplastic pipe (Zhao et al., 1998)], and 5% to 7.5% [i.e. PVC pipes (Eagle, 2009),

based on the pipe working pressure]; (ii) preventing buckling of the pipe wall; and (iii)

limiting any potential for ‘wall crush’ that may occur due to vertical loads (soil and live

loads) applied directly above the pipe.

Pipes must have sufficient strength to achieve its intended function. Pipe strength

is the ability to resist stresses, mainly internal pressure, soil pressure, live loads, and

longitudinal bending moments/stresses. Pipe stiffness describes the load-deformation

characteristics of flexible plastic pipes, thus can be used to determine the deformations

of the pipe wall. The pipe durability, however, is the ability to resist corrosion, abrasion

and deleterious environmental conditions, and it is a significant parameter for

determining the design service life for the performance of any pipe type.

2.13 Applied Loads on Buried Pipes

2.13.1 Foreword

The loads borne by the pipes depend primarily on the pipe type. For rigid pipes,

the pipe is expected to resist any vertical pressure as well as any horizontal reacting

earth pressure. However, flexible pipes affected by vertical loads experience

deformations resulting in a supporting horizontal soil pressure (Moser, 1990). Typical

loads applied on buried pipes can be classified into dead and live loads. Dead loads

generally remain static throughout the life of the pipes, and it considers any sole dead

65

load. On the other hand, live loads always change in position and/or magnitude. The

most common live loads borne by pipes are vehicular loads (i.e., trucks or trains).

The design of any buried pipe system should consider both types of superimposed

loads, dead loads and live loads as separate design parameters (American Water

Works Association, 2002). The level of soil stresses on the pipelines depends mainly on

the soil nature soil, its degree of saturation, and other relevant factors, such as seasonal

meteorological variations, which cause periodical changes in soil moisture. It should be

noted that the superimposed loads on the underground pipes can vary significantly

throughout the life of the pipeline due to the rise of any of the following conditions

construction/ installation conditions, routine operational conditions, and extreme loading

conditions, such as landslides, and earthquake-induced ground movements

(Wijewickreme and Weerasekara, 2010).

2.13.2 Soil Load

Design earth loads can be determined using the Marston load theory (Marston,

1930). According to the theory, it was assumed that the weight of the backfill partly

resisted by frictional shear forces at the trench walls developed with time. In addition,

the apparent cohesion of the soil was ignored when determining the equilibrium of

vertical forces (Moser, 1990). Marston load theory determined the load on a buried pipe

based on the soil column weight, or central prism, which was modified by a factor that

incorporated the effect of the relative movement between the side columns of soil, or

external prisms to the central prism (Marston and Anderson, 1913). The frictional forces,

between soil prisms, can then be determined using Rankine’s theory. The soil load

applied to buried pipe for rigid pipes was determined as shown in Equations (2.34) and

(2.35). Values of the ratio of lateral to vertical earth pressure and coefficient of friction

against sides of the trench were previously estimated for different typical soil types

66

(Moser, 1990). The coefficient of friction generally ranged from 0.3 to 0.5, which

corresponded to the angle of friction values ranging from 17o to 27o (Moser, 1990).

(2.34)

(2.35)

where;

ws is the soil load on a buried pipe,

is the unit weight of backfill,

W is the horizontal width of the pipe trench,

Cd is a load coefficient for the trench,

’ is the coefficient of friction between backfill and sides of trench,

K is the ratio of active lateral pressure to vertical pressure, and

H is the height of backfill on top of the pipe.

The load on flexible pipes was found to be different from rigid pipes. It can be

determined if the relative stiffness of pipe and backfill is reasonably estimated

(Cameron, 2005). Theoretically, when both stiffnesses of soil and pipe are assumed to

be equal, then the load can be approximated to the prism load. The prism load can be

computed by the weight of the soil column directly applied above the pipe as shown in

Equation (2.36). The prism load neglects any external factors acting on pipes, such as

side wall friction. Therefore, the prism load approach can be used as a conservative

approach for flexible pipes (American Water Works Association, 2002). The actual load

on pipe may be more or less than the prism load based on the soil-pipe stiffness. When

the pipe stiffness is less than the soil stiffness, which is common for flexible pipes, the

soil above the pipe redistributes the load away from the pipe into the soil (Petroff, 1990).

67

(2.36)

where;

P is soil prism load on the pipes at a depth of H due to soil weight.

Most of flexible pipes would fall into the case of which the load imposed on the

pipe is less than the soil prism load over the pipe as shown in Equation (2.37) (Moser,

1990). For comparison purposes, the ratio of soil load applied on a flexible pipe to a rigid

pipe can be calculated as shown (2.38) which is equal to the ratio of the pipe diameter to

the trench width.

(2.37)

(2.38)

where;

D is the outside diameter of the pipe,

W is the horizontal width of the pipe trench, and

Cd is the load coefficient for the trench.

2.13.3 Live Loads

The live load is the other type of loading acting on underground pipes. Pipelines

underneath roads or railways also experience live loads. The live loads can be classified

into two types, concentrated live loads (such as a truck wheel), or distributed live loads

(Moser, 1990). The effect of live loads on a pipe decreases as the depth of soil cover

increases. The determination of live load is important for underground pipes with shallow

depths (such as less than 1.5 m) (American Water Works Association, 2002). The

method developed by Hall and Newmark (1978) uses a load coefficient (Cs) for

calculating concentrated loads as shown in Equation (2.39). The loaded surface area is

considered as a rectangle, and a truncated pyramid is punched through as what can be

determined by Equation (2.40) (Watkins and Anderson, 1999).

68

(2.39)

(2.40)

where;

Wsc is the live load on pipe due to a concentrated load,

Wsd is the live load on pipe due to a distributed load,

P is a concentrated surface load,

Pd is the intensity of a distributed surface load,

F is the impact factor,

Cs is a load coefficient, and

L is effective length of the pipe (1 m is typically used).

2.14 Pipe Stresses

Figure 2.10 shows the three principal stresses in a pipe, namely longitudinal stress

(σx) circumferential or tangential stress (σy), and radial stress (σz). According to most

design practices of pipelines, the radial stress σz is negligibly small (Young et al., 2011).

The major stresses σx and σy are considered uniform throughout the thickness of the

wall as the pipe thickness is much smaller than the pipe diameter, especially when the

ratio of the mean pipe radius to the wall thickness is more than 10.

Tangential stresses and associated bending moments are typically caused by pipe

crushing under the application of external loads. Rigid pipes and many flexible pipes

are not designed to resist high longitudinal stresses (Moser, 1990). Any uniform axial

compressive stress can produce a uniform axial tensile stress on the pipe wall as

illustrated in Figure 2.11 (Ng, 1994). Longitudinal stresses can also be produced by

different loading and environmental actions, including thermal expansion and

longitudinal bending. Longitudinal bending occurs when pipes are bent due to any

directional changes as shown in Figure 2.12 (Ng, 1994).

69

Figure 2.10: Principle stresses of a pipeline (Ng, 1994)

Figure 2.11: Stresses in a pipeline under longitudinal extension (Ng, 1994)

Figure 2.12: Stresses in a pipeline under longitudinal bending conditions (Ng, 1994)

70

2.15 Pipe Deformations

2.15.1 Horizontal Deformation

Spangler (1941) reported that flexible pipes provide limited inherent stiffness

compared to rigid pipes. By using laboratory testing results, Spangler handled the

influence of surrounding soils on the change in pipe shape as illustrated in Equation

(2.41) that is known as developed Iowa Formula. The determination of the pipe

deformation was based on the “ring theory”, assuming that the loss in vertical diameter

is compensated by an increase of the same magnitude in the horizontal diameter,

whereas the deformed pipe shape is elliptical as shown in Figure 2.13 (Moser, 1990).

This basic assumption imposes the importance of using advanced finite element

analysis and methods to better evaluate the actual behavior of pipelines.

The pipe deformation is expressed as the ratio of vertical deformation to horizontal

enlargement of the pipe diameter. The soil load in Iowa Formula can be determined

using the definition of the soil column load on underground pipe developed by (Marston

and Anderson, 1913).

(2.41)

where;

is the change in pipe diameter (flexible pipe),

is a deflection lag factor,

is the bedding constant,

Ep is the elastic modulus for the pipe material (kPa),

I is the moment of inertia of the pipe wall per unit length (m3),

E is the modulus of passive resistance of soil (kPa),

r is the mean pipe radius (m), and

71

is the soil load on the buried pipe (kN/m).

Spangler’s equation (Iowa Formula) was based on three main assumptions as follows:

The vertical deformation is equivalent to the horizontal deformation,

The pipe deformation is elliptical, and

The modulus of soil reaction is considered constant for the backfill material.

In the case of sustained loading conditions, the deflection lag factor (Dl) was found to

increase with time (Howard, 1985). Prevost and Kienow (1994) reported that the angle

of bedding support may be assumed around 90° in the determination of the deformation

and moments at a bedding constant (K) value of 0.012.

2.15.2 Bending Displacement

There are different loads that may produce axial bending stresses in a pipe. Some

of the major causes of bending displacements include the following (Moser, 1990);

The differential settlements of the structure that is directly connected to the pipe;

Settlements of the pipe bedding (i.e., erosion of the soil),

Seasonal rise and fall of soil,

Non-uniform foundation conditions, and

Tree-root growth induced pressure.

These loads can be significant, highly variable, and localized and can cause damages to

underground pipeline networks as being discussed in the following section. Figure 2.14

shows a mechanism of pipe bending displacement.

72

Figure 2.13: Pipe deformation diagram (Ring Theory)

Figure ‎2.14: Pipe displacement due to axial bending

73

2.16 Buried Pipe Damages

2.16.1 Failure Mechanisms Associated with Soil Movements

Failure in pipes is typically defined as a state attributable to break, collapse, or

bending so that the structure no longer fulfills its purpose even though there is no

catastrophic failure (Cassa, 2008; Matthews et al., 2013). There are different pipe failure

mechanisms that are caused or deemed to be caused by driving conditions, original pipe

laying or successive changes in ground conditions (MacKellar and Pearson, 2003). The

excessive loads applied on pipes may cause failure due to crushing or compression in

the pipe wall, pipe wall bending, longitudinal bending, excessive deflection, tensile

failure, or buckling. (Matthews et al., 2013). Failures may also occur in the pipe, joints,

or mechanical fitting valves/meters. Figure 2.15 shows the most common soil movement

induced failure modes for pipe networks.

Longitudinal cracking is a failure nature mainly occurring in large diameter pipes

causing the pipe to break off completely (Cassa, 2008) . Circumferential (bending) pipe

breaks are, however, typically found in small diameter pipes (i.e. water mains), and it

normally develops around the circumference of the pipe (Cassa, 2008). This failure type

is typically caused by either the pipe breaking under bending condition or by the ground

forces pulling the pipe aside (Cassa, 2008). This failure can happen as a result of the

significant soil displacements along the pipeline (i.e. soil swell-shrink related movements

(Rajani et al., 1996). Morris (1967) and Clark (1971) reported that volume change of

clays can be considered as a significant factor towards the high number of water main

breaks. Other failure modes affecting the joints or the mechanical fitting of pipe networks

can also be induced because of the soil movement. These elements can be subjected to

seal failure or pull out due to the differential movement along the pipeline.

74

Figure 2.15: Common soil movement induced failure modes for pipe networks after (Cassa, 2008)

Soil Movement Induced Pipeline Systems' Failures

Pipe

Longitudinal Crack

Circumferential Crack

Joints

Seal Failure

Pull Out

Mechanical Fitting Valve/Meter

Seal Failure

75

2.16.2 Case Studies

Large amounts of water may enter the soil during the rainy seasons and result in

excessive soil heave. Conversely, a significant reduction in the moisture content during

the dry seasons may lead to soil settlements. The soil upward/downward deformations

were found to be induced on underground pipes to such magnitudes as to cause them

to fail, especially in the case of small diameter pipes. Newport (1981) observed that the

high breakage rates often occurred following very hot and dry summers. Ground

movements induce bending stresses on underground pipelines.

Gould (2011) investigated the effect of the seasonal variations in weather

conditions on the failure rates of Australian water reticulation pipes. The seasonal

variation in pipe failure numbers was found to occur due to excessive ground

movements caused by the shrinking and swelling of clay soils. Gould et al. (2009) also

concluded that the highest failure rates occurred between December and May. It was

clear that the failure rates were corresponding to particular climate events. Clayton et al.

(2010) analyzed monitoring data of pipelines installed in London clay over a two-year

period. It was reported that noticeable ground movements, in the order of 3 to 6 mm/m

pipe length, were observed in the vertical and horizontal directions.

Hu and Hubble (2007) analyzed the breakage rates of asbestos cement (AC)

water mains for the period from 1980 to 2004 in the City of Regina, Canada. The study

investigated the rainfall deficit and the freezing index and compared them with the

annual and winter pipe failure rates. It was noted that the annual and winter breakage

rate peaks occurred at the peak rainfall deficits. Hudak et al. (1998) also studied pipe

breakages of cast iron (CI) and PVC pipes buried in soils with different shrinkage and

swelling potentials. The results showed that the highest density of pipe breaks took

place in areas with the highest plasticity indices for 150 mm and 200 mm diameter

76

pipes. There were significant numbers of pipe breaks in June and July, which were

attributed to the occurrence of short periods of rain applied on dry soil.

Mordak and Wheeler (1988) presented historical data for four AC water main

assets located in different sites in the United Kingdom. One site was characterized by a

clay soil deposit while the others had sand and gravel deposits. The AC water main,

buried in clay, was observed to have two failure peaks, corresponding with the long dry

periods experienced during two hot summers. Most of the pipe failures occurred during

the summer (dry) months. Pipe failures for the three other areas, with sand and gravel

soils, took place randomly throughout the year.

Baracos et al. (1955) studied the pipe failure rates of cast iron (CI) water pipes in

the City of Winnipeg, Canada from 1948 to 1953. The clay deposit in the area was

characterized as highly plastic with a significant swelling/shrinkage potential. The

monthly circumferential pipe failure rates formed a cyclic pattern that occurred in

September and January. A close correlation was derived between the circumferential

failure rate pattern and the monthly weather changes, including mean precipitation,

temperatures, and depth of snow cover.

2.17 Related Numerical Modeling Studies in the Literature

The use of numerical modeling allows for assessing the effect of a wide range of

variables in a timely and efficient manner. Significant improvements in the capabilities of

computer software have improved the numerical modeling techniques. A considerable

amount of numerical studies has been completed in order to understand the complex

interactions between the soil and pipe (Wijewickreme, 2012). The scale of soil geometry

is typically large, and therefore, the soil properties can be averaged and simulated as a

continuum (Chen and Baladi, 1985; Chen and Liu, 2012). The mechanical behavior of

soil can be analyzed using the theoretical framework of the continuum mechanics of

77

solids (Chen, 1990). The continuum soil-structure interaction models, in the form of the

stress-strain partial differential equations, were initially developed based on the

governing linear elastic continuum relationships. Then, a range of assumptions can be

made in order to develop the equations in a closed form. Continuum models typically

allows for the simulation of a wide range of soil parameters (Colasanti and Horvath,

2010). Continuum modeling methods have also resulted in a better understanding of

soil-pipe interaction problems (Wijewickreme, 2012).

There are two main modeling approaches in the area of soil-structure interaction of

underground pipelines, the Winkler Spring Approach, and the Finite Element Analysis.

In the Winkler Spring Approach (Winkler, 1867), the pipe is assumed to perform as a

thin strip and the soil medium is represented by spring elements (Ng, 1994). The springs

are generally mounted transverse to the pipe axis in order to simulate the load transfer

associated with soil movement acting perpendicular to the longitudinal pipe axis (Ng,

1994). Winkler s hypothesis is still being used as the main subgrade model in soil-

structure interaction applications. Significant improvements have been made to the

Winkler Spring Approach in order to reflect the different physical aspects of the soil-

structure interaction.

Rajani et al. (1996) presented a Winkler Spring Approach for jointed water mains

consisting of PVC and cast or ductile iron pipes. A sensitivity study was implemented to

categorize the key variables in the performance of buried water mains. The results

concluded that temperature changes and the soil-pipe reaction modulus had a

considerable role on the water main breaks. The Winkler Spring Approach is a

conceptual approach to model boundary and loading conditions and often fails to

simulate the soil physical behavior in a precise manner (Dutta and Roy, 2002). The

Winkler model does not consider the continuity of soil mass and assumes no soil-pipe

interaction between the locations of soil springs along the pipeline (Rajani and

78

Tesfamariam, 2004). The disadvantage of the Winkler model is that it is primarily based

on the simulation of the soil pressure with respect to absolute pipe displacement without

incorporating the influence of any rigid body movements occurred by the soil mass

(Trickey and Moore, 2007).

The Finite Element Analysis is an advanced numerical approach where the soil is

geometrically defined by nodes and represented by finite structural units. A geometric

model is solved as a mathematical model and the behavior is described by differential

equations and boundary conditions. The behavior of a flexible pipe with non-uniform soil

support was modeled by Zarghamee (1986) as a cylindrical shell installed in an elastic

medium. It was found that the internal pipe pressure did not reduce the resulting flexural

strains given that the overturn produced by the pipeline haunch supports were

insufficient (Ban, 2008; Zarghamee, 1986).

Zhan and Rajani (1997) conducted a finite element analysis to evaluate the

influence of various backfill materials and burial pipe depths on the behavior of buried

PVC and ductile pipes. The analysis demonstrated that the use of a Controlled Low

Strength Material (CLSM) as a backfill as opposed to traditional materials (i.e. sand and

clay) resulted in reduced stresses on PVC pipes under traffic loads. McGrath (1998)

conducted a study on the soil-pipe interaction during installation of flexible and rigid

pipes. It was concluded that pipe behavior is greatly influenced by installation methods

and soil characteristics (i.e. compaction and backfill characteristics). Trickey and Moore

(2007) performed a numerical analysis for buried pipes of varying stiffness and

embedment depths. It was found that the burial depth had a little impact on the peak

deformation for stiff (rigid) pipes located close to the ground surface. However, for

flexible pipe, the peak deformation decreased significantly as embedment depth

increased.

79

Barbato et al. (2010) used a linear elastic finite element modeling technique to

study the effects of geometric and mechanical parameters that characterize the soil-

structure interaction for a buried pipe located under highways. The study concluded that

the soil-pipe interaction considerably depends on the pipe material and stiffness as well

as the geometric parameters defining the pipe trench. Rajeev and Kodikara (2011)

completed a numerical study of an experimental pipe, buried in expansive soil. The pipe

was assumed to perform as a linear elastic material. The soil, however, was modeled as

a nonlinear elastic material. The study predicted the magnitude of soil movements with

the change of water flow. It was also concluded that despite the recognized influence of

expansive soils on the behavior of underground pipelines, the research effort directed at

the numerical modeling of the soil-pipe interaction behavior is limited (Rajeev and

Kodikara, 2011).

The previous technical review shows that finite element methods and techniques

can be used effectively to investigate the behavior of various types of buried pipelines.

These studies have presented a database for the behavior of buried pipes under certain

field conditions. As reported, the amount of studies on the engineering performance of

pipes under various weather conditions is quite limited. Further studies are considered

necessary in order to better understand the field performance of buried pipes as well as

the effects associated with the unsaturated soil conditions.

80

CHAPTER 3: FIELD INVESTIGATION

3.1 General

The National Research Council Centre for Sustainable Infrastructure Research

(NRC-CSIR) instrumented an underground water main (Hu and Vu, 2011). The

instrumentation program was established to monitor the performance of an underground

pipe section placed in a well-developed area in the City of Regina where frequent pipe

breakages have occurred in recent years. This research utilized the monitoring data of

the instrumented pipe and surrounding soils. The data consisted mainly of pipe

displacements, soil water content, and ground temperatures. The field measurements

were used to better understand soil-pipe interactions under field conditions, to validate

the results of the numerical analysis, and to back calculate some critical properties of the

soil-pipe system. This chapter presents the main details of the instrumented site, and

how the pipeline was instrumented in the field.

The field site was located in Emerald Park Road, in Whitmore Park area (south

Regina), Saskatchewan. The front yards of the houses and park area were covered with

grass. Mature trees were also present in the park and in front of the houses. A new PVC

water main, 0.15 m in diameter, was installed at a depth of approximately 2.9 m below

the finished ground surface. The instrumented pipe section was installed to replace an

existing AC (Asbestos Concrete) broken pipe section. A group of high quality sensors

were buried in the backfill as well as in the native soil (Regina clay) surrounding the

trench. All instruments were properly calibrated to provide high-quality data in the field

after installation (Hu and Vu, 2011). The essential soil properties and initial conditions for

the numerical model were determined through a laboratory program that was conducted

on the collected samples from the field (Hu and Vu, 2011).

81

3.2 Field Program Details

Figure 3.1 shows the site details and general layout. A pipe trench, 2.4 m wide and

8.2 m long, was excavated to replace the existing AC pipe section. The original water

main at this location was found at a depth of 2.9 m below the ground surface. It

consisted of a 4 m long Class 150 AC pipe section, 150 mm in diameter, and was

installed about 50 years ago. The installed instrumented section consisted of a 4 m long

C900 PVC water main pipe and had a nominal diameter of 150 mm. PVC pipes are the

primary pipe type currently used for water pipe section replacements in the area. PVC

pipes have relatively low stiffness, and are therefore, very reactive to soil deformations.

The end joints consisted of PVC plastic sleeves, which were fitted over the ends of the

pipe section, where two rubber rings were used between the pipe and the sleeves. The

joints were designed to be quite flexible, and could accommodate some deformations.

Tables 3.1 and 3.2 provide a summary of the pipes properties of the original AC and

PVC pipes encountered in this investigation at the field site.

Installation of soil sensors was carried out at different depths below the ground

surface for both the native and backfill soils. The field instrumentation system was

comprised of strain gauges, thermocouples, earth pressure cells, displacement

transducers, and soil moisture (WCR) probes. Figure 3.2 shows the main types of

measurements collected for the native and backfill soils and pipe. The instrumentation

program monitored the vertical displacement, and temperature of the pipeline at specific

locations along with some essentials properties of the native clay soil, mainly

temperature, soil pressure, and volumetric water content. The sensors locations were

selected to cover a decent range of measurements of the soil and pipe at different levels

(Hu and Lotfian, 2006).

82

Figure ‎3.1: Site location and layout

Figure 3.2: Summary of field instrumentation types

Volumetric

Water Content

Soil MeasurementsPipe Measurements

Field Monitoring Data

Temperature Soil PressureVertical Displacement

83

Table 3.1: Dimensions and properties of the existing AC pipe section

Property Value

Nominal diameter (mm) 150

Outer diameter (mm) 183

Inside diameter (mm) 153

Length (m) 4

Elastic modulus (GPa) 20 – 25

Ultimate tensile strength (MPa) 25

Strain to failure (%) 0.1

Poisson’s ratio 0.3

Thermal coefficient (× 10–6/°C) 8.5

Table 3.2: Dimensions and properties of the Instrumented PVC pipe

Property Value

Nominal diameter (mm) 150

Outer diameter (mm) 175

Inside diameter (mm) 155

Length (m) 4

Elastic modulus (GPa) 2.8

Ultimate tensile strength (MPa) 48

Strain to failure (%) 10

Poisson’s ratio 0.42

Thermal coefficient (× 10–6/°C) 79

84

3.3 Instrumentation Details

Table 3.3 presents some of the necessary details of the sensors installed at the

site. Figure 3.3 presents a section layout showing the locations of the sensors that were

installed on site. A total of thirty (30) volumetric water content sensors (WCR) were

installed in the backfill and surrounding native soil to observe the moisture content

profiles. The WCR probes were calibrated for the native soil conditions (Hu et al., 2008).

Four pressure cells were placed in the vertical and horizontal directions to measure both

vertical and horizontal soil pressures. The pressure cells with a 350 kPa capacity were

installed in the trench backfill at a depth of about 170 mm above the top of the pipe.

Calibration curves for the pressure cells were provided by the manufacturer (Hu and

Lotfian, 2006). In addition, twelve (12) thermocouples were installed on the pipe exterior

surface; and thirty three (33) thermocouples were placed in the backfill and the native

soil surrounding the trench.

Three custom-built Geokon Model 4450 vibrating wire displacement transducers

were installed to measure the pipe displacements with time. The rod extensometers

were installed with one end of each extensometer being attached to the instrumented

pipe section, and the other end was installed in a test hole and anchored to the ground

at a depth of about 6.6 m below the ground surface. The test holes were backfilled with

bentonite plug to prevent water flow into the test holes. An automatic data acquisition

system was installed to collect and monitor pipe deformation, volumetric water content,

earth pressure, pipe displacements, and soil temperatures. The type of data acquisition

system used was manufactured by Campbell Scientific and the model was CR1000-55.

The data-logging equipment was kept in a metal instrumentation box in the field, and the

collection frequency was every hour.

85

Table 3.3: Summary of the main details of the instruments installed at the field site

Property Instrument Type Range (unit)

Accuracy (unit)

Pipe vertical displacement

Displacement (Custom-built 4450) (Geokon)

0 to 300 (mm) ± 0.1 (%)

Temperature Thermocouples (T type) (Veriteq Instruments)

-190 to + 350 (oC) ± 0.40 (°C)

Soil pressure Pressure cell (Model 4810X) (Geokon)

0 to 350 (kPa) ± 0.1 (%)

Soil volumetric water content

Water content reflectometer (CS616)(Campbell Scientific)

0 to 100 (%) ± 2.5 (%)

Figure 3.3: Schematic of the installed sensors layout at the field site

86

3.4 Climate Data

The climate data for the study area was retrieved from a nearby Environment

Canada meteorological station (Regina International Airport, Regina, Saskatchewan).

Figures 3.4 to 3.8 show the climate data including daily precipitation, air temperature,

wind speed, net radiation, and relative humidity covering a range from November 2005

to November 2007. The presented climate data was used as input for the numerical

modeling program. The total precipitation accounted for rainfall (primarily during the

summer) and snow (primarily during the winter). It is also to be noted that the historical

temperature data indicated that the monthly mean temperature is below 0 °C for 5

months from November to March every year.

87

Figure 3.4: Daily and cumulative precipitation at the field site

Figure 3.5: Daily air temperature at the field site

Nov-05 Feb-06 May-06 Aug-06 Nov-06 Feb-07 May-07 Aug-07 Nov-07

Date (mmm-yy)

0

10

20

30

40

Pre

cip

itation (

mm

)

0

200

400

600

800

1000

Cum

ula

tive P

recip

itation (

mm

)

Precipitation

Cumulative Precipitation

Nov-05 Feb-06 May-06 Aug-06 Nov-06 Feb-07 May-07 Aug-07 Nov-07

Date (mmm-yy)

-40

-20

0

20

40

Te

mpe

ratu

re (

oC

)

88

Figure 3.6: Daily wind speed at the field site

Figure 3.7: Daily net radiation at the field site

Dec-05 Mar-06 Jun-06 Sep-06 Dec-06 Mar-07 Jun-07 Sep-07

Date (mmm-yy)

0

20

40

60

Win

d S

pe

ed

(km

/hr)

Nov-05 Feb-06 May-06 Aug-06 Nov-06 Feb-07 May-07 Aug-07 Nov-07

Date (mmm-yy)

0

5

10

15

20

25

Ne

t R

ad

iation

(M

J/m

2/d

ay)

89

Figure 3.8: Daily relative humidity at the field site

Nov-05 Feb-06 May-06 Aug-06 Nov-06 Feb-07 May-07 Aug-07 Nov-07

Date (mmm-yy)

20

40

60

80

100

Re

lative

Hu

mid

ity (

%)

90

3.5 Backfill and Bedding Soils

The backfill and bedding soils in the trench consisted of mixed concrete (MC) and

sand, respectively. The grain size distribution curves for the bedding sand and mixed

concrete are shown in Figure 3.9. Clean sand was used as a bedding material below the

pipe. One sand lift was placed and compacted from the bottom of the trench to the

spring-line of the pipe (approximately 0.36 m deep). The trench was backfilled with MC

from the spring-line of the pipe up to the pavement base level.

The mixed concrete (MC) consisted of a blend of recycled crushed concrete and

sub-base gravel at a ratio of 2:1 by weight (Hu and Lotfian, 2006). The MC was installed,

in lifts of 150 mm, to the level of the underside of the pavement according to the

Construction Specifications of the City of Regina (Hu and Vu, 2011). The trench backfill

was compacted using a vibrating plate compactor. The water content and dry density

were measured in the field using a nuclear densometer.

91

Figure 3.9: Grain size distribution for backfill materials (sand and mixed concrete)

0.01 0.1 1 10 100

Particle Size (mm)

0

20

40

60

80

100

Perc

en

tag

e F

ine

r (%

)

Mixed Concrete

Sand

92

3.6 Soil Profile and Properties at the Field Site

Soils encountered in the study area were identified by two test holes before the

installation of the instrumented pipe section. The first test hole (TH1) was drilled in the

street, about 1.7 m west of the pipe. The second (TH2) was in the park, about 6 m east

of the pipe. Soils generally consisted of lacustrine clay which extended to a depth of

about 9.5. Clay till layer was encountered below the clay and extended to the bottom of

the test holes. Laboratory testing was reported for selected soil samples, including index

tests, soil-water characteristic curves, saturated hydraulic conductivity tests, swelling

tests, consolidation tests, and measurement of soil suction (Hu and Vu, 2011).

The upper soil layer was identified as highly plastic lacustrine clay (Regina clay)

that was typically silty, moist and stiff to very stiff. Figure 3.10 shows the water content

and dry unit weight profiles at the field site. The clay was also found to be over-

consolidated with an over-consolidation ratio of approximately 5.0. The measured

swelling pressure ranged from 500 to 550 kPa. The clay was also identified to exhibit

significant swelling and shrinkage characteristics upon wetting and drying.

The grain size distribution curve for both the top Regina clay and clay till samples

were plotted in Figure 3.11. Tables 3.4 and 3.5 show a summary of the key geotechnical

index properties for Regina clay and clay till soils. The grain size distribution curve

shows that the Regina clay consists of silt and clay maximum content (size less than 74

m) in the range of 98.7% - 99.9%. The liquid limit (LL), plastic limit (PL), and plasticity

index (PI) were measured to be in the range of (64 % - 94%), (23% - 35%), and (37% -

66%), respectively. The measured water content for the field samples ranged from 23%

to 35%. The soil specific gravity (Gs), average void ratio (e), and average dry unit weight

(yd) were found to be 2.73, 0.95, and 1540 kg/m3, respectively. The field matric suction

was found to be in the range of 700 kPa to 3000 kPa, and the approximated swelling

index (Cs) was measured to be approximately 0.09.

93

In general, the clay is considered non-plastic when it has a moisture content that is

less than the plastic limit (White, 1949). Therefore, the clay would likely be cracked or it

has a tendency to develop cracks with any reduction in its moisture content. Based on

the soil moisture content results, the upper two to three meters of the clay were found to

have a natural moisture content that was less than that of the lower clay layer. In

addition, the upper clay layer was also found to have an average moisture content that

was less than the plastic limit demonstrating that the clay was desiccated. (Hu and Vu,

2011) also illustrated that large surficial cracks, approximately 10 mm wide, were

observed in the study area.

94

Figure 3.10: Water content and dry unit weight profiles with depth at the field site

10 15 20 25 30 35 40

Water Content (%)

16

12

8

4

0

Dep

th (

m)

TH1

TH2

Clay

Clay Till

13 14 15 16 17 18 19

Dry Unit Weight (kN/m3)

16

12

8

4

0

Dep

th (

m)

TH1

TH2

Clay

Clay Till

95

Figure 3.11: Grain size distribution of Regina clay and clay till

0.001 0.01 0.1 1 10

Particle Size (mm)

20

40

60

80

100

Perc

en

tag

e F

ine

r (%

)

Regina Clay

Clay Till

Silt and clay content (74m)

- For Regina Clay = 99.9 (%)- For Clay Till = 95.8 (%)

96

Table 3.4: Geotechnical index properties of Regina clay

Soil Property (unit) Value

Specific gravity 2.73

Average dry density, yd (kg/m3) 1540

Moist unit weight, yw (kN/m3) 19.87

Natural water content, w (%) 23 – 35

Liquid limit, LL (%) 64 – 94

Plastic limit, PL (%) 23 – 34

Plastic index, PI (%) 37 – 66

Void ratio, e 0.95

Poisson’s ratio, (predicted) 0.33

Field suction-filter paper (kPa) 700 – 3000

Saturated hydraulic conductivity (m/s) 1.6 × 10–9 to 2.8 × 10–8

swelling index, Cs 0.09

Corrected swelling pressure (kPa) 500 – 550

Gravel content (≤62 mm) (%) 0.0 – 0.4

Sand content (≤5 mm) (%) 0.1 – 1.3

Silt and clay content (≤74 mm) (%) 98.7 – 99.9

97

Table 3.5: Geotechnical index properties of glacial clay till

Soil Property (unit) Value

Dry density, yd (kg/m3) 1900

Moist unit weight, yw (kN/m3) 21.2

Natural water content, w (%) 12 – 20

Liquid limit, LL (%) 38 – 44

Plastic limit, PL (%) 14 – 20

Plastic index, PI (%) 23 – 28

Field suction-filter paper (kPa) 2000

Gravel content (≤62 mm) (%) 0 – 3.6

Sand content (≤5 mm) (%) 4.2 – 31.8

Silt and clay content (≤74 mm) (%) 66.5 – 95.8

98

3.7 Unsaturated Soil Characteristics

Figure 3.12 also shows the soil-water characteristic curve (SWCC) of Regina clay

predicted using the laboratory measurements of soil suctions based on Fredlund and

Xing Fitting equation (Fredlund and Xing, 1994). The development of the SWCC for the

naturally deposited clay soil is discussed later in the thesis. The SWCC for the clay till

and backfill materials (sand and mixed concrete) were estimated based on the index

properties and the grain size distributions using the methods documented by (Fredlund

and Xing, 1994; Fredlund et al., 2002). Figures 3.13 and 3.14 show the soil-water

characteristic curve (SWCC) for clay till, sand and mixed concrete.

The measured saturated hydraulic conductivity of the clay ranged from 1.6 X 10-9

m/sec to 2.8 X 10-8 m/sec. The saturated hydraulic conductivity of the clay till, sand and

mixed concrete were estimated based on the grain size distribution, using Hazen’s

equation (Hazen, 1892), to be 1.0 X 10-8 m/sec, 2.30 X 10-05 m/sec and 7.50 X 10-06

m/sec, respectively. The unsaturated hydraulic conductivity functions for the native clay,

clay till and the backfill materials were estimated, using the Leong and Rahardjo

equation (Leong and Rahardjo, 1997), and based on the saturated hydraulic

conductivity. Hydraulic conductivity functions were plotted for Regina clay, clay till, sand,

and mixed concrete in Figures 3.15 and 3.16.

99

Figure ‎3.12: Soil water characteristic curve (SWCC) for Regina Clay

Figure 3.13: Soil water characteristic curve (SWCC) for clay till

0.1 1 10 100 1000 10000 100000 1000000

Matric Suction, (kPa)

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

Vo

lum

etr

ic W

ate

r C

on

ten

t, (

vo

l/vo

l)

Fredlund and Xing SWCC Fit

Laboratory Measurements

Saturation Suction = 0.1 kPa

Residual Volumetric Water Content, s= 0.12

Air Entry Value, AEV103.5 kPa

1 10 100 1000 10000 100000 1000000

Matric Suction, (kPa)

0

0.1

0.2

0.3

0.4

0.5

Vo

lum

etr

ic W

ate

r C

on

ten

t, (

vo

l/vo

l)

Fredlund and Xing SWCC Fit

Saturation Suction= 1 kPa

Air Entry Value, AEV63.4 kPa

Residual Volumetric Water Content, s= 0.19

100

Figure 3.14: Soil water characteristic curves (SWCCs) for the backfill materials

Figure 3.15: Hydraulic conductivity functions for Regina clay and clay till

0.1 1 10 100 1000 10000 100000 1000000

Matric Suction, (kPa)

0

0.1

0.2

0.3

0.4

Vo

lum

etr

ic W

ate

r C

on

ten

t, (

vo

l/vo

l)

Sand

Mixed Concrete

Saturation Suction = 0.1 kPa

Residual Volumetric Water Content, s

- For Sand = 0.05

- For Mixed Concrete = 0.17

Air Entry Value, AEV

- For Sand = 5 kPa

- For Mixed Concrete = 0.92 kPa

Fredlund and Xing SWCC Fit

0.1 1 10 100 1000 10000 100000 1000000

Matric Suction, (kPa)

1E-012

1E-011

1E-010

1E-009

1E-008

1E-007

Hyd

rau

lic C

on

du

ctivity,

k (

m/s

ec)

Regina Clay

Clay Till

101

Figure 3.16: Hydraulic conductivity functions for the backfill materials

0.1 1 10 100 1000 10000 100000 1000000

Matric Suction, (kPa)

1E-008

1E-007

1E-006

1E-005

0.0001

Hydra

ulic

Co

nd

uctivity,

k (

m/s

ec)

Sand

Mixed Concrete

102

3.8 Soil Moisture Field Data

Figure 3.17 shows the volumetric water content (VWC) measurements at three

levels 0.45 m, 2.92 m, and 4.0 m at the field site in the clay deposits surrounding the

pipe trench. As a general trend, for clays at low levels (2.92 m, and 4.0 m), had

experienced relatively small variations in volumetric water contents which were found to

be up to 5 %. The variation of volumetric water contents corresponded to the seasonal

variation in the climate conditions, without significant transits during rainfall or snowmelt

events. The clay soil at a higher level had noticeable variations in volumetric water

content which typically corresponded to the seasonal variation in the climate conditions

and the rainfall or snowmelt events. The volumetric water content at a depth of 0.45 m

was found to be as low as 15 % to as high as 36 %. The volumetric water content (VWC)

measurements were used to predict the variation in soil suction using the soil water

characteristic curve (SWCC) as showed in Figure 3.18. The pattern of the variations of

volumetric water contents was observed for the soil suction. The maximum and

minimum soil suction for the clay deposit at the two levels 2.92 m and 4.0 m were

estimated to be in the order of 1150 kPa and 480 kPa; however they were found to be in

the order of 36500 kPa and 2760 kPa at a depth of 0.45 m.

103

Figure 3.17: Volumetric water content in the clay deposit at various levels

Figure 3.18: Estimated soil suction in the clay deposit at various levels

Apr-06 Jun-06 Jul-06 Sep-06 Oct-06 Dec-06 Jan-07 Mar-07 Apr-07

Date (mmm-yy)

0

10

20

30

40

50

60

Vo

lum

etic W

ate

r C

on

ten

t, (

%)

0

10

20

30

40

Da

ily p

recip

itatio

n (

mm

)

Precipitation

Depth=0.45 m

Depth=2.92 m

Depth=4.0 m

Apr-06 Jun-06 Jul-06 Sep-06 Oct-06 Dec-06 Jan-07 Mar-07 Apr-07

Date (mm/dd/yy)

100

1000

10000

100000

Estim

ate

d S

uction

, (

kP

a)

0

5

10

15

20

25

Da

ily p

recip

ita

tio

n (

mm

)

Precipitation

Depth=0.45 m

Depth= 4.0 m

Depth=2.92 m

104

3.9 Air and Soil Temperatures

Figure 3.19 shows the air and ground temperatures collected in a 1-year period

from April 2006 to April 2007 following the pipe installation. The soil temperatures

experienced maximum and minimum peak temperatures corresponding to the air

temperature. The peak temperatures occurred at different time durations contingent on

the depth below the ground surface. In general, the shallower the depth, the earlier the

peak temperature occurred. At a depth of 0.45 m, the peak temperatures occurred at a

time that was close to the peak air temperature. However, there was a delay for the

maximum and minimum peak temperatures to occur at the pipe level. Detailed

discussions concerning the soil temperature with time are included later in the thesis.

3.10 Measured Pipe Displacements

The instrumented pipe displacement varied significantly throughout the

monitoring period. Figures 3.20 and 3.21 show the variation in pipe displacement with

time. The pipe experienced significant displacement peaks, in a 12-month period, a

relatively small one in the winter season and large ones in the spring/summer season.

Pipe displacements were also found to be different along the pipeline which were

attributable to the performance of the pipe end restraints. Overall, the increase in soil

pressure resulted in downward pipe displacements. It is anticipated that the decrease in

soil moisture content caused the soil to shrink, which in return, reduced the soil friction.

The soil pressure borne by the pipe would normally increase as a result of any reduction

in soil friction (Moser, 1990). Upward pipe displacements were also observed and were

closely correlated to seasonal variations in weather conditions.

105

Figure 3.19: Air and soil temperature observed at the field site

Nov-05 Feb-06 May-06 Aug-06 Nov-06 Feb-07 May-07 Aug-07 Nov-07

Date (mmm-yy)

-40

-20

0

20

40

Te

mpe

ratu

re (

OC

)

Depth = 0.45m

Depth = 1.0m

Air

Depth = 2.92m

Depth = 4.0m

106

Figure 3.20: Pipe displacements and soil pressures at the field site

Figure 3.21: Pipe displacements at the field site

-20

40

100

160

220

280

-5

0

5

10

Apr-06 Jul-06 Oct-06 Jan-07 Apr-07

So

il P

ressure

(kP

a)

Pip

e D

isp

lacem

ent (m

m)

Date (mmm-yy)

Vertical Pressure

Horizontal Pressure

Maximum Displacement

Average Displacement

0

10

20

30

40

-5

0

5

10

Apr-06 Jul-06 Oct-06 Jan-07 Apr-07

Daily

Pre

cip

itatio

n (m

m)

Pip

e D

isp

lacem

ent (m

m)

Date (mmm-yy)

Precipitation

Maximum Displacement

Average Displacement

107

CHAPTER 4: LOAD-DEFORMATION ANALYSIS UNDER DIFFERENT SOIL

AND PIPE CONDITIONS

4.1 Problem Statement

A technically sound design of buried pipes has to be based on the pipe

characteristics, internal and external loads, and surrounding conditions such as backfill

and side fill materials, and installation depth. The main objective of this study was to

investigate the influence of the most significant design parameters of buried pipes. A

numerical program was established and used to estimate the elastic deformations of a

soil-pipe system under different soil and loading conditions through real-life scenarios.

The finite element modeling approach was used to account for volume change effects,

soil layering, and displacement of the pipe relative to the soil. The model was validated

using previously developed empirical design equations. The numerical analysis was

performed in order to predict the elastic deformations occurring due to various backfill

materials around the pipes, native soils around the trench, and diverse applied loading

magnitudes.

4.2 Governing Equations

The soil-pipe system was modeled as an elastic continuum. The elastic continuum

solution for buried pipes can be represented by Equation (4.1) as suggested by (Klar et

al., 2004).

[PS](u) = (F) (4.1)

where;

[PS] is the stiffness matrix of the pipe,

(u) is the pipe displacement, and

(F) is the force vector representing the soil loading.

108

Generally, the total stress distribution in a soil element is controlled by the static

equilibrium of applied forces. The balance of forces would result in a set of partial

differential equations’ governing static equilibrium of forces (Chou and Pagano, 1992).

The governing stress-strain equation, in Cartesian coordinates, is presented in terms of

displacements in the x and y directions (u and v) as shown in Equations (4.2) to (4.9).

σ

σ

(4.2)

σ

σ

(4.3)

σ

(4.4)

σ

(4.5)

σ

(4.6)

(4.7)

(4.8)

(4.9)

where;

E is elasticity parameter,

σx and σy are the stresses in the x- and y- direction, and

σxy is the shear stress, Fx and Fy are the body forces in the x- and y- direction and

are equal to 0 and –gρ, respectively.

Combining the Equations (4.2) and (4.3) would result in Equations (4.10) and (4.11).

(4.10)

(4.11)

109

4.3 Modeling Overview, Geometry and Boundary Conditions

A two-dimensional soil-pipe model was developed to perform load-deformation

analysis as summarized in Figure 4.1. The analysis was conducted for a typical

underground pipe system that was under different soil and loading conditions. Figure 4.2

shows the model geometry and dimensions. The soil-pipe system was modeled under

strain conditions. The material used to fill the pipe trench was subdivided into three

different layers bedding, backfill, and cover. The natural soil surrounding the trench, the

soil in the trench, and the road pavement structures were modeled as separate layers. A

parametric study was initially conducted to evaluate the most significant design

parameters as reported in (Saadeldin et al., 2013b). The main features of the model can

be summarized as follows: (i) the soil and pipe materials were modeled in accordance

with the elastic continuum approach; (ii) live loads due to vehicular traffic were modeled

as a concentrated static load magnified by a dynamic amplification factor, and (iii) the

effect of variable water elevation was not modeled.

The appropriate boundary conditions were applied along the borders of the finite

element model. The bottom of the model was defined as a fixed boundary, where the

horizontal and vertical displacements were set equal to zero. However, the vertical sides

of the model had only a horizontal displacement equal to zero. Loads specified for the

analysis include the in-situ soil pressure in the soil profile and the assumed traffic loads.

Figure 4.3 presents the geometry and generated mesh of the model. The main input

parameters of the sensitivity analysis for the load-deformation analysis are summarized

in Table (4.1). The pipe was modeled as a 2D plane wall pipe with constant wall

thickness (t). The pipe used was a PVC pipe with (D/t) and initial Young’s modulus (Ep)

of 2.8 GPa. The road was modeled by a sub-pavement base of thickness equal to 0.25

m, and an asphalt layer of 0.1 m thickness.

110

Figure 4.1: Summary of soil-pipe interaction model for parametric study analysis

Figure 4.2: Model geometry and definition of the problem

111

Figure 4.3: Generated mesh

(X and Y units are in meters)

112

Table 4.1: Summary of the main input parameters

Parameter (unit) Value

Pipe diameter to thickness ratio, D/t 10

Soil bedding thickness, hp (m) 0.15

Pavement thickness, tp (m) 0.1

Sub-base thickness, tbs (m) 0.25

Poisson’s ratio of pipe, p 0.4

Poisson’s ratio of clay, μs 0.3

Modulus of elasticity of clay, Es (kPa) 9500 - 47500

Modulus of elasticity of pipe, Ep (GPa) 2.8 - 280

Pipe diameter, D (m) 0.15 - 0.6

Cover thickness, hc (m) 0.3 - 2.0

Traffic load, T.L. (kN) 0.0 - 100

113

4.4 Numerical Modeling versus Analytical Results

Some of the model results, in the case of no applied surface traffic load and no

pavement structure, were compared with the results obtained by using Spangler’s

equation, illustrated earlier in this thesis. The soil load (wc) in Spangler’s equation was

obtained using the Marston load theory. In addition, as defined by Moser (1990), the

load coefficients (Cd) were determined to be 1.55 and 0.84 for a trench width of two and

six times the pipe diameter, respectively. It was observed that the results of the

numerical analysis matched the analytical results obtained by Spangler’s equation for

wider trench configuration. These findings and similarities led to higher deformations for

trenches of less width (i.e. twice the pipe diameter), as shown in Figure 4.4. It is

important to note that Spangler’s equation uses an individual value of the soil elastic

modulus to represent the backfill soil. However, the numerical simulation is capable of

fully considering the variation in elastic properties between the backfill soil and the native

soil surrounding the pipe trench.

114

Figure 4.4: Effect of the trench width on the maximum pipe deformations

10000 20000 30000 40000 50000

Modulus of Elasticity, Ef (kPa)

0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

Pip

e D

efo

rmation,

D/D

(%

)

Native soil condition, Es=9500 kPaTrench

W

D

Ef

Es

12D

Spangler's equation

Modeling results

W= 2D

W= 6D

W= 6D

W= 2D

115

4.5 Pipe Deformations

The study analyzed the soil-structure interaction system under different soil and

pipe conditions. The field behavior of pipelines is significantly influenced by the trench

backfill, and the natural soil surrounding the trench. The surrounding backfill material

provides considerable support for underground flexible pipelines. It is well known that the

narrower the trench, the lighter the load applied to the pipe, and the pipe has to support

this load. However, in the case of flexible pipes, the pipe tends to rely heavily on the

surrounding soil to carry the applied loads. A flexible pipe principally derives its

resistance strength from the passive pressures induced by the relative movement of the

sides of the pipe against the surrounding soil (Moser, 1990).

Figure 4.5 shows the influence of soil elastic modulus on the surface displacement

under different surface traffic loads (T.L.) (i.e. 0 kN/m, 10 kN/m, 25 kN/m, 50 kN/m, 75

kN/m, and 100 kN/m). The increase in the soil elasticity, as an indicator of the soil

strength, resulted in reduced surface displacements and pipe deformations. It is a

practical and more efficient way of reducing the influence of earth loads on the pipe. The

higher strength of backfill reaction can be obtained by applying efficient compaction

conditions and using granular soils.

Figure 4.6 shows the influence of the soil cover thickness (hc) on pipe deformation

occurring under different surface traffic load magnitudes (i.e., 0 kN, 10 kN, 25 kN, 50 kN,

75 kN, and 100 kN). As a general trend, higher installation depths increased the effect of

soil load and decreased the impact of the surface load on underground pipelines.

However, pipe deformations were found not to increase monotonically with the increase

in soil cover under high surface loads.

116

Figure 4.5: Effect of soil modulus of elasticity on the overall soil surface displacement

Figure 4.6: Effect of soil cover height and loading conditions on pipe deformation

T.L.= 100 kN/m

T.L.= 75 kN/m

T.L.= 50 kN/m

T.L.= 25 kN/m

T.L.= 10 kN/m

10000 20000 30000 40000 50000

Soil Modulus of Elasticity, Es (kPa)

2

4

6

8

10

12

Su

rfa

ce

Dis

pla

ce

me

nt

(mm

)

Initial Condition, Es=9500 kPa

Ef/ Es=5

Ef/ Es=3

Ef/ Es=2

Ef/ Es=1.5

Ef/ Es=1.25

0 0.4 0.8 1.2 1.6 2

Soil Cover Height, hc (m)

0.1

0.2

0.3

0.4

0.5

0.6

0.7

Pip

e D

efo

rmation,

D/D

(%

)

Initial Condition, Es=9500 kPa

0 kN

10 kN

25 kN

50 kN

75 kN

100 kN

117

The performance of buried pipes was analyzed under different backfill material

conditions. As a general trend, higher backfill elasticity properties were found to make a

significant contribution to the behavior of flexible pipes. This response is a result of the

better confinement provided to the sides of the pipe with filling material that is stiffer than

the natural soil surrounding the trench. The performance of buried pipes was analyzed

under different backfill material conditions. Figure 4.7 shows the maximum pipe

deformations ( D/D) due to the increase in soil elasticity ratio between the backfill soil in

the trench and the native soil (Ef/Es). Two main cases were studied, including, trenches

with a width of two and six times the pipe diameter. When the backfill soil strength was

less than the native soil (Ef/Es < 1), the rate of increase in pipe deformation for the wide

trench was much higher than in the case of the narrow trench. Conversely, when the

backfill soil strength is greater than the native soil (Ef/Es > 1), the pipe deformation with a

wide trench was lower than in the case of a narrow trench.

In the event of a trench width of two times the pipe diameter, the resulting

reduction in the maximum pipe deformation was up to about 35% with an increase in soil

elasticity factor (Ef/Es) of 5. However, increasing (Ef/Es) to 5 in a trench width of six times

the pipe diameter resulted in a reduction in the pipe deformation up to about 60%.

Higher backfill strengths were found to make a considerable contribution to the behavior

of flexible pipes. This response is a result of the better confinement provided to the sides

of the pipe with filling material that is stiffer than the natural soil surrounding the trench.

These results confirm that flexible pipes mainly derive their ability to resist loads from the

lateral pressure of the soil along the sides of the pipes.

Pipe deformation increased with the decrease in pipe diameter and elastic

modulus as shown in Figure 4.8. It was generally observed that PVC pipes are subjected

to higher pipe deformations than rigid pipes due to their lower elastic modulus. The rate

118

of increase in pipe deformation with the decrease in pipe diameter was also found to be

different between 0.6 m to 0.3 m and 0.3 m to 0.15 m.

Field installation details and design criterion can be developed to determine the

optimized parameters as a function of the pipe characteristics (i.e., strength and

diameter), natural and backfill soil conditions, and surface loading conditions. Based on

the modeling results, it was observed that the pipe shares load with the soil beside it.

Loads are transferred through the compacted backfill material and the undisturbed

trench walls. Therefore, the trench width plays a role in controlling the net applied loads

on the pipe. It is essential to minimize the effect of trench width on the load reaching the

pipe. This cannot be achieved if the backfill material is not compacted. Therefore, it must

be recognized that adequate working space would then need to be provided. If working

space is insufficient, proper compaction for the backfill material beside and beneath the

pipe may not be achievable.

119

Figure 4.7: Effect of the backfill modulus of elasticity on pipe deformations

Figure 4.8: Effect of the soil cover thickness and loading conditions on pipe deformations

0 1 2 3 4 5

Modulus of Elasticity Ratio, Ef/Es

0

0.1

0.2

0.3

0.4

0.5

Pip

e D

efo

rmation,

D/D

(%

)

W= 2D

W = 6D

Native soil condition, Es=9500 kPa WTrench

D

Ef

Es

12D

0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6

Pipe Diameter, D (m)

0.2

0.3

0.4

0.5

0.6

0.7

Pip

e D

efo

rmation,

D/D

(%

)

Ep = 2.8 GPa

Ep = 28 GPa

Ep = 280 GPa

120

CHAPTER 5: PIPE RESPONSE TO RELATIVE SATURATION OF

SURROUNDING SOIL

5.1 Problem Statement

A large amount of water may enter soils during rainy seasons and result in

excessive soil heave whereas a significant reduction in water content during dry seasons

may result in settlement of soils as shown in Figure 5.1. Unsaturated soil surrounding

the pipes may, in return, swell and shrink which may change the loading on the pipes.

Pipeline design guidelines, such as those proposed by (Nyman, 1984), are typically

based on the assumption that the foundation soil is either fully saturated or dry. Until

now, the behavior of unsaturated soils on pipelines is not well understood. The modeling

technique adopted in Chapter 4 was modified to address the performance of a

hypothetical pipeline buried in an unsaturated clay soil.

The main objective of this study was to investigate the performance of small pipes

buried in unsaturated soil. The modeling analysis aimed to capture the pipe

displacements that occurred due to the change in soil suction associated with changes

in the soil moisture content. The soil suction was estimated based on field

measurements and was then used as an input for the model. The soil upward movement

was then simulated in order to draw conclusions regarding the behavior of pipes under

saturation of surrounding soils.

5.2 Governing Equations

The constitutive relationships for the modeling of unsaturated soil conditions were

developed as an extension of the saturated soil constitutive equations and utilized two

independent measures consisting of the total normal stress and soil suction. Soil

deformations occurred due to the change in total soil volume were then defined as the

summation of the normal strains in x- and y- directions (Fredlund and Morgenstern,

121

1976). The incremental elastic forms of these equations were provided in Equations

(5.1) to (5.8) (Fredlund and Vu, 2003). The main coefficients of volume change were

calculated as a function of the soil matric suction and net normal stress. The governing

stress-strain equations can be presented in terms of displacements in the x- and y-

directions (u and v). The stress-strain relation was assumed to be linear within each

stress and strain increment; however, the elasticity parameters, E and H, were allowed

to change in magnitude between increments.

(5.1)

(5.2)

(5.3)

(5.4)

(5.5)

(5.6)

(5.7)

(5.8)

where;

μs is the Poisson’s ratio for the soil,

Fx and Fy are x and y components of the body force vector,

E is the elasticity parameter as a function of the net normal stress,

H is the elasticity parameter as a function of the change in matric suction,

Cs and Cm are swelling indices obtained from net normal stress plane and matric

suction plane, respectively,

122

Ce, C11, C22, and C33 are the stiffness tensor components as defined by (Fredlund

and Gitirana Jr, 2005),

dσ is the net normal stress, and

d is the change in matric suction.

5.3 Modeling Overview, Geometry and Boundary Conditions

The proposed model was developed to solve the soil-structure interaction

equations under unsaturated soil conditions (Saadeldin et al., 2013b). Figure 5.2

describes the soil-displacement modeling approach for the analysis. Figure 5.3 presents

the geometry and boundary conditions used for the two-dimensional analysis. The soil

mass was divided into two main zones. The first zone (inactive zone) is covered with

pavement, which significantly reduces the infiltration of precipitation into the zone,

whereas the second zone (active zone) is directly subjected to various weather

conditions. The change in volumetric water content in the area studied was found to be

as low as 5%, as high as 20%, and corresponded well with the seasonal variations in

climate conditions.

Changes in soil suction subsequently occurred within the active soil depth. The

SWCC was estimated according to the unimodal Fredlund and Xing Fitting equation and

based on index properties, including water content (w), specific gravity (Gs), and dry unit

weight ( d). Soil suction profiles were predicted using the SWCC for the defined initial

and final soil moisture conditions. The soil under the pavement was assumed to have a

negligible matric suction change, a result of limited vertical moisture flows. Free

movement in the vertical direction was allowed at the top boundary and horizontal

movements of both the left and right sides were fixed. The lower boundary was fixed in

both directions. The boundary conditions of the pipelines were defined independently of

the side boundaries of the model to represent different end restraints.

123

The pipeline had a nominal diameter of 0.15 m, a total length of 6.5 m, and a

length of 4.0 m in the active zone. The pipe depth within the active zone was presented

by a relative depth ratio (DR) which was equal to the pipe depth divided by the total

active depth of the soil. The pipe elastic modulus and ultimate tensile strength were 2.8

GPa and 48 MPa, respectively. Figure 5.4 presents the adopted mathematical model of

the soil-pipe interaction. The field natural volumetric water content (θ) of Regina clay

was found to be around 40%. The field soil suction was measured to be around 2000

kPa. When water enters the soil, the soil suction gradually decreases. To account for

this decrease in soil suction for modeling purposes, a set of soil suction values (i.e. 1000

kPa, 500 kPa, 180 kPa, and 38 kPa) were modeled. This corresponds to an approximate

increase in volumetric water content (∆θ) of 5 %, 10%, 15%, and 20%. The soil suction

variation in the active zone was assumed to extend to the whole active depth. Some of

the main input parameters of the soil-pipe model are summarized in Table 5.1. The

modeling results were also obtained for a range of pipe conditions (i.e., elastic modulus,

depth, and end restraints).

124

Figure 5.1: Schematic diagram of the effect of dry and wet soil conditions on the pipe performance after (Rajeev et al., 2012)

Figure 5.2: Summary of the modeling procedure for displacement analysis

Input Processing

Material Index

PropertiesEstimated SWCC

Volume Change

IndicesInitial Suction State

Poisson's Ratio VWC Change, θ

Boundary Conditions Final Suction

Elasticity Modulus

Two Dimensional Soil-pipe Interaction Model

Displacement Data

125

Figure 5.3: Geometry for the two-dimensional soil-pipeline model

Figure 5.4: Theoretical model for the analysis of a buried pipe under unsaturated soil conditions

126

Table 5.1: Initial material parameters of the PVC pipe

Parameter Value

Nominal diameter, D (mm) 150

Total pipe length, L 6.5

Pipe length in active zone , X 4.0

Poisson’s ratio, µp 0.4

Elastic modulus, Es (GPa) 2.8

Ultimate tensile strength (MPa) 48

127

5.4 Numerical Modeling versus Analytical Results

The results of the cumulative swelling movement occurring at the ground surface

were validated against the results determined by Equation (5.9) suggested by (Fredlund

and Rahardjo, 1993a).

(5.9)

where;

Sh is the surface heave (m),

Pf is the final stress state in the soil layer (kPa),

Po is the initial stress state in the soil layer (kPa),

hl is the layer thickness (m),

Cs is the swelling index, and

e0 is the initial void ratio.

The initial state of stress was defined as the measured swelling pressure which is a

function of the soil depth. The soil profile was divided into fifteen layers. The average

swelling pressure of Regina clay was found to be as low as 80 kPa and as high as 200

kPa within the active soil depth. The final state of soil stress was equal to the net

effective overburden pressure. The resulting cumulative surface heave at the ground

surface ranged from 60 mm to 115 mm. The maximum surface heave resulting from the

modeling analysis was found to be 80 mm, which falls within the range of the estimated

values.

5.5 Soil Saturation - Pipe Displacement Analysis

Figure 5.5 illustrates the change in maximum soil surface displacements along with

the change in soil suction that occurred within the active soil zone. The plot indicates a

significant influence of the change in soil suction on the soil displacements. Maximum

soil displacement was predicted to be about 80 mm at the ground surface, due to a total

128

change in suction from 2000 kPa to 38 kPa. The figure also shows the maximum upward

displacements at the pipe level due to the same change in soil suction within the active

zone at a relative depth ratio (DR) of 0.9. The maximum pipe displacement was found to

be around 10 mm, at a depth of 2.7 m below the surface. The results of the upward soil

movement profiles can lead to an optimized pipe burial depth under the change in soil

suction and at the selected factor of safety. Similarly, avoiding the excessive fluctuation

in soil suction close to the area of the underground infrastructure may be an effective

way for minimizing soil movements and resulting displacement of underground

infrastructure.

Figure 5.6 presents the maximum pipe displacements predicted at the pipe level

for different pipe depth ratios of 0.5, 0.67, and 0.9. When the pipeline was installed at

shallower depth, it was subjected to higher displacement under the same variation in

water content and degree of saturation. The resulting pipe displacements increased

from 6% to 28% with the decrease of the relative depth ratio (DR) from 0.9 to 0.5. The

magnitude of upward soil movement was mainly affected by the change in soil elasticity

as a result of the increase in the total normal stress with depth.

Figures 5.7 and 5.8 show the normalized pipe displacement (pipe displacement/

pipe diameter, or d/D) profiles found at the pipe level along the pipeline length (X/L) at

different water content variations in the active zone. These two figures were obtained for

hinged and fixed end constraint conditions. In the case of hinged end restraints, the

maximum vertical displacement occurred at the left edge of the pipe in the active zone,

and it decreased along the length of the pipe. On the other hand, for fixed end restraints,

the maximum vertical displacement occurred around the approximate centre of the

active zone and decreased near the two edges of the pipeline.

The difference in the maximum normalized vertical displacement between the

hinged and fixed end restraint cases was found to be quite small (less than 0.2%) due to

129

the low modulus of elasticity for PVC pipes. For the studied cases, the maximum

normalized pipe displacement was found to be about 6%, corresponding to the relative

increase in volumetric water content, about 20% (approximately corresponds to the

change from the field condition to full saturation). The maximum pipe displacements

were approximately 2.1%, 3.7%, and 5.3%, with a relative increase in volumetric water

content of 5%, 10%, and 15% in the active zone, respectively.

In the case of significant variation environmental conditions applied on the ground

surface, burial depth was found to be a significant factor affecting the performance of

buried pipes. Burying pipelines should be completed deep enough to avoid significant

soil movements. The field ground surface conditions influence the amount of water

entered or lost from the soil structure and, therefore, it controls the burial depth. It is

important to take in consideration the actual condition of the ground surface in the

selection criteria of the burial depth of underground pipes. The right practice is to have

different burial depths with different surface conditions such as fully sealed surfaces

(concrete, pavement in good condition with no vegetation); semi-sealed surfaces (loose

fitting paving, loose fitting paving slabs), surfaces with vegetation trees and grass, and

unsealed surfaces.

130

Figure 5.5: Influence of the pipe burial depth on the pipe displacements

Figure 5.6: Maximum displacements versus normalized volumetric water content

Maximum Ground Displacement

Maximum Pipe Displacement

0 400 800 1200 1600 2000

Change in Soil Suction (kPa)

0

20

40

60

80

Dis

pla

cem

en

t (m

m) Initial suction = 2000 kPa

DR= 0.9

DR = 0.5

DR = 0.67

DR = 0.9

0 10 20 30 40 50

Normalized Volumetric Water Content, /i (%)

0

10

20

30

Maxim

um

Norm

aliz

ed

Pip

e D

isp

lace

men

t, d

/D (

%)

d/D = -0.005(/i)2 + 0.83(/i)

d/D = -0.003(/i)2+ 0.53(/i)

d/D = -0.001(/i)2 + 0.185(/i)

131

Figure 5.7: Pipe displacements due to the variation in soil moisture content under hinged end restraints

Figure 5.8: Pipe displacements due to the variation in the soil moisture content under fixed end restraints

0 0.2 0.4 0.6 0.8 1

Distance along the Pipe, X/L

-1

0

1

2

3

4

5

6

7

No

rma

lized

Pip

e D

ispla

ce

me

nt,

d/D

(%

)

Hinged-hinged Pipeline End Restrains

Active zone Inactive zone

= 20%

= 15%

= 10%

= 5%

Ep=2.8 GPa, DR=0.9

0 0.2 0.4 0.6 0.8 1

Distance along the Pipe, X/L

-1

0

1

2

3

4

5

6

7

No

rmaliz

ed

Pip

e D

ispla

ce

men

t, d

/D (

%)

Fixed-fixed Pipeline End Restrains

Active zone Inactive zone

= 20%

= 15%

= 10%

= 5%

Ep=2.8 GPa, DR=0.9

132

Figures 5.9 and 5.10 show the distribution of normalized pipe displacement along

the pipe for various properties of pipe materials. It was observed that the end restraints

of flexible pipes did not have an influence on the maximum displacement magnitudes of

the pipe (d/D). However it affected the distribution of displacement along the pipeline. On

the other hand, the end pipeline restraints showed a considerable effect on rigid pipes.

The maximum normalized pipe displacements were found to be around 6 % and 1.5 %

for hinged and fixed end restraints, respectively.

133

Figure 5.9: Pipe displacements in case of both hinged and fixed end restraints for a low elastic modulus magnitude (i.e. PVC pipe)

Figure 5.10: Pipe displacements in case of hinged and fixed end restraints for a high elastic modulus magnitude (i.e. steel pipe)

0 0.2 0.4 0.6 0.8 1

Distance along the Pipe, X/L

-1

0

1

2

3

4

5

6

7

No

rmaliz

ed

Pip

e D

ispla

ce

men

t, d

/D (

%) Fixed

Active zone Inactive zone

= 20 %

Ep=2.8 GPa, DR=0.9

Hinged

0 0.2 0.4 0.6 0.8 1

Distance along the Pipe, X/L

0

1

2

3

4

5

6

No

rma

lized

Pip

e D

isp

lace

men

t, d

/D (

%)

Fixed

Active zone Inactive zone

= 20 %

EP=280 GPa, DR=0.9

Hinged

134

CHAPTER 6: MATHEMATICAL FORMULATIONS OF SOIL-WATER

INTERACTION UNDER UNSATURATED SOIL CONDITIONS

6.1 Problem Statement

The mathematical formulation of unsaturated soil properties is necessary for the

computational modeling of water flow in soil deposits. As part of the formulation, it is

crucial to determine the soil water characteristic curve (SWCC) to identify the soil

response to transient water flow conditions. Transient water flow changes the stress

state in the soil structure, and consequently, the soil deforms in response to the changes

in the stress state and results in a new equilibrium state. The volume-mass constitutive

surfaces are the other key properties that provide an overall theoretical framework for

the soil state variables. The negative pore water pressure typically increases when a

load condition is being applied to an unsaturated soil structure. The consolidation

process may then take place when the excess pore water pressure is allowed to

dissipate and the soil volume will, subsequently, change with time. Therefore, for

unsaturated soils, the mechanical stress and pore water pressure should be coupled

together.

Unsaturated soils are often characterized by a variety of heterogeneities

(fractures/cracks) which affect water movement in soils (Novak et al., 2000). Infiltration

of precipitation and/or surface water into low-permeable soils (i.e. clays) is normally slow

(Novak et al., 2000). The presence of cracks increases the water flow into soil structure

and, in return, the soil water content. The depth of cracking is, therefore, a crucial

parameter that needs to be examined in order to understand the effects of desiccation

cracks. In the case of modeling saturated soils, the hydraulic conductivity can be

characterized by a lumped parameter because it contains relatively stable macropores

(Reynolds, 1993). Unsaturated cracked soils, however, cannot be modeled the same

135

way as saturated soils. Soil cracks are unstable and their geometry changes with time

depending upon the soil water content (Novak et al., 2000).

6.2 Methodology

In this chapter, a bimodal SWCC equation was developed for the native clay

deposit based on laboratory soil moisture-suction measurements. The proposed bimodal

SWCC equation was constructed by means of two distinct logistic power regression

models. A factor quantifying the change in the soil hydraulic conductivity was also used

to simulate the increase in the hydraulic conductivity of the cracked soil structure. In

addition, the hydraulic conductivity functions were developed for the cracked and un-

cracked soil layers. The constitutive surfaces were then established based on the

developed bimodal SWCC and representative consolidation test results. An unsaturated

seepage modeling analysis was also performed using the developed mathematical

framework. The model was utilized to simulate the application of net surface flux on a

soil column consisting of highly plastic clay over clay till. The sensitivity of the seepage

analysis, attributable to the variations in the SWCC and hydraulic characteristics, was

also demonstrated.

6.3 Fracture Depth Formulation

Morris et al. (1992) derived relationships between the cracking depth, depth to

ground water table, and surface suction as presented in Equation (6.1). The approach

assumed that the soil suction decreases linearly from the suction at the ground

surface (S0) to zero at the ground water table (dw) as presented in Equation (6.2).

The relationship also incorporated a coefficient representing the shear strength of the

soil as presented in Equations (6.3), (6.4) and (6.5).

136

(6.1)

(6.2)

(6.3)

(6.4)

(6.5)

Where;

is the cracking depth,

S is the suction,

S0 is the suction at the ground surface,

z is the depth below the ground surface,

is the distance from the ground surface to the ground water table,

is the compression modulus,

μ is the Poisson's ratio,

is the unit weight of soil,

is the coefficient for the angle of shearing resistance,

is the soil friction angle with respect to the total stress, and

is soil friction angle with respect to the soil suction.

Lau (1987) and Fredlund and Rahardjo (1993a) published a mathematical method

for determining the depth of cracking as a function of shear strength based on the

Rankine theory of lateral earth pressure as shown in Equation (6.6). The method

assumed that at rest, the coefficient earth pressure and the net horizontal stress were

equal to zero at the bottom of the soil crack. Figure 6.1 shows the mathematical model

for desiccation cracks and a typical soil suction profile presenting the negative pore-

water pressure as a linear function of distance above the groundwater table. The

137

equation is a function of the ground water depth, elasticity ratio, soil unit weight and

Poisson's ratio of soil. The relationship can generally be used to estimate the cracking

depth for a range of the elasticity ratio of soil.

μ

(6.6)

where;

is the depth of cracking,

/H is the elasticity parameters ratio. For an initially saturated clay, it ranges from

0.15 - 0.2 (Lau, 1987),

is the unit weight of water, and

is a variable used to permit the pore-water pressure to be represented as a

percentage of the hydrostatic profile.

Other variables were previously defined.

138

Figure ‎6.1: A typical desiccated soil profile and idealized matric suction profile, after

(Lau, 1987)

139

Values of the estimated cracking depth were determined using the two approaches

discussed above and presented in Equations (6.1) to (6.6). At the time of the field

investigation, there was no definite ground water table up to the explored depth of 16 m.

Due to the relatively granular and seepage nature of the clay till, the ground water depth

was then assumed to be either located at the maximum explored depth (16 m) or just

beneath the clay deposit (9.5m). The assumed ground water table levels are considered

reasonably representative of both highly wet conditions during rainy conditions and the

creation of a perched groundwater table within the clay till. Figures 6.2 to 6.4 show the

variation in estimated cracking depth at different ground water depths and different soil

conditions.

It was found that the linear elastic method resulted in higher estimated cracking

depths. Based on the analytical results, a cracking depth of up to 3 m can be expected

and can reasonably represents the field conditions as previously discussed. Based on

the results shown in Figure 6.2, a cracking depth of 3 m may be expected at a surface

suction of 100 kPa. In addition, based on the results shown in Figures 6.3 and 6.4, a

cracking depth of 3 m, may however, be expected at an elasticity ratio (E/H) of

approximately 0.15 and 0.11 for a ground water depth of 16 m and 9.5 m, respectively,

assuming that the hydrostatic profile factor is approximately 1.5.

140

Figure 6.2: Estimated cracking depth at different ground water depths (linear elastic theory)

Figure 6.3: Estimated cracking depth at a ground water depth of 9.5m (shear strength approach)

0 200 400 600 800 1000

Surface Soil Suction, So

0

4

8

12

16

Cra

ckin

g D

epth

, d

cr (m

)dw

= 16m

dw= 9m

1 1.2 1.4 1.6 1.8 2

Hydrostatic Profile Factor, fw

0

1

2

3

4

Cra

ckin

g D

epth

, d

cr (m

)

E/H = 0.2

E/H = 0.15

E/H = 0.1

E/H = 0.05

Ground Water Depth, dw= 9.5 m

141

Figure 6.4: Estimated cracking depth at a ground water depth of 16 m (shear strength approach)

1 1.2 1.4 1.6 1.8 2

Hydrostatic Profile Factor, fw

1

2

3

4

5

6

Cra

ckin

g D

epth

, d

cr (m

)

E/H = 0.2

E/H = 0.15

E/H = 0.1

E/H = 0.05

Ground Water Depth, dw= 16 m

142

6.4 Hydraulic Conductivity Formulation

The upper three meters of the clay layer were assigned a higher hydraulic

conductivity to simulate the effect of the seasonal development of desiccation cracks. A

hydraulic conductivity factor labeled (fcr) was assumed to stand for the increase in the

hydraulic conductivity due to the propagation of cracks. The hydraulic conductivity factor

was defined as the ratio between the hydraulic conductivity of the cracked soil structure

to the hydraulic conductivity of the natural soil (un-cracked) structure (fcr = kcr/kn). In view

of the fact that the crack development and distribution is a dynamic process and is

generally associated with the water infiltration/exfiltration processes. The factor was then

defined to be a function of the soil suction as a stress state that reflects the soil moisture

conditions.

The maximum increase in the hydraulic conductivity (a value represented by fmax)

is specified at the residual VWC and residual soil suction (Ψr) assuming that the cracks

are fully developed. However, the factor was assigned a value of 1 at completely

saturated condition (Ψsat). Using this definition, Equation 6.7 was incorporated into the

modeling program to simulate the final hydraulic conductivity of the top soil layer with

time.

Ψ Ψ

Ψ Ψ (6.7)

where;

is the hydraulic conductivity factor defining the ratio between cracked and un-

cracked soil condition,

is a variable used to represent the maximum hydraulic conductivity ratio

between cracked and un-cracked soil at an entirely dry condition,

Ψ is the soil suction,

Ψ is the soil suction corresponding to the residual VWC, and

143

Ψ is the soil suction corresponding to the saturated VWC.

6.5 Development of a Bimodal SWCC

Precise representation of the SWCC is essential for the seepage analysis of

saturated-unsaturated soil deposits. A bimodal SWCC was developed using the

measured data sets obtained from laboratory testing of the soil suction at different water

content levels. Figure 6.5 shows the measured data points for the field samples as two

sets, namely (1) and (2) that were obtained using the pressure plate and filter paper

methods as discussed in (Hu and Vu, 2011). The proposed equation of the bimodal

SWCC was developed using two distinct logistic regression models fitting the laboratory

data sets before and after a turning point. The turning point appeared at a soil suction of

100 kPa as shown in Figure 6.5. The developed equation and the corresponding fitting

parameters are included in Equation (6.8). It is important to mention that the presented

fitting equation was found to match a previous SWCC unimodal equation proposed by

(Gardner, 1958). However, the fitting parameters are predicted in the form of two sets,

in order to establish a bimodal curve. Figure 6.5 also shows the predicted SWCC using

the Fredlund and Xing fitting equation and based on the GSDC of the native clay.

θ θ

θ

θ

θ

θ

(6.8)

where;

θ is the volumetric water content (VWC) (vol/vol),

θt and t are the volumetric water content and soil suction at the turning point (0.53

and 100 kPa, respectively),

is the soil suction (1 ≤ ≤ 106 kPa),

θsat is the saturated volumetric water content (0.63) (vol/vol),

144

ai and bi are the curve fitting parameters for the first part of the curve (3.844, and

2.542, respectively), and

af and bf are the curve fitting parameters for the second part of the curve (3.56,

and 7.464, respectively).

The developed bimodal SWCC fitted well the laboratory data measurements and

showed two air entry values as initial and final values of 15 kPa and 200 kPa,

respectively. The bimodal SWCC was, however, positioned below the unimodal fit. Also,

with an increase in the soil suction, the difference between the unimodal and bimodal

SWCCs increased. This difference was considered as a good representation of the

divergence between the bases of developing each of these curves. Unimodal SWCCs

established based on the grain size distribution of the soil may take into account the

native formation of the particles. However, it is important to note that clay of dispersed

structure would result in a different soil water interaction behavior than that of fluctuated

structure even if both of them have the same grain size. The bimodal was, alternatively,

developed based on a reasonable range of measured VWC and soil suction values.

Therefore, it was able to better capture the structure of the clay. By comparing both the

unimodal and bimodal soil curves, it is clear that the bimodal curve, at the same soil

suction, shows greater water moisture reduction than the unimodal one.

A parametric study was conducted to determine the SWCC using different values

of fitting parameters. Figures 6.6 and 6.7 show the effect of change in fitting parameters

on the SWCC shape. It was noted that the fitting parameters can be used to plot a wide

range of variation in soil suction and VWC. It was also noted that the variation in the

parameter (a) did not result in a change in the slopes of the SWCC, and it controlled

more the position of the SWCC in relation to the best fit. However, the parameter (b) had

more control on the slope of the plotted SWCC. As shown in the figure, the bimodal

SWCC consisted of two SWCC functions that were superimposed to create one function

145

for the cracked soil as defined by (Durner, 1994). The two SWCC functions along with

their distinct parameters were developed representing the intact and cracked parts of the

soil. The bimodal SWCC was found to be a practical tool to simulate the soil as a dual-

porosity medium (Köhne et al., 2002). Using the developed bimodal SWCC, the soil

behavior can then be modeled as a combination of two materials averaged over the

whole soil volume demonstrating the change in the soil behavior caused by the soil

cracks (Fredlund et al., 2002).

146

Figure 6.5: Bimodal SWCC and laboratory suction measurements

Figure 6.6: SWCC shapes at different values for the fitting parameter (a)

1 10 100 1000 10000 100000 1000000

Matric Suction, (kPa)

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

Vo

lum

etr

ic W

ate

r C

on

ten

t, (

vo

l/vo

l)

Laboratory Measurements Set (1)

Laboratory Measurements Set (2)

Fredlund and Xing SWCC Unimodal Fit

Proposed SWCC Bimodal Equation Fit

Residual Water Content,s= 0.085

Initial Air Entry Value, AEVi

15 kPa

Final Air Entry Value, AEVf

200 kPa

1 10 100 1000 10000 100000 1000000

Matric Suction, (kPa)

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

Vo

lum

etr

ic W

ate

r C

on

ten

t, (

vo

l/vo

l)

Bimodal SWCC Equation

0.6a

0.8a

1.2aa

147

Figure 6.7: SWCC shapes at different values for the fitting parameter (b)

1 10 100 1000 10000 100000 1000000

Matric Suction, (kPa)

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

Vo

lum

etr

ic W

ate

r C

on

ten

t,

(vo

l/vo

l)

Bimodal SWCC Equation

0.6b

b1.2b

0.8b

148

6.6 Volume-mass Constitutive Relationships

The constitutive surfaces can typically be constructed using the relationships of the

void ratio, water content, and degree of saturation with the change in soil suction and net

normal stress. The unsaturated soil structure interaction under loading and unloading as

well as wetting and drying processes can be modeled using the volume-mass

constitutive surfaces. These constitutive surfaces define the primary soil state variables

(i.e., void ratio, water content and degree of saturation) as a function of the two

independent stress state variables (i.e. matric suction and net normal stress) forming 3D

surfaces. Using these constitutive surfaces, all soil state variables can be obtained over

the entire ranges of possible pore pressure and stress values. Matyas and Radhakrishna

(1968) recommended that the soil element can be graphically characterized as a point in

a system (i.e., a state point) where the coordinate axes stand for the state parameters, in

order that changes in the stress state can be defined as a state path. Fredlund and

Rahardjo (1993b) presented the main principles to construct 3D constitutive surfaces as

summarized below (Zhang, 2004):

(i) Void ratio versus net normal stress plot when the matric

suction is equal to zero, [e=f(σmean, = 0)],

(ii) Void ratio versus suction plot when the net normal

stress is equal to zero [e=f(σmean =0, )],

(iii) Water content versus net normal stress plot when the

matric suction is equal to zero, [w=f(σmean, = 0)],

(iv) Water content versus suction plot when net normal stress

equals zero, [w=f(σmean = 0, )],

(v) Degree of saturation versus mechanical stress plot when

the matric suction is equal to zero [S=f(σmean, = 0)], and

149

(vi) Degree of saturation versus matric suction plot when

net normal stress is equal to zero [S =f(σmean =0, )].

Two constitutive relationships are considered necessary for the two stress state

variables, one is a change in mechanical stress and the other is a change in matric

suction (Zhang, 2004). In order to define the constitutive relationships, the bimodal

SWCC was used to define the relationship between the matric suction and VWC. In

addition, previous laboratory measurements presented by Fredlund (1967) was used to

plot the relationship between the normalized VWC and mean normal stress. Figure 6.8

presents the relationship between the normalized VWC and mean normal stress as

defined in Equation (6.9). Figure 6.9 shows the different plots for the relationship

between the VWC and net normal stress using different values for the input fitting

parameters.

Likewise, previous laboratory measurements presented by Fredlund (1967) were

used to define the relationship between the normalized VWC and mean normal stress as

defined in Equation (6.10). This relationship was used to predict the degree of saturation

at different VWC levels. Figure 6.10 presents the relationship between the VWC and

degree of saturation. Figure 6.11 present the different shapes of the relationship at

different fitting parameters. Figures 6.12, 6.13, and 6.14 show the predicted volume-

mass constitutive surfaces. These constitutive surfaces were utilized to predict the

unsaturated soil properties of native clay. The plotted constitutive surfaces demonstrate

the change in the unsaturated soil properties for a range of the net normal stress and

soil suction. They incorporated the use of the bimodal SWCC that represents a good fit

with the laboratory measurements.

150

θ

θ σ σ σ

(6.9)

where;

σ σ and are fitting parameters, and σ is the net normal stress.

For Regina clay, best-fit curves were found with the following numerical values:

= 0.285593, 0.1815003, and -0.17729

θ

θ (6.10)

where;

and are fitting parameters,

is the degree of saturation, and

θ is the volumetric water content (VWC) (vol/vol).

For Regina clay, best-fit curves were found with the following numerical values:

b = 0.007576, c= 97.92621, and d = 2.17763

151

Figure 6.8: Normalized volumetric water content versus mean normal stress

1 10 100 1000

Mean Normal Stress, (kPa)

0

0.1

0.2

0.3

0.4

No

rmaliz

ed V

olu

me

tric

Wa

ter

Con

tent,

Obtained based on the Results of Fredlund (1967)

The Proposed Fitting Equation

= (a

+ b

log

mean)(

-1

d)

R2 = 0.999

152

Figure 6.9: Parametric analysis for normalized volumetric water content versus mean normal stress relationship

0

0.4

0.8

1.2

1.6

No

rma

lize

d V

olu

me

tric

Wate

r C

onte

nt,

Regina Clay (After Fredlund (1967))

0

1

2

3

No

rmaliz

ed V

olu

me

tric

Wa

ter

Con

tent,

1 10 100 1000

Mean Normal Stress, mean

(kPa)

0

0.2

0.4

0.6

No

rmaliz

ed V

olu

me

tric

Wa

ter

Con

tent,

b = 0.182, d = -0.177

a = 0.286, d = -0.177

a = 0.286, b

= 0.182

= (a

+ b

log

mean)(

-1

d)

a= 0.286

a= 0.1

a =

0.5

b= 0.182

b

= 0.1

b = 0

.3

d =

-0.1

77

d = -0

.1

d = -0

.3

R2 = 0.999

153

Figure 6.10: Degree of saturation versus volumetric water content

0 0.2 0.4 0.6 0.8 1

Volumetric Water Content, (vol/vol)

0

20

40

60

80

100

Deg

ree o

f S

atu

ration

, S

(%

)

Regina Clay (after Fredlund (1967))

The proposed equation

S = c d

b+d R2= 0.993

154

Figure 6.11: Parametric analysis for degree of saturation versus volumetric water content relationship

0

20

40

60

80

100

De

gre

e o

f S

atu

ratio

n,

S (

%)

Regina Clay (After Fredlund (1967))

0

20

40

60

80

100

De

gre

e o

f S

atu

ration

, S

(%

)

c = 100

c = 97.93

c = 95

0 0.2 0.4 0.6 0.8 1

Volumetric Water Content, (vol/vol)

0

20

40

60

80

100

De

gre

e o

f S

atu

ratio

n, S

(%

)

b =

0.02

b =

0.0

01

d =

2.1

8

d =

5

c = 97.93, d = 2.18

b = 0.008, d = 2.18

c = 97.93, b = 0.008

S = c d

b+d

b =

0.0

08

d = 1

R2= 0.993

155

Figure 6.12: Volumetric water content constitutive surfaces

Figure 6.13: Void ratio constitutive surfaces

110

1001000

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

Vo

lum

etr

ic W

ate

r C

on

ten

t,

(vo

l/v

ol)

0.6-0.7

0.5-0.6

0.4-0.5

0.3-0.4

0.2-0.3

0.1-0.2

0-0.1

1

10

100

1000

0

0.2

0.4

0.6

0.8

1

1.2

Vo

id R

atio

, e

1-1.2 0.8-1

0.6-0.8 0.4-0.6

0.2-0.4 0-0.2

156

Figure 6.14: Degree of saturation constitutive surfaces

110

1001000

0

20

40

60

80

100

De

gre

e o

f Sa

tura

tio

n,

S (%

)

80-100

60-80

40-60

20-40

0-20

157

6.7 Water Flow Mathematical Formulation

Above the ground water table, the pressure heads are negative, and the hydraulic

head can be determined indirectly through field measurements of the suction head or

negative pore-water pressure. As discussed earlier in this thesis, Fredlund and

Rahardjo (1993b) presented a general partial differential equation based on Richards’

equation for the transient moisture flow through saturated/unsaturated soils as illustrated

in Equation (6.11).

(6.11)

where;

h is the total head (m),

is the slope of the soil water characteristic curve, and

is the unit weight of water (kN/m3).

6.8 Parametric Study

6.8.1 General

A parametric study was performed to characterize the effect of the change in

unsaturated soil properties on the water seepage analysis for a soil column. The

numerical model validated the reliability of the mathematical framework and quantified

the effects of different boundary conditions. The modeling of soil moisture variation was

based on the governing partial differential equations of seepage through saturated-

unsaturated soils as discussed in the previous sections. The finite element modeling was

implemented using the commercial finite element program FlexPDE (PDE Solutions Inc.,

2014). The FlexPDE's model script was written based on the developed mathematical

framework and then the operations were executed to transform the description form of

the partial differential equations to a finite element model, run the analysis and produce

graphical demonstration of the results.

158

6.8.2 Methodology

The soil column consisted of a highly plastic clay (Regina clay) over clay till. The

width and height of the column were 3 m and 16 m, respectively. The finite element

mesh for the problem was generated by FlexPDE as presented in Figure 6.15. Cell sizes

were typically controlled by the spacing between explicit points in the domain boundary.

The developed initial mesh consisted of triangular finite elements with an average size of

0.4 m over the arbitrary two-dimensional problem domain.

The seepage analysis was performed using the following boundary conditions:

At the top of the column, a net flux was applied on the soil column. The

magnitude of the applied flux was assumed to be 0.75 mm/day, 1 mm/day, and 2

mm/day,

At the bottom of the column, net flux equaled zero,

The sides of the column had no applied flow boundary condition,

The highly plastic clay layer was divided into two layers, the upper layer

extended to a depth of (dcr), and the bottom clay layer extended to a depth of

9.5m, and

The initial VWC was assumed to be 0.3, 0.36, 0.25 for the top clay, bottom clay,

and clay till layers, respectively. The corresponding initial soil suction values

were then determined to be 2800 kPa, 1600 kPa, and 1400 kPa, respectively.

The seepage modeling results were obtained in terms of soil parameters such as VWC

and soil suction. The variation in these parameters as functions of both depth and time

were obtained at the following depths below the ground surface, 1.5 m (within the top

desiccated clay layer), 6.25 m (within the bottom clay layer), and 10 m (i.e., near the top

of the clay till layer).

159

Figure 6.15: Soil column configuration for the parametric study

160

6.8.3 Results and Discussion

A series of analyses using both unimodal and bimodal SWCCs were conducted

under a net surface flux of 1 mm/day and hydraulic conductivity factor of 1 as shown in

Figures 6.16 and 6.17. The pattern for the change in VWC was different, and roughly

followed the shape of the utilized SWCC characteristics. It is clear that the bimodal

SWCC resulted in a rapid reaction for the change in soil moisture conditions to the

applied net flux in comparison with the unimodal SWCC. The top clay layer approached

a full saturation condition in about five times the time required for the bimodal SWCC

model. The results suggest that the shape and characteristic of the SWCC have an

influence on the outcome of the seepage analysis and, therefore, the SWCC should be

formulated precisely to reflect the field conditions. Cracked soils were found to exhibit

bimodal behavior because of the pore space distribution discrepancy in the soil matrix

structure that is typically created by the cracks. The bimodal SWCC function resulted in

allowing greater infiltration and was considerably different from the unimodal one.

Therefore, it is clear that the use of bimodal SWCC provided a more practical method for

the simulation of the soil water interaction for highly plastic clays.

At a depth of 1.5 m and under a net surface flux of 1 mm/day and a maximum

cracking factor (fmax) of 1, 50, and 200, the variation in the resulting hydraulic

conductivity factor and the hydraulic conductivity of the top clay were plotted with time as

shown in Figures 6.18 and 6.19. The initial hydraulic conductivity of the clay was in the

order of 10-12 m/sec. The resulting initial hydraulic conductivity of the clay was in the

order of 10-10 m/sec (i.e., an increase of two orders of magnitude), and 10-11 m/sec (i.e.,

an increase of one order of magnitude) corresponding to fmax of 200 and 50, respectively.

The hydraulic conductivity was almost the same after approximately 235 days when the

soil approached near saturation condition.

161

Figure ‎6.16: Predicted suction profiles for Bimodal SWCC versus Unimodal SWCC

Figure 6.17: Predicted volumetric water content profiles for Bimodal SWCC versus Unimodal SWCC

0 500 1000 1500 2000 2500

Elapsed Time, days

0

1000

2000

3000

Matr

ic S

uctio

n, (

kP

a)

I = 1 mm/dayDepth = 1.5 m

B

A: Bimodal SWCCB: Unimodal SWCC

A

0 500 1000 1500 2000 2500

Elapsed Time, days

0.2

0.4

0.6

0.8

Vo

lum

etr

ic W

ate

r C

onte

nt,

(vol/vol)

I = 1 mm/dayDepth= 1.5 m

A

B

A: Bimodal SWCCB: Unimodal SWCC

162

Figure 6.18: Hydraulic conductivity factor versus elapsed time

Figure 6.19: Hydraulic conductivity versus elapsed time

0 60 120 180 240 300 360

Elapsed Time, days

0

20

40

60

Hydra

ulic

Conductivity F

acto

r, f

cr

fmax= 1

fmax= 50

fmax= 200

Full Saturation

I = 1 mm/dayDepth = 1.5 mdcr = 3 m

0 60 120 180 240 300 360

Elapsed Time, days

1E-012

1E-011

1E-010

1E-009

1E-008

Hydra

ulic

Conductivity, k (

m/s

ec)

fmax= 1

fmax= 50

fmax= 200

I = 1 mm/dayDepth = 1.5 mdcr = 3 m

163

Figure 6.20 shows the variation in soil suction with the change in the hydraulic

conductivity of the clay. When the (fmax) increased from 50 to 200, the change in soil

suction was not as noticeable as the change from 1 to 50. In addition, the change in the

resulting VWC took less time to start with the increase in the hydraulic conductivity. This

overall behavior was reasonably able to model the dynamicity of the change of soil

hydraulic characteristics with the formation/ discontinuation of soil cracks.

The change in soil moisture was studied for different cases where the hydraulic

cracking factor was applied to different soil depths representing the increase in the depth

of the top desiccated clay layer. Using a maximum cracking factor (fmax) of 50, the

change in soil moisture characteristics were predicted for different cracking depths of 1

m, 2 m, and 3 m as shown in Figure 6.21. There were noticeable differences in the VWC

profiles with the increase in cracking depth. Due to the increase in the hydraulic

conductivity, the change in VWC started earlier and resulted in higher VWC magnitudes.

However, after a certain time period (approximately 50 days), increasing cracking depth

resulted in lower VWC magnitudes. This behavior reflected the effect of the change in

the seepage process of the cracked soil structure versus the un-cracked one. When the

soil is cracked to a certain depth, the water seepage is expected to be more rapid, and

water reaches a greater depth. However, the change in VWC at a certain depth would

take longer to reach saturation condition.

After approximately 140 days, the resulting VWC profile was almost the same with

the increase in cracking depth reflecting the small effect of soil cracking when the soil

reached near saturation condition. This behavior corresponded to the change in

hydraulic conductivity factor with time as described earlier. These results demonstrate

the significance of properly defining the depth of cracking in soil water interaction

analysis.

164

Figure 6.20: Predicted suction versus elapsed time at cracking factor magnitudes

Figure 6.21: Predicted volumetric water content versus elapsed time at different cracking depths

0 60 120 180 240

Elapsed Time, days

1

10

100

1000

Matr

ic S

uctio

n, (

kP

a)

fmax = 1

fmax = 50

fmax = 200

I = 1 mm/dayDepth = 1.5 mdcr = 3m

0 60 120 180 240 300 360

Elapsed Time, days

0.2

0.3

0.4

0.5

0.6

0.7

Vo

lum

etr

ic W

ate

r C

on

ten

t,

dcr= 0 m

dcr= 1 m

dcr= 2 m

dcr= 3 m

Full Saturation

fmax = 50, I = 1 mm/day

Depth = 1.5 m

165

Figures 6.22 and 6.23 show the variation in VWC and soil suction with time at

three different depths 1.5 m, 6.26 m, and 10 m below the surface. In general, there was

an increase in VWC with time as a result of the water seepage process. The VWC

gradually increased until it reached a constant value corresponding to the full saturation

condition. The change in VWC followed a bimodal pattern. A time delay for the start of

the change in VWC occurred and increased with the increase in depth reflecting the time

at which the soil water interaction started. Likewise, the time required to reach saturation

was different at each level.

Finally, the effect of the change in the applied surface flux was also investigated

as shown in Figures 6.24 and 6.25. Both VWC and soil suction gradually changed with

time and finally reached a constant value corresponding to those values at the full

saturation condition. There were differences in time at which the soil became saturated

for the different applied net fluxes. Increasing the magnitude of the applied flux resulted

in a rapid soil saturation progression. However, the variation in VWC and soil suction

followed almost the same pattern.

166

Figure 6.22: Predicted suction versus elapsed time at different depths

Figure 6.23: Predicted volumetric water content versus elapsed time at different depths

0 60 120 180 240 300 360

Elapsed Time, days

0

1000

2000

3000

Matr

ic S

uctio

n, (

kP

a)

Depth = 1.5 m

Depth = 6.25 m

Depth = 10 m

fmax = 1.0

I = 1 mm/daydcr = 3 m

0 60 120 180 240 300 360

Elapsed Time, days

0.2

0.3

0.4

0.5

0.6

0.7

Volu

me

tric

Wa

ter

Conte

nt, (

vol/vol)

Depth = 1.5 m

Depth = 6.25 m

Depth = 10 m

fmax = 1.0, I = 1 mm/day,

dcr = 3 m

167

Figure 6.24: Predicted volumetric water content versus elapsed time at different net surface flux magnitudes

Figure 6.25: Predicted suction versus elapsed time at different net surface flux magnitudes

0 60 120 180 240 300 360

Elapsed Time, days

0.2

0.3

0.4

0.5

0.6

0.7

Vo

lum

etr

ic W

ate

r C

onte

nt,

(vol/vol)

I = 2 mm/day

I = 1 mm/day

I = 0.75 mm/day

Full Saturation

fmax = 50

Depth = 1.5 mdcr = 3m

0 60 120 180 240 300 360

Elapsed Time, days

1

10

100

1000

Matr

ic S

uctio

n, (

kP

a)

I = 2 mm/day

I = 1 mm/day

I = 0.75 mm/day

fmax = 50

Depth = 1.5 mdcr = 3m

168

CHAPTER 7: SOIL-PIPE-ATMOSPHERE INTERACTION UNDER FIELD

CONDITIONS

7.1 Problem Statement

Predicting soil moisture content, soil suction, and temperature profiles is a key

element to study the field performance of underground pipelines. Developing a

framework to simulate real cases and incorporate more complex boundary conditions

and advanced unsaturated soil parameters is necessary. Desiccation cracks occur at

some stage due to the soil shrinkage during evapotranspiration processes (mainly, soil

water loss). The development of soil cracks influences the hydraulic conductivity and the

SWCC of the near-surface clay layers, and consequently, the moisture content changes

with time. Therefore, incorporating the development of soil cracks in the simulation of the

soil behavior is critical.

A key property that is vital for the implementation of unsaturated soil principles is

the soil water characteristic curve (SWCC). Bimodal SWCCs have not been heavily used

or verified in the modeling of the unsaturated soil deposits in many field applications, and

therefore, was used in this research study. Climatic conditions are other key factors of

the problem’s boundary conditions. However, data availability is a challenge constraining

not only the features of the modeling process but also the resulting predictive ability and

accuracy. Therefore, weather data from a nearby Environment Canada Meteorological

Station (ECMS) were analyzed for their applicability to the current study.

7.2 Methodology

The main objective of the proposed model was to investigate the response of a

150 mm PVC pipe buried in a highly plastic clay deposit under field conditions. A model

was established to incorporate the daily climatic conditions and to examine the effect of

the atmospheric conditions on the soil and pipe behavior. The required soil properties

169

and initial conditions for the model were determined from the laboratory results of the

collected samples from the field. To accurately model the field performance, advanced

interpretations were developed and validated to better simulate the development of

surface desiccation cracks. The factor quantifying the change in the soil hydraulic

conductivity, developed in this research, was used to model the increase in the hydraulic

conductivity of cracked soil structure compared to an un-cracked one. In addition, the

bimodal SWCC equation was utilized in this model to simulate the surficial highly plastic

clay layer surrounding the pipe trench.

The general processes and major steps of the numerical program are presented

in Figure 7.1. The program aimed to predict the displacement that occurred in the pipe

due to the change in climatic conditions with time. The model was established based on

a series of equations for heat and mass transfer in saturated-unsaturated soil structures

and a bimodal SWCC. The numerical model incorporated the simulation of the actual

evaporation of soil profile under field conditions. The prediction of final soil suction is the

connecting element between the two processes. The seepage analysis was established

to output time dependent suction changes influenced by the applied climatic conditions.

The stress-strain analysis used the suction data from the seepage analysis for its initial

and final conditions and then outputted the soil displacements. The entire analysis

utilized a general-purpose partial differential equation solver FlexPDE version 6 (PDE

Solutions Inc., 2014). The monitoring data of a section of an instrumented water pipe

buried in highly expansive native clay was also utilized. The field measurements were

utilized to develop the understanding of the soil-pipe interaction under field conditions,

and to quantify the effect of the variation of different soil-pipe parameters such as ground

surface conditions.

170

7.3 Geometry and Boundary Conditions

The modeling was undertaken using a two-dimensional model having the

atmospheric boundary condition at the top and no-flow boundary condition at the bottom

as well as left and right sides. Figure 7.2 presents the geometry and boundary conditions

used for the two-dimensional analysis. The pipe location and depth, pipe trench

dimensions and ground surface conditions were considered in accorance with the field

site conditions. Snow presence during the period between November and March was

assumed as total precipitation and applied right after the ground temperature rose to

above zero, representing the thawing process of the accumulated snow on the ground

surface. The surface boundary condition consisted of two zones to simulate the field site.

The first zone was covered with vegetation and, therefore, it was in full

interaction with the daily atmospheric conditions as well as the park watering during the

spring-summer season (from June to October). The park watering was assumed to be

applied twice a week with the amount of 1.81 mm per day in accordance with the field

site records. However, the second zone was covered with a pavement structure, which

reduced the soil-atmosphere interaction to some degree. Figure 7.3 shows the model

mesh for the analysis. Free movement in the vertical direction was allowed at the top

boundary and the horizontal movements of both left and right sides were fixed. The

lower boundary was fixed in both horizontal and vertical directions. The PVC pipe had a

nominal diameter of 0.15 m and was buried at a depth of 2.9 m below the ground

surface. Table 7.1 shows a summary of the key geotechnical index properties for the

native clay (Regina clay), the backfill (mixed concrete), and the pipe bedding (sand).

Table 7.2 shows a summary of the main parameters of the stress-strain modeling

analysis.

171

Figure 7.1: Soil-pipe-atmosphere modeling processes

Figure ‎7.2: Schematic diagram showing the field site conditions

172

Figure 7.3: The developed mesh for the modeling analysis

173

Table 7.1: Geotechnical index properties for the clay, mixed concrete and sand

Soil Property (unit) Clay Mixed Concrete Sand

Specific gravity, Gs 2.7 2.7 2.65

Average dry density, d (kN/m3) 15.40 18.30 18.20

Average wet density (kN/m3) 19.87 22 20.9

Natural water content, w (%) 23-35 - -

Initial void ratio, e0 0.95 0.5 0.45

Liquid limit (%) 64-94 - -

Plastic limit (%) 23-34 - -

Plastic index (%) 37-66 - -

Swelling index, Cs (Consolidation test)

0.09 - -

Swelling index, Cm (SWCC test) 0.08 - -

Saturated hydraulic conductivity (m/sec)

1.6 X 10-9 to 2.8 X 10-8

7.50 X 10-6 2.30 X 10-5

Table 7.2: Summary of the main parameters of the stress-strain modeling analysis

Parameter (unit) Value

Modulus of elasticity of mixed concrete (MPa) 80

Modulus of elasticity of sand (MPa) 60

Modulus of elasticity of pipe (GPa) 2.8

Poisson’s ratio of pipe 0.2

Poisson’s ratio of clay 0.4

Poisson’s ratio of mixed concrete 0.35

Poisson’s ratio of sand 0.3

Pipe diameter (m) 0.15

Pipe burial depth (m) 2.9

174

7.4 Mathematical Formulation

7.4.1 Evapotranspiration Process

The term (Ep) was used to indicate the evapotranspiration flux from the soil surface

to the atmosphere. The analysis in this study was performed using the modified Penman

Equation (Penman, 1948) for actual evapotranspiration (Wilson et al., 1997; Wilson,

1990; Wilson et al., 1994). The potential evapotranspiration at the soil-atmosphere

boundary was calculated as shown in Equations (7.1) to (7.7).

(7.1)

(7.2)

(7.3)

(7.4)

(7.5)

(7.6)

(7.7)

where;

Ep is the evapotranspiration flux (m/day),

PE is the potential evapotranspiration (m/day),

RHa is the relative humidity at the soil surface,

is the air vapour pressure of air (mmHg),

is the air vapour pressure at the soil surface (mmHg),

is the saturation air vapour pressure (mmHg),

is the saturation air vapour pressure at the soil surface (mmHg),

175

is the slope determined from the saturation vapour pressure versus temperature

curve (mmHg/ºF) and can be determined as shown in Equation (7.6) (Zotarelli et

al., 2014),

QN is the net available radiant energy at the surface (m/day)(Gray, 1973);

is the psychrometer constant (0.27 mm Hg/ºF),

RHa is the air relative humidity,

Ta is the atmospheric air temperature,

Ua is the wind speed (m/day), and

Re is the net radiation.

7.4.2 Water Flow Equations

The transient moisture flow through saturated-unsaturated soil condition was

determined according to Equation (7.8) (Fredlund and Rahardjo, 1993b).

(7.8)

where;

h is the total pressure head,

kx is the hydraulic conductivity in x-direction,

ky is the hydraulic conductivity in y-direction,

is the slope of the soil water characteristic curve, and

is the unit weight of water.

The partial differential equation intended for conductive heat flow in soils was

identified according to Equation (7.9) (Pentland et al., 2001).

176

(7.9)

where;

λx and λy are the soil thermal conductivity in x- and y- direction,

T is the temperature,

c is the mass specific heat, and

ρ is the soil density.

The term cρ is designated to the volumetric specific heat capacity of the soil. The mass

specific heat (c) was assumed to be constant and estimated to be 1.2 x 103 J/kg.K based

on the average of typical specific heat of soils (De Vries, 1963; List, 1966; Oke, 2002).

The thermal conductivity for the frozen and unfrozen backfill soil was estimated in

accordance with Equations (7.10) and (7.11) (Farouki, 1986; Kersten, 1949).

(7.10)

] (7.11)

Likewise, the thermal conductivity for the frozen and unfrozen native clay soil was

estimated in accordance with Equations (7.12) and (7.13) (Farouki, 1986; Kersten, 1949)

(7.12)

(7.13)

7.4.3 Stress-strain Equations

The governing nonlinear stress-strain equations are presented in terms of

displacements in the x- and y- direction (u and v). The incremental elastic forms of these

equations were provided in Equations (7.14) and (7.15) (Fredlund and Vu, 2003). The

incremental elastic forms of these equations were incorporated in Equations (7.16) and

(7.17) (Fredlund and Vu, 2003). The elasticity parameters were calculated using the

volume change indices and an assumed Poisson’s ratio.

177

(7.14)

(7.15)

(7.16)

(7.17)

where;

μs is the Poisson’s ratio for the soil,

Fx and Fy are x- and y- components of the body force vector,

E is the elasticity parameter as a function of the net normal stress,

H is the elasticity parameter as a function of the change in matric suction,

dσ is the net normal stress,

is the change in matric suction,

E is the elasticity parameter,

µ is Passions' ratio,

Cs and Cm are swelling indices obtained from net normal stress condition and

matric suction condition, respectively, and

Ce, C11, C22, and C33 are the stiffness tensor components as defined by (Fredlund

and Gitirana Jr, 2005).

7.5 Climate Data

The climate is a contributing factor to the seasonal variations in the moisture

content and temperature of the surficial soil layer. Air temperature and precipitation data

for the analyzed site were obtained from a nearby Environment Canada meteorological

station (Regina International Airport, Regina, Saskatchewan). The climate data,

including daily precipitation, air temperature, wind speed, net radiation, and relative

humidity were reported earlier among the field investigation details covering the time

178

span of field monitoring data and modeling duration. The total precipitation accounts for

rainfall and snow throughout the year.

7.6 Pavement Boundary Conditions

The pavement structure may experience different degrees of structural failures

(e.g. cracks) during its operations and, therefore, allows a percentage of the total surface

atmosphere applied flux to interact with the underlying material. The assumption of

having no applied flux because of the surface pavement structure is then invalid and is

not representative of the field conditions. In order to capture the variations in the VWC in

the trench backfill, a number of analyses have been completed to find out the amount of

the net flux transferred to the backfill material through the pavement structure. A

seepage ratio through the pavement (Rp) was defined as the ratio of the surface flux

below and above the pavement structure. The seepage ratio (Rp) was assigned different

values ranging from 0% to 100%. The resulting volumetric content was then compared

with the field measurements to predict the representative amount of applied flux through

the pavement structure.

7.7 Soil Temperature Analysis Results

Figures 7.4 and 7.5 compare the model outputs and the measured soil

temperature at depths of 0.45 m and 2.92 m, respectively, in the pipe trench below the

ground surface. The results show that the model captured the temperature variation for

the backfill material surrounding the pipe. The model outputs were in close agreement

with the field data near the ground surface. The model generated slightly lower

temperature values with a discrepancy of 6 °C to 9 °C. The backfill material experienced

periodic temperature changes due to the seasonal changes in air temperature. In the

soils above the pipe level, the temperature varied more in proximity to the ground

surface.

179

Soil temperatures experienced maximum and minimum peak temperatures

corresponding to the air temperature. The peak temperatures occurred at different time

durations contingent on the depth below the ground surface. In general, the shallower

the depth, the earlier the peak temperature occurred. At a depth of 0.45 m, the peak

temperatures occurred at a time that was close to the peak air temperature. However,

there was a delay of approximately 50 days for the maximum and minimum peak

temperatures to occur at the pipe level. Table 7.3 included a comparison of the change

in soil temperature from the field measurements and modeling results. The change in

soil temperature at the pipe level (2.92 m) was found to be around 13.8 oC and 13.7 oC

based on the modeling results and field measurements, respectively. This difference

increased significantly with the decrease in depth and reached approximately 42.1 oC

and 36.4 oC at a depth of 0.45 m, based on modeling results and field measurements,

respectively.

180

Figure 7.4: Soil temperature versus time in the pipe trench

Figure 7.5: Soil temperature versus time at a depth of 2.92 m in the pipe trench

Nov-05 Feb-06 May-06 Aug-06 Nov-06 Feb-07 May-07 Aug-07 Nov-07

Date (mmm-yy)

-40

-20

0

20

40

Te

mpera

ture

(OC

)

Air

Field Measuremets (0.45m)

Modelling Results (0.45 m)

Nov-05 Feb-06 May-06 Aug-06 Nov-06 Feb-07 May-07 Aug-07 Nov-07

Date (mmm-yy)

-40

-20

0

20

40

Te

mpe

ratu

re (

OC

)

Air

Field Measuremets (2.92 m)

Modelling Results (2.92m)

181

Table 7.3: Summary of the modeling results for the pipe trench for the period from April 2006 to April 2007

Parameter Depth

(m) Field Modeling

Soil Temperature Change, T (oC)

0.45 From -9.7 to 26.7

(36.4)

From -18.7 to 23.4

(42.1)

2.92 From 3.6 to 17.3

(13.7) From 5.4 to 19.2

(13.8)

182

7.8 Soil Moisture Analysis Results

The change in soil moisture due to the change in climatic conditions was modeled

over a 365-day period, between the 25th of April, 2006 and the 25th of April, 2007, for the

native clay and the backfill material. Figure 7.6 shows the variation in the VWC and

suction for the native clay versus depth with time obtained from the numerical modeling

analysis. The variation in VWC reflected well the seasonal variations in the climatic

conditions and the rainfall events. Soil layers close to the ground surface were highly

affected by the surface atmosphere conditions. The bimodal SWCC provided the model

with a good capability to capture the variation in the moisture content of the top clay

layer.

As a general trend, the VWC and soil suction varied considerably with time within

the upper two to three meters. It was clear that higher fluctuations in moisture content

occurred near the ground surface and diminished with depth. Figure 7.7 shows the

variation in VWC of the native clay deposits surrounding the pipe trench at a depth of

0.45 m from the field investigation and the numerical modeling analysis. The maximum

change in VWC was predicted to be in the order of 20.5% versus 17% from the field

measurements as summarized in Table 7.4. However, the change in VWC dropped

substantially with depth. The change in VWC at a depth of 2.92 m was predicted to be

5.5 % compared to 6 % based on the field measurements. However, the change in VWC

at a depth of 4 m was predicted to be 5 % compared to 4.5 % based on the field

measurements.

It was found that the numerical modeling results were in close agreement with field

measurements for unfrozen soil conditions (i.e. between April and November). Since the

moisture sensors in the field were not capable of measuring reliable volumetric water

contents readings during frozen soil conditions, the corresponding field measurements

have not been included. The climate data obtained from Environment Canada were

183

found to be adequate for the modeling of the soil moisture change in the field. Freezing

temperatures and snow falling started mid-November and continued as winter advanced.

Therefore, the sudden increase in the predicted VWC following the winter reflected the

thawing process of accumulated snow on the ground.

Figure 7.8 shows the effect of increasing the applied flux on the backfill in the pipe

trench. The net applied flux corresponds to full contact between the ground and the

atmosphere. The change in VWC of the backfill was predicted under different variations

in the applied surface flux as described earlier in this paper. The effect of the applied flux

decreased with the increase in soil depth. A normalized change in the VWC of more than

50% was predicted assuming that the full atmospheric induced net flux was applied to

the backfill material. Figure 7.9 shows the VWC of the backfill material at a depth of 0.45

m below the ground surface. Although the pipe trench was covered with a pavement

structure in the field, the backfill experienced changes in moisture content with time.

Table 4 shows the modeling results and field measurements for the change in VWC for

the backfill material at a depth of 0.45 m. The change in the VWC was measured in the

field to be approximately 13% at a depth of 0.45 m. It was, however, predicted that a

change of 14 % in the case of 30 % of the net atmosphere flux was applied on the pipe

trench.

In accordance with the modeling results and the field measurements, 30% of the

net flux was found to get through the pavement structure and interacted with the

underlying backfill material. In this study, the backfill material consisted of non-expansive

soil and, therefore, the variation in the soil moisture content did not result in additional

soil volume change. However, soil movements are still expected due to the swell-shrink

behavior of the clay surrounding the trench.

184

Figure ‎7.6: VWC and suction profiles at various levels for the native clay

Figure 7.7: VWC at a depth of 0.45 m versus time for the native clay

0 10 20 30 40 50

Volumetric Water Content, (%)

6

5

4

3

2

1

0

De

pth

(m

)

100 1000 10000 100000

Suction, (kPa)

6

5

4

3

2

1

0

10 Days (May, 5th)

45 Days (June, 10th)

120 Days (Aug., 25th)

150 Days (Sep., 25th)

330 Days (March, 25th)

Apr-06 Jun-06 Jul-06 Sep-06 Oct-06 Dec-06 Jan-07 Mar-07 Apr-07

Date (mmm-yy)

0

10

20

30

40

50

Vo

lum

etr

ic W

ate

r C

on

ten

t,

(%

)

Modelling Results (0.45 m)

Field Measurements

Precipitation

0

10

20

30

Pre

cip

ita

tio

n (

mm

)

185

Table 7.4: VWC change of the native clay for the period from 25th April 2006 to 25th April 2007

Depth (m) VWC Change, Δθ (%)

Field Modeling

0.45 17 20.5

2.92 6 5.5

4.0 5 4.5

186

Figure 7.8: Volumetric water content change in the pipe trench

Figure 7.9: Volumetric water content versus time in the pipe trench

0 20 40 6010 30 50

Normalized Volumetric Water Content, /o (%)

3

2.5

2

1.5

1

0.5

De

pth

(m

)

Fluxp = Assumed Net flux

through the pavement cracks = Rp * Net Atmospheric Flux

Rp= 100%

R p= 40%

R p= 30%

R p=

20%

Rp

= 1

0%

o = Initial VWC

Apr-06 Jun-06 Jul-06 Sep-06 Oct-06 Dec-06 Jan-07 Mar-07 Apr-07

Date (mmm-yy)

0

10

20

30

40

Volu

me

tric

Wa

ter

Co

nte

nt,

(%

)

Modelling Results (0.45 m)

Field Measurements (0.45 m)

0

10

20

30

Pre

cip

ita

tio

n (

mm

)

187

7.9 Soil and Pipe Displacements with Time

Figure 7.10 shows the variation in pipe displacements obtained from the field

monitoring program and the modeling analysis. The field data are the average of

displacements recorded at three locations along the pipe. The total pipe displacements

obtained from the modeling analysis were determined considering the swell-shrink

displacements for the native material surrounding the trench and the change in the net

normal stress of the backfill. It is clear that pipe displacements varied with time due to

the change in climatic conditions. Figure 7.11 shows the daily vertical displacements

predicted at the ground surface for the clay surrounding the trench and covered with

surface vegetation. The soil displacements corresponded to the precipitation and

evaporation events. Abrupt displacements up to 7.5 mm per day occurred at the ground

surface due to the change in the moisture content.

Although the pipe trench was backfilled with a non-expansive material to 3.0 m

below the ground surface and covered with a pavement structure, the pipe experienced

fluctuating movements due to the native clay surrounding the trench. The average pipe

displacement was measured in the field to vary from -2.3 mm to 5.5 mm. The predicted

pipe displacements changed from -1.8 mm to 3.1 mm. The difference between the

modeling results and the field measurements may be attributed to the fact that the

freeze-thaw analysis was not incorporated. It is also to be noted that the magnitude of

the pipe displacement is expected to be more if the pipe was to be backfilled with the

native clay material.

188

Figure 7.10: Pipe displacements versus time

Figure 7.11: Daily ground displacements (native clay)

May-06 Jun-06 Jul-06 Aug-06 Sep-06 Oct-06

Date (mmm-yy)

-10

-5

0

5

10

15

To

tal D

isp

lace

men

t (m

m)

Modelling Results

Field Measurements (Average Records along the Pipe)

0

4

8

12

16

20

Pre

cip

ita

tio

n (

mm

)

Apr-06 Jul-06 Oct-06 Jan-07 Apr-07Jun-06 Sep-06 Dec-06 Mar-07

Date (mmm-yy)

-5

-4

-3

-2

-1

0

1

2

3

4

5

6

7

8

Da

ily S

we

ll S

hri

nk D

isp

lace

me

nt

(mm

)

0

10

20

30

Pre

cip

ita

tio

n (

mm

)

Modelling Results (at the Ground Surface)

189

CHAPTER 8: SUMMARY, CONCLUSION AND FUTURE WORK

This chapter summarizes the main findings of this research study, including

engineering significance of the research, field observations, related technical

interpretations, and the results of the numerical modeling analysis.

8.1 Engineering Significance and Applications

The elevated annual breakage rates of buried pipes, especially small diameter

pipes, are a significant problem in arid and semi-arid regions, worldwide. Historical pipe

breakage rate records and field observations indicated that the pipe failure is strongly

correlated with ground movements and seasonal variation in climatic conditions. The

understanding of the correlation between pipe behavior and the most critical geo-

environmental conditions (i.e., soil moisture changes) is well developed. Finite element

methods can be used effectively to analyze this type of problems. The number of studies

on the behavior and performance of pipes under various weather conditions is still quite

limited. Consequently, the aim of this research was to enhance the current knowledge by

studying the performance of buried pipes under different field geo-environmental

conditions and the effects due to changes in unsaturated soil conditions. A better

understanding of unsaturated clays and pipe interactions can ultimately help reduce the

rate of buried pipe breakages.

8.2 Summary of Results

8.2.1 Field Investigation

Field monitoring data of the underground water main and surrounding soil provided

an understanding of the effect of soil on the behavior of buried pipes. It was found that

the soil and pipe behavior was influenced by the seasonal changes and correlated well

with the changes in the local meteorological conditions. The soil temperature, volumetric

190

water content (VWC), and suction were good indicators of the changes in the soil

behavior with the time throughout the year. The general pattern of variation in soil

temperatures was affected by air temperature and the depth below ground surface. The

volumetric water content of the backfill and the surrounding native soils at shallow

depths showed significant variations with respect to extreme weather events, especially

late March to early April as a representation of snowmelt event, and from May to June as

a representative of rainfall events. The effect of the seasonal variation decreased with

depth below ground surface.

As a general trend, clays at the pipe level of approximately 2.92 m, experienced

relatively small variations in volumetric water contents (VWCs), and were found to be in

the range of ± 5 %. The clay soil at a higher level, near the ground surface, had

prominent variations of VWC which corresponded well to the seasonal variation in the

climate conditions and the rainfall or snowmelt events. The change in VWC was found to

reach as high as 20 % at a depth of approximately 0.5 m below the ground surface. The

observations also confirmed that the wet-dry seasonal events can induce significant

displacements on the pipes. Theses displacements magnitude varied along the pipeline.

8.2.2 Load-deformation Analysis

The finite element modeling approach was established using the elastic continuum

theory. The soil-pipe system was considered as an elastic continuum with the

consideration of volume change effects, soil layering, and displacement of the pipe

relative to the soil. The performance evaluation study of the soil-pipe interaction included

the identification and modeling of a typical buried pipeline arrangement. The modeling

approach provided comparable results of maximum pipe deformations with the empirical

design equation (Spangler’s Equation). The key results of this performance evaluation

model can be summarized as follows:

191

The geometric configuration of the pipe trench, including the cover height

and trench width, significantly influenced the overall deformations of the soil-

pipe system. Increased trench width along with enhanced soil conditions

were found to provide more lateral support when the flexible pipe deforms.

The effect of the field configuration, such as the trench width, in the

calculation of pipe deformations, was found to be essential for capturing the

pipe deformations under field conditions.

Flexible pipes are subjected to higher deformations due to their low stiffness.

Pipe deformation was found to decrease with the increase in pipe diameter,

and elastic modulus of backfill materials. This effect is a result of the

improved confinement provided for the sides of the pipe.

8.2.3 Pipe Response to Unsaturated Soil Conditions

The behavior of a buried pipe in unsaturated clay soil was modeled under

saturation of the surrounding active layer. The adopted approach, through simulations of

possible scenarios, provided better understanding of the field behavior of underground

pipes. The pipe response was found to depend on the pipe depth, stiffness, and end

restraints. The main results of this model can be summarized as follows:

An increase in soil water content, corresponding to a decrease in soil suction,

resulted in significant upward displacements. The pipe displacements were highly

influenced by pipe burial depth. The upward movements of pipelines can be

notably reduced by a slight increase in burial depth.

For the case study dealing with a pipe buried at a depth of 2.7 m (a relative depth

ratio (DR) of 0.9 within the active soil zone):

o Maximum soil displacements at the ground and pipeline level were predicted

to be about 80 mm, and 10 mm, respectively. This was a result of a total

192

change in volumetric water content from 20% (close to the field condition) to

60% (close to the full saturation condition).

o Maximum pipe displacements of 2.1%, 3.7%, 5.3%, and 6.0% were reported

as results of the relative increase in volumetric water content of 5%, 10%,

15%, and 20%, respectively.

The end restraints of an underground flexible pipe (PVC pipe) affected the

distribution of displacement along the pipeline but had a minimal influence on the

maximum pipe displacement magnitude. However, the end restraints of an

underground rigid pipe (i.e. steel pipe) affected the distribution and magnitude of

displacement along the pipeline.

In case of fixed end restraints and under the same change in moisture content of

the soil, the maximum upward displacement of PVC pipes was found to be about

four times the displacement of steel pipes.

8.2.4 Soil-water Interaction in Highly Plastic Clays

A comprehensive theoretical framework for the mathematical formulations of

unsaturated soil properties was developed. The Soil Water Characteristic Curve (SWCC)

and hydraulic conductivity function for the native highly plastic clay were formulated to

reflect field conditions. The mathematical formulation of the flow through saturated-

unsaturated soil structure was developed taking into consideration the effects of soil

cracking, the use of a bimodal SWCC based on real soil moisture-suction

measurements, and the different boundary conditions. The developed bimodal SWCC

equation was based on two distinct logistic regression models fitting laboratory data sets

before and after a turning point. The volume-mass constitutive relationships were

established to provide a clear understanding of the saturated-unsaturated soil behavior

under changes in net normal stress and soil suction. In addition, the sensitivity of

193

transient seepage modeling to the characteristics of the SWCC and hydraulic

conductivity were clearly demonstrated. The main conclusions of the study can be

summarized as follows:

The soil moisture content test results demonstrated that the top two to three

meters of the clay in the study area had a water content that was less than the

plastic limit. Surficial cracks were also observed within this layer in the field. In

addition, based on the performed theoretical analysis, a cracking depth of up to

three meters would be highly expected.

Using the bimodal SWCC, the soil behavior can be modeled as a dual-porosity

medium. The developed bimodal SWCC consisted of two SWCC functions along

with their distinct parameters representing the intact part of the soil and the

cracked one.

Volume-mass constitutive relationships were established based on the bimodal

SWCC and representative consolidation results. These relationships were used to

predict the saturated-unsaturated soil properties under changes in soil suction and

net normal stress. The volume-mass constitutive relationships led to improved

predictions of the soil water interaction.

A seepage model was effectively established to simulate the development of soil

cracks as a dynamic process and a function of the soil water interaction process.

The seepage modeling study demonstrated that:

o The SWCC is a critical property affecting the modeling of the soil water

interaction and has a significant influence on the seepage modeling results.

The bimodal SWCC provided a practical technique for the simulation of the

soil water interaction for cracked soils. The use of bimodal SWCC reflected

the increase in water infiltration because of the pore space distribution

194

discrepancy in the soil matrix structure that is created by the cracks. In

addition, the bimodal SWCC resulted in a rapid change in the volumetric water

content (VWC) with time compared with an unimodal one.

o The hydraulic conductivity factor defined as a function of the soil suction

provided a reasonable demonstration of the increase in the hydraulic

conductivity of the desiccated clay layer. The increase in the soil depth within

which the hydraulic conductivity factor was applied, representing the increase

in the vertical extent of soil cracks, caused the change in soil moisture to

reach a greater depth, however, the soil took more time to reach saturation.

o Both VWC and soil suction gradually changed with time, and finally achieved a

constant value of the full saturation condition. The variation in VWC and soil

suction followed the same pattern with the increase in the applied net flux.

8.2.5 Soil-pipe-atmosphere Interaction under Field Conditions

The results of field and numerical investigations of the relationships between the

soil-pipe system and the seasonal changes in local meteorological conditions were well

demonstrated. The developed bimodal SWCC equation was based on two distinct

logistic regression models fitting laboratory data sets before and after a turning point.

The numerical analysis results were compared with the field measurements and the

following conclusions can be reported:

The numerical results were generally in agreement with field measurements. The

model was able to predict changes in soil moisture content, suction, and soil

temperature with time. Although the pipe trench was covered by a pavement

structure in the field, the backfill material showed a pattern of periodic variations in

VWCs. It was found that at least 30% of the net surface flux resulting from the daily

195

precipitation and evapotranspiration was in fact exchanged through the pavement

structure into the backfill for the simulated site.

The general pattern of variation in soil temperatures near the ground surface

closely followed that of the air temperature, but peak soil temperatures lagged

behind the peak air temperatures with depth. A lag of up to 50 days was observed

at the bottom of the pipe trench.

The change in soil temperature at the pipe level (2.92 m) was found to be around

15 oC. This difference increased significantly with the decrease of the depth and

reached 40 oC at a depth of 0.45 m.

The results of the seepage analysis were found to be representative of the

unfrozen soil conditions. Freezing temperatures and snowfall started mid-

November and continued as winter advanced. A sudden increase in VWCs

following the winter was predicted by the model reflecting the thawing process of

accumulated snow depth above ground.

The change in soil moisture for the backfill and the surrounding native soils

showed a highly fluctuating pattern at shallow depths as a result of the different

weather events. The depth of seasonal variations in VWC was found to extend

below the pipe level (to a depth of 4 m).

The pipe experienced varying displacements with time. The maximum predicted

pipe displacement was found to be in the order of 3.1 mm due to the change in soil

moisture of the native clay surrounding the pipe trench. The model predictions for

the pipe displacement were comparable to the average displacements along the

pipe that were obtained from the field measurements. The pipe displacement is

expected to increase if native clay was used to backfill the pipe trench.

196

8.3 Conclusion

With the increase in occurrence and severity of failures of underground pipe

systems in areas of expansive soil deposits, there is a great need to establish modified

design methods or construction techniques for these systems. This study utilized the

results of a field instrumentation program for the hydro-mechanical analysis of small

diameter underground pipes. It was confirmed that the pipe experiences varying

displacements with time as a result of different weather events. The model developed in

this study found that the depth of seasonal fluctuations in soil moisture condition

exceeded the normal pipe burial level (extends to a depth of approximately 4 m). In

addition, the backfill material in the pipe trench, showed a pattern of periodic variations

in moisture condition although the trench was covered by a pavement structure. The

numerical simulation predicted that at least 30% of the net surface flux, resulting from

the daily precipitation and evapotranspiration, permeated the pavement structure. From

a practical standpoint, it was concluded that the use of highly plastic clay soil as a

backfill material is not recommended for underground small diameter pipelines. In fact, it

is highly recommended that the design guidelines of buried pipelines should take into

consideration the volume change characteristics of the native soil deposits (i.e., highly

plastic clay soils) surrounding the pipe itself or the trench in case of using a granular

backfill material. The key criterion is to specify a minimum pipe burial depth for different

pipe types in order to minimize the resulting soil swell-shrink induced displacements.

197

8.4 Future Work

Based on the conclusions of the current research, the following recommendations

were made for future studies:

This thesis developed a method for studying the soil-pipe-atmosphere interaction

under local field conditions. The developed methods can be applied to other

locations analyzing the effects of different climate change scenarios.

The obtained results are encouraging and it would be of interest to extend the

scope of field testing to include freeze-thaw effects on pipelines buried in

expansive soil deposits.

More research can be implemented to characterize the water infiltration through

pavement surface discontinues/cracks and its influence on underground structures

buried in expansive soil deposits.

Future extensions of the developed model methods or applications in this thesis

are possible. One of which can incorporate the use of adaptive moving mesh

technique.

198

REFERENCES

Agricultural Technology Centre (2004) Soil moisture meters. www1.agric.gov.ab.ca. Accessed March 01, 2015.

Aitchison G (1965) Soils properties, shear strength, and consolidation. Proceedings of 6th International Conference on Soil Mechanics and Foundation Engineering, Toronto. 319-321.

Al Nakshabandi G, Kohnke H (1965) Thermal conductivity and diffusivity of soils as related to moisture tension and other physical properties. Agricultural Meteorology 2(4):271-279.

Albrecht BA, Benson CH (2001) Effect of desiccation on compacted natural clays. Journal of Geotechnical and Geoenvironmental Engineering 127(1):67-75.

Alonso EE, Gens A, Josa A (1990) A constitutive model for partially saturated soils. Géotechnique 40(3):405-430.

American Water Works Association (2002) Pipe–Design (PVC): Design and Installation vol M23. AWWA, USA. 167.

Anderson D (2010) Vertisolic soils of the Prairie region. Prairie Soils Crops 3:29-36. Anderson T (2005) Fracture mechanics: fundamentals and applications. CRC press,

Florida, USA. 610. Ayad R, Konrad J-M, Soulie M (1997) Desiccation of a sensitive clay: application of the

model CRACK. Canadian Geotechnical Journal 34(6):943-951. Ban HK (2008) Performance of buried concrete pipe under different environmental

conditions. Ph.D. Thesis, The Pennsylvania State University. 217. Baracos A, Hurst WD, Legget RF (1955) Effects of physical environment on cast-iron

pipe. American Water Works Association:1195-1206. Barbato M, Bowman M, Herbin A (2010) Performance of buried pipe installation.

FHWA/LA.10/467. Louisiana Transportation Research Center. 123. Baver LD, Gardner WH, Gardner WR (1972) Soil physics. Wiley New York. 498. Bear J (1972) Dynamics of fluids in porous media. Dover Science, New York, USA. 764. Becker BR, Misra A, Fricke BA (1992) Development of correlations for soil thermal

conductivity. International Communications in Heat and Mass Transfer 19(1):59-68.

Bishop A (1959) The effective stress principle. Teknisk Ukeblad 39:859-863. Bishop AW, Blight G (1963) Some aspects of effective stress in saturated and partly

saturated soils. Geotechnique 13(3):177-197. Black PE (2007) Revisiting the thornthwaite and mather water balance. JAWRA Journal

of the American Water Resources Associatio 43(6):1604–1605. Blight G, Williams A (1971) Cracks and fissures by shrinkage and swelling. Proc. 5th

Regional Conference for Africa on soil mechanics and foundation engineering(Angola). 1.16-11.21.

Briones A, Uehara G (1977) Soil elastic constants: II. Application to analysis of soil cracking. Soil Science Society of America Journal 41(1):26-29.

Cameron DA (2005) Analysis of buried flexible pipes in granular backfill subjected to construction traffic. University of Sydney. .

Canadian Geotechnical Society (1978) Canadian foundation engineering manual. Canadian Geotechnical Society.

Casagrande A (1937) Seepage through dams. New England Water Works 51(2):295-336.

Cassa AM (2008) A numerical investigation into the behaviour of leak openings in pipes under pressure. University of Johannesburg. South Africa.

199

Chao GK, Chao KC, Overton DD, Nelson EJ (2015) Foundation Engineering for Expansive Soils. John Wiley & Sons, USA. 416.

Chen F (1988) Foundations on expansive soils. Developments in soils geotechnical engineering vol 54. Elsevier, New York, USA.

Chen FH, Ma GS (1987) Swelling and shrinkage behavior of expansive clays. 6th International Conference on Expansive Soils, New Delhi. 127-129.

Chen L (1990) Development in geotechnical engineering, Limit analysis and soil plasticity vol 52. Elsevier Publishing, USA. 477.

Chen W-F, Baladi GY (1985) Soil plasticity: theory and implementation. Elsevier Publishing, USA. 244

Chen W-F, Liu X (2012) Limit analysis in soil mechanics. Elsevier Publishing, USA. Childs EC, Collis-George N (1950) The permeability of porous materials. Proceedings of

the Royal Society of London A: Mathematical, Physical and Engineering Sciences Series A(201):392-405.

Chou PC, Pagano NJ (1992) Elasticity: tensor, dyadic, and engineering approaches. Courier Corporation, USA. 290.

Christiansen E, Sauer EK (2002) Stratigraphy and structure of Pleistocene collapse in the Regina Low, Saskatchewan, Canada. Canadian Journal of Earth Sciences 39(9):1411-1423.

Christiansen E, Schmid B (2005) Glacial geology of southern Saskatchewan. Proceedings of the 58th Canadian Geotechnical Conference, Saskatoon, Canada.

Christiansen EA (1979) The Wisconsinan deglaciation, of southern Saskatchewan and adjacent areas. Canadian Journal of Earth Sciences 16(4):913-938.

Clark CM (1971) Expansive-soil effect on buried pipe. American Water Works Association 63(7):424-427.

Clayton CR, Xu M, Whiter JT, Ham A, Rust M (2010) Stresses in cast-iron pipes due to seasonal shrink-swell of clay soils. Proceedings of the ICE-Water management 163(3):157-162.

Colasanti RJ, Horvath JS (2010) Practical subgrade model for improved soil-structure interaction analysis: software implementation. Practice periodical on structural design and construction 15(4):278-286.

Corapcioglu MY (1984) Land subsidence — A state-of-the-art review. Fundamentals of transport phenomena in porous media. Springer. 369-444.

Corte A, Higashi A (1964) Experimental research on desiccation cracks in soil. Cold regions research and engineering lab U.S. Army Materiel Command Cold Regions Research & Engineering Laboratory, Wilmette, Illinois, USA. 76.

Croney D, Coleman J (1948) Soil thermodynamics applied to the movement of moisture in road foundations. The 7th International Congress Applied Mechanics. 163-177.

Das BM (2006) Principles of geotechnical engineering. Cengage Learning, USA. 704 Davies JT, Rideal EK (1961) Interfacial phenomena. Academic Press, New York, USA.

474. Day RW (1999) Geotechnical and foundation engineering: design and construction.

McGraw-Hill Professional, USA. de FN Gitirana Jr G, Fredlund DG (2004) Soil-water characteristic curve equation with

independent properties. Journal of Geotechnical and Geoenvironmental Engineering 130(2):209-212.

De Vries D (1952) The thermal conductivity of soil. Mededelingen van de Landbouwhoge school te Wageningen 52 (1):1-73, translated by Building Research Station (Library Communication No. 759), England.

200

De Vries D (1963) Thermal properties of soils. Physics of plant environment. North-Holland Publishing Co., Amsterdam.

Dobchuk B, Niemenen G, O'kane M, Strunk R, Barbour S (2009) Performance monitoring of soil cover field trials at the regina municipal landfill. WASTECON 2009: The Power of Solid Waste, California, USA. 22-24.

Driscoll R (1983) The influence of vegetation on the swelling and shrinking of clay soils in Britain. Geotechnique 33(2):93-105.

Durner W (1994) Hydraulic conductivity estimation for soils with heterogeneous pore structure. Water Resources Research 30(2):211-223.

Dutta SC, Roy R (2002) A critical review on idealization and modeling for interaction among soil–foundation–structure system. Computers & structures 80(20):1579-1594.

Eagle J (2009) Depth of burial for PVC pipe. J-M Manufacturing Company Inc. www.jmeagle.com. Accessed March 01, 2014.

Elkady TY (2014) Unsaturated characteristics of undisturbed expansive shale from Saudi Arabia. Arabian Journal of Geosciences 7(5):2031-2040.

Ellis J, Acton D, Clayton J (1965) The Soils of the Regina Map Area-72 I. Saskatoon, Canada.

Farouki OT (1986) Thermal properties of soils. Trans Tech Publications, USA. Fourier JBJ (1878) The analytical theory of heat. The University Press. 466 Fredlund D (1979) Second Canadian geotechnical colloquium: appropriate concepts and

technology for unsaturated soils. Canadian Geotechnical Journal 16(1):121-139. Fredlund D, Gitirana Jr G (2005) Unsaturated soil mechanics as a series of partial

differential equations. Proceedings of International Conference on Problematic Soils. 27.

Fredlund D, Morgenstern N (1976) Constitutive relations for volume change in unsaturated soils. Canadian Geotechnical Journal 13(3):261-276.

Fredlund D, Rahardjo H, Leong L, Ng CW (2001) Suggestions and recommendations for the interpretation of soil-water characteristic curves. The 14th Southeast Asian Geotechnical Conference, Hong Kong.

Fredlund D, Vanapalli S (2002) Shear Strength of Unsaturated Soils Methods of Soil Analysis: Part 4 Physical Methods. Soil Science Society of America, USA. 329-361.

Fredlund DG (1967) Comparison of soil suction and one-dimensional consolidation characteristics of a highly plastic clay. National Research Council Canada, Division of Building Research, Canada.

Fredlund DG (1975) Engineering properties of expansive clays. University of Saskatchewan Transportation and Geotechnical Group, Canada.

Fredlund DG (2000) The 1999 RM Hardy lecture: The implementation of unsaturated soil mechanics into geotechnical engineering. Canadian Geotechnical Journal 37(5):963-986.

Fredlund DG (2006) Unsaturated soil mechanics in engineering practice. Journal of geotechnical and geoenvironmental engineering 3:286-321.

Fredlund DG, Rahardjo H (1993a) An overview of unsaturated soil behaviour. Geotechnical Special Publication. American Society of Civil Engineers (ASCE). 1-31.

Fredlund DG, Rahardjo H (1993b) Soil mechanics for unsaturated soils. John Wiley & Sons. 517

Fredlund DG, Vu HQ (2003) Numerical modelling of swelling and shrinking soils around slabs-on-ground. Proceeding, Post-Tensioning Institute Annual Technical Conference, Huntington Beach, CA, USA.

201

Fredlund DG, Xing A (1994) Equations for the soil-water characteristic curve. Canadian Geotechnical Journal 31(4):521-532.

Fredlund MD, Wilson GW, Fredlund DG (2002) Use of the grain-size distribution for estimation of the soil-water characteristic curve. Canadian Geotechnical Journal 39(5):1103-1117.

Fricke BA (1997) Soil thermal conductivity: effects of saturation and dry density. University of Missouri. USA.

Gallage C, Kodikara J, Uchimura T (2013) Laboratory measurement of hydraulic conductivity functions of two unsaturated sandy soils during drying and wetting processes. Soils and Foundations 53(3):417-430.

Gardner W (1958) Some steady-state solutions of the unsaturated moisture flow equation with application to evaporation from a water table. Soil science 85(4):228-232.

Gitirana Jr GdFN (2005) Weather-related geo-hazard assessment model for railway embankment stability. University of Saskatchewan. Saskatoon, Canada.

Gould S, Boulaire F, Marlow D, Kodikara J (2009) Understanding how the Australian climate can affect pipe failure. Australia's National Water Conference and Exhibition, Ozwater'09, Australian Water Association. 1-8.

Gould SJF (2011) A study of the failure of buried reticulation pipes in reactive soils. PhD. Thesis, Monash University. Australia.

Gray DM (1973) Handbook on the principles of hydrology. Water Information Center, USA.

Griffith AA (1921) The phenomena of rupture and flow in solids. Philosophical transactions of the royal society of london Series A, containing papers of a mathematical or physical character:163-198.

Hall WJ, Newmark NM (1978) Seismic design criteria for pipelines and facilities. Journal of the Technical Councils of ASCE 104(1):91-107.

Hanson JA, Hardin BO, Mahboub K (1994) Fracture toughness of compacted cohesive soils using ring test. Journal of geotechnical engineering 120(5):872-891.

Hazen A (1892) Some physical properties of sands and gravels: with special reference to their use in filtration. 24th Annual Report. Massachusetts State Board of Health, Boston, Mass., USA. 539–556.

Hillel D (1982) Introduction to soil physics vol 364. Academic Press, New York, USA. Holtz RD, Kovacs WD (1981) An introduction to geotechnical engineering. Prentice-Hall,

USA. Houlsby G (1997) The work input to an unsaturated granular material. Géotechnique

47(1):193-196. Howard AK (1985) The USBR equation for predicting flexible pipe deflection.

Underground Plastic Pipe, ASCE. 37-55. Hu Y, Hubble D (2007) Factors contributing to the failure of asbestos cement water

mains. Canadian Journal of Civil Engineering 34(5):608-621. Hu Y, Lotfian K (2006) Field installation of an instrumented section of water main pipe on

Emerald Park Road, October 24 to 31, 2005. National Research Council Centre for Sustainable Infrastructure Research, Regina, Saskatchewan.

Hu Y, Vu H, Hubble D (2008) Water content measurement in highly plastic clay using dielectric based probes. Canadian Geotechnical Journal 47(3).

Hu Y, Vu HQ (2011) Analysis of soil conditions and pipe behaviour at a field site. Canadian Geotechnical Journal 48(6):847-866.

Hu Y, Wang D, Cossitt K, Chowdhury R (2010) AC pipe in North America: Inventory, breakage, and working environments. Journal of Pipeline Systems Engineering and Practice 1(4):156-172.

202

Hudak P, Sadler B, Hunter B (1998) Analyzing underground water-pipe breaks in residual soils. Water Engineering and Management 145(12):15-20.

Hung VQ (2003) Uncoupled and coupled solutions of volume change problems in expansive soils. Ph.D. Thesis, University of Saskatchewan. Canada.

Jennings J (1961) A revised effective stress law for use in the prediction of the behaviour of unsaturated soils. Conference on Pore Pressure and Suction in Soils, London, UK. 26-30.

Johansen O (1975) Thermal conductivity of soils. PhD. Thesis, University of Throndeim, Norway.

Johnston K, Haug M (1992) Impact of wet-dry freeze-thaw cycles on the hydraulic conductivity of glacial till. Proceedings of the 45th Canadian Geotechnical Conference, Canadian Geotechnical Society, Toronto, Canada. 26-28.

Jones LD, Jefferson I (2012) Expansive soils vol 1. Geotechnical engineering principles, problematic soils and site investigation. ICE Publishing, UK.

Kersten MS (1949) Laboratory research for the determination of the thermal properties of soils. University of Minneapolis, Minn., USA.

Khalili N, Khabbaz M (1998) A unique relationship of chi for the determination of the shear strength of unsaturated soils. Geotechnique 48(5).

Klar A, Vorster T, Soga K, Mair R (2004) Soil-pipe-tunnel interaction: comparison between winkler and elastic continuum solutions. Department of Engineering, University of Cambridge, UK. 332.

Kodikara J, Barbour S, Fredlund DG (2002) Structure development in surficial heavy clay soils: a synthesis of mechanisms. 23(3):25-40.

Köhne JM, Köhne S, Gerke H (2002) Estimating the hydraulic functions of

dual‐permeability models from bulk soil data. Water resources research 38(7):26-21-26-11.

Lambe TW (1960) A mechanistic picture of shear strength in clay. Research Conference on shear strength of cohesive soils, ASCE. 555-580.

Lambe TW, Whitman RV (2008) Soil mechanics SI version. John Wiley & Sons, USA. Lau J (1987) Desiccation cracking of clay soils. University of Saskatchewan. Saskatoon,

Canada. Lawn B, Wilshaw R (1975a) Indentation fracture: principles and applications. Journal of

Materials Science 10(6):1049-1081. Lawn B, Wilshaw T (1975b) Fracture of Brittle Solids. Cambridge, University Press, UK. Lee F-H, Lo K-W, Lee S-L (1988) Tension crack development in soils. Journal of

geotechnical engineering 114(8):915-929. Lee I, Ingles O (1968) Strength and deformation of soils and rocks. Soil mechanics,

selected topics (ed IK Lee):195-294. Leong EC, Rahardjo H (1997) Permeability functions for unsaturated soils. Journal of

Geotechnical and Geoenvironmental Engineering 123(12):1118-1126. LIoret A, Alonso E (1980) Consolidation of unsaturated soils including swelling and

collapse behaviour. Géotechnique 30(4):449-477. List RJ (1966) Smithsonian meteorological tables, 6th edn. Smithsonian Institute,

Washington, D.C. Liu Y (2005) Three dimensional finite element modeling of pavement subsurface

drainage systems. University of Kentucky. USA. Lu Y-m (1969) Swell properties of desiccated Regina clay. University of Saskatchewan.

Canada. Ma C, Eggleton RA (1999) Cation exchange capacity of kaolinite. Clays and Clay

Minerals 47(2):174-180.

203

MacKellar S, Pearson D (2003) Nationally agreed failure data and analysis methodology for water mains. Volume 1: Overview and Findings. Report Ref. No. 03/RG/05/7. UK Water Industry Research.

Manosuthikij T (2008) Studies on volume change movements in high PI clays for better design of low volume pavements. University of Texas at Arlington. USA.

Marshall T (1958) A relation between permeability and size distribution of pores. Journal of Soil Science 9(1):1-8.

Marston A (1930) The theory of external loads on closed conduits in the light of latest experiments. Bulletin No. 96. Iowa Engineering Experimental Station, Ames, Iowa, USA.

Marston A, Anderson A (1913) The theory of loads on pipes in ditches and tests of cement and clay drain tile and sewer pipe. Iowa State College Bulletin 31.

Matthews J, Nestleroth B, Condit W, Thomson J (2013) Evaluation of Opportunities to Improve Structural Inspection Capabilities for Water Mains: Large Diameter Cast Iron Pipe Cincinnati, Ohio, USA.

Matyas EL, Radhakrishna H (1968) Volume change characteristics of partially saturated soils. Géotechnique 18(4):432-448.

Miller CJ, Mi H, Yesiller N (1998) Experimental analysis of desiccation crack propagation in clay liners. Journal of the American Water Resources Association/(USA) 34:677-686.

Milly P (1984) A linear analysis of thermal effects on evaporation from soil. Water Resources Research 20(8):1075-1085.

Mitchell J (1991) Conduction phenomena: from theory to geotechnical practice. Geotechnique 41(3):299-340.

Mitchell J (1993) Fundamentals of Soil Behavior. John Wiley and Sons, Inc., New York, USA.

Mollard J, Kozicki P, Adelman T (1998) Some geological, groundwater, geotechnical and geoenvironmental characteristics of the Regina area, Saskatchewan, Canada vol 42. Urban Geology of Canadian Cities. Geological Association of Canada, Canada.

Mordak J, Wheeler J (1988) Deterioration of asbestos cement water mains, final report to the Department of the Environment. Water Research Center, Wiltshire, UK.

Morris PH, Graham J, Williams DJ (1992) Cracking in drying soils. Canadian Geotechnical Journal 29(2):263-277.

Morris R (1967) Principal causes and remedies of water main breaks. American Water Works Association:782-798.

Moser A (1990) Buried pipe design. McGraw-Hill, USA. 219. Muñoz-Carpena R, Shukla S, Morgan K (2004) Field devices for monitoring soil water

content. University of Florida Cooperative Extension Service, Institute of Food and Agricultural Sciences, USA.

Nahlawi H, Kodikara J (2006) Laboratory experiments on desiccation cracking of thin soil layers. Geotechnical & Geological Engineering 24(6):1641-1664.

Natural Resources Canada N (2010) Atlas of Canada, Age of Rocks. geogratis.gc.ca. Accessed December 01, 2014.

Nelson JD, Miller DJ (1997) Expansive soils: problems and practice in foundation and pavement engineering. John Wiley & Sons, USA.

Nelson JD, Overton DD, Durkee DB (2001) Depth of wetting and the active zone. Expansive Clay Soils and Vegetative Influence on Shallow Foundations. The Geo-Institute of the American Society of Civil Engineers, Houston, TX, USA.

Newport R (1981) Factors influencing the occurrence of bursts in iron water mains. Aqua 3:274–278.

204

Ng C, Chen R (2008) Recompacted, natural and in-situ properties of unsaturated decomposed geomaterials. Geotechnical and Geophysical Site Characterization: Proceedings of the 3rd International Conference on Site Characterization (ISC'3, Taipei, Taiwan, 1-4 April 2008). BOOK Keynote papers (258 pages)+ CD-ROM full papers, CRC Press. 117.

Ng CW, Menzies B (2007) Advanced unsaturated soil mechanics and engineering. CRC Press. 690.

Ng PCF (1994) Behaviour of buried pipelines subjected to external loading. University of Sheffield. UK. 339.

Nguyen T (1999) Solution of saturated/unsaturated seepage problems using a general partial differential equation solver. University of Saskatchewan. Canada.

Norrish K (1954) The swelling of montmorillonite. Discussions of the Faraday society 18:120-134.

Novak V, Šimåunek J, Genuchten MTv (2000) Infiltration of water into soil with cracks. Journal of Irrigation and Drainage Engineering 126(1):41-47.

Nuth M, Laloui L (2008) Effective stress concept in unsaturated soils: Clarification and validation of a unified framework. International journal for numerical and analytical methods in Geomechanics 32(7):771-801.

Nyman DJ (1984) Guidelines for the seismic design of oil and gas pipeline systems. Committee on Gas and Liquid Fuel Lifelines of the ASCE Technical Council on Lifeline Earthquake Engineering, ASCE, New York, NY, USA. 473.

Oh WT, Garga VK, Vanapalli SK (2008) Shear strength characteristics of statically compacted unsaturated kaolin. Canadian Geotechnical Journal 45(7):910-922.

Oke TR (2002) Boundary layer climates. Routledge. PDE Solutions Inc. (2014) User Guide - FlexPDE 6. www.pdesolutions.com. Accessed

July 01, 2014. Penman HL (1948) Natural evaporation from open water, bare soil and grass.

Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences, The Royal Society. 120-145.

Pentland J, Gitirana Jr G, Fredlund D (2001) Use of a general partial differential equation solver for solution of mass and heat transfer problems in geotechnical engineering. 4th Brazilian Symposium on Unsaturated Soils, NSAT 2001. 29-45.

Petroff LJ (1990) Review of the relationship between internal shear resistance and arching in plastic pipe installations. Buried plastic pipe technology. American Society for Testing and Materials (ASTM), Baltimore, USA. 266-280.

Pham HQ (2005) A volume-mass constitutive model for unsaturated soils. University of Saskatchewan. Canada.

Prevost R, Kienow K (1994) Basics of flexible pipe structural design. Journal of transportation engineering 120(4):652-671.

Pruett RJ, Murray HH (1991) Clay mineralogy, alteration history, and economic geology of the Whitemud Formation, southern Saskatchewan, Canada. Clays and Clay Minerals 39(6):586-596.

Raats P (1984) Mechanics of cracking soils. Proceedings ISSS Symposium on Water and Solute Movement in Heavy Clays Soils, The Netherlands. 23-38.

Rajani B, Tesfamariam S (2004) Uncoupled axial, flexural, and circumferential pipe soil interaction analyses of partially supported jointed water mains. Canadian Geotechnical Journal 41(6):997-1010.

Rajani B, Zhan C, Kuraoka S (1996) Pipe soil interaction analysis of jointed water mains. Canadian Geotechnical Journal 33(3):393-404.

205

Rajeev P, Chan D, Kodikara J (2012) Ground–atmosphere interaction modelling for long-term prediction of soil moisture and temperature. Canadian Geotechnical Journal 49(9):1059-1073.

Rajeev P, Kodikara J (2011) Numerical analysis of an experimental pipe buried in swelling soil. Computers and Geotechnics 38(7):897-904.

Reynolds W (1993) Unsaturated hydraulic conductivity, field measurement. Soil sampling and methods of analysis. Lewis Publishers, USA. 633-644.

Rice H, Forman S, Patry L (1959) A study of some profiles from major soil zones in Saskatchewan and Alberta. Canadian Journal of Soil Science 39(2):165-177.

Richards B (1965) Measurement of free energy of soil moisture by the psychrometric technique, using thermistors. Moisture Equilibria, London, UK. 39-46.

Richards B (1967) Moisture flow and equilibria in unsaturated soils for shallow foundations. Permeability and capillarity of soils. American Society for Testing and Materials(ASTM), USA. 4-34.

Richards LA (1931) Capillary conduction of liquids through porous mediums. Journal of Applied Physics 1(5):318-333.

Robertson GP, Sollins P, Ellis BG, Lajtha K (1999) Exchangeable ions, pH, and cation exchange capacity. Standard soil methods for long-term ecological research. Oxford University Press, New York, USA. 106-114.

Rutten MM, Steele-Dunne SC, Judge J, van de Giesen N (2010) Understanding heat transfer in the shallow subsurface using temperature observations. Vadose Zone Journal 9(4):1034-1045.

Saadawi H, Wainer G (2003) Improving the fnite element method using cell-DEVS. Summer Computer Simulation Conference, Society for Computer Simulation International. 348-353.

Saadeldin R, Henni A (2013) An overview of infrastructure planning of sustainable buried pipelines. GeoMontreal 2013, The Canadian Geotechnical Society, Montreal, Canada.

Saadeldin R, Henni A (2016) A Novel Modeling Approach for the Simulation of Soil-Water Interaction in a Highly Plastic Clay Geomechanics and Geophysics for Geo-Energy and Geo-Resources:In Press.

Saadeldin R, Hu Y, Henni A (2013a) Performance of buried pipes under different soil and loading conditions. GeoMontreal 2013, The Canadian Geotechnical Society, Montreal, Canada.

Saadeldin R, Hu Y, Henni A (2015) Numerical analysis of buried pipes under field geo-environmental conditions. International Journal of Geo-Engineering 6(1):1-22.

Saadeldin R, Hu Y, Henni A (2016) Soil-pipe-atmosphere Interaction under Field Conditions.Submitted.

Saadeldin R, Hu Y, Siddiqua S, Henni A (2013b) Response of buried pipes to unsaturated soil conditions. Pipelines and Trenchless Construction and Renewals—A Global Perspective, American Society of Civil Engineers (ASCE). 462-473.

SAI Platform (2010) Water Conservation Technical Briefs-TB6-Irrigation scheduling. www.saiplatform.org. Accessed December 05, 2015.

Saskatchewan Geological Society (2003) Geology and mineral and petroleum resources of Saskatchewan. Miscellaneous Report 2003-07 (Regina, SK: Saskatchewan Geological Survey, Saskatchewan Industry and Resources).

Satyanaga A, Rahardjo H, Leong E-C, Wang J-Y (2013) Water characteristic curve of soil with bimodal grain-size distribution. Computers and Geotechnics 48:51-61.

Skempton A (1953) The colloidal activity of clays. Selected Papers on Soil Mechanics:106-118.

206

Skempton A (1961) Effective stress in soils, concrete and rocks. Pore pressure and suction in soils. Butterworths, London, UK. 4–16.

Soil Science Society of America (2008) Glossary of soil science terms. SSSA, Madison, WI, USA.

Stearn CW, Carroll RL, Clark TH (1979) Geological evolution of North America. John Wiley & Sons, Wiley, USA. 545.

Suleiman MT (2002) The structural performance of flexible pipes. Iowa State University. Ames, Iowa 160.

Terzaghi VK (1936) The shearing resistance of saturated soils and the angle between the planes of shear. Proceedings of the 1st international conference on soil mechanics and foundation engineering, Harvard University Press Cambridge, MA. 54-56.

Thomas PJ (1998) Quantifying properties and variability of expansive soils in selected map units. Virginia Polytechnic Institute and State University. Virginia, USA.

Trenhaile AS (2004) Geomorphology: a Canadian perspective. Don Mills, Ont.: Oxford University Press, New York, USA.

Trickey SA, Moore ID (2007) Three-dimensional response of buried pipes under circular surface loading. Journal of Geotechnical and Geoenvironmental Engineering 133(2):219-223.

Uni-Bell PVC Pipe Association (1991) Handbook of PVC pipe: design and construction. Uni-Bell PVC Pipe Association, Dallas, TX, USA. 474.

Van Der Merwe D (1964) The prediction of heave from the plasticity index and the percentage clay fraction of soils. 6(6):103-107.

Vanapalli S, Pufahl D, Fredlund D (1998) The effect of stress state on the soil-water characteristic behavior of a compacted sandy-clay till. Proceedings of the 51st Canadian Geotechnical Conference, Edmonton, Canada. 81-86.

Veihmeyer FJ (1964) Evapotranspiration. Hand Book of Applied Hydrology. McGraw-Hill, New York, USA. 11.11-11.38.

Vu HQ, Fredlund DG (2004) The prediction of one-, two-, and three-dimensional heave in expansive soils. Canadian Geotechnical Journal 41(4):713-737.

Wallace WS (1948) Geology of Canada vol 3. University Associates of Canada, Toronto, Canada. 23-26.

Walsh KD, Colby CA, Houston WN, Houston SL (2009) Method for evaluation of depth of wetting in residential areas. Journal of geotechnical and geoenvironmental engineering 135(2):169-176.

Wang J-J, Zhu J-G, Chiu C, Zhang H (2007) Experimental study on fracture toughness and tensile strength of a clay. Engineering Geology 94(1):65-75.

Watkins RK, Anderson LR (1999) Structural mechanics of buried pipes. CRC Press, Boca Raton, USA. 444.

Wheeler S, Karube D (1995) State of the art report: Constitutive modelling. The First International Conference on Unsaturated Soils, Paris, France. 1323-1356.

White NF, Duke HR, Sunada DK, Corey AT (1970) Physics of desaturation in porous materials. Journal of the Irrigation and Drainage Division 96(2):165-191.

White WA (1949) Atterberg plastic limits of clay minerals. American Mineralogist 34:508-512.

Wijewickreme D (2012) Role of geotechnical engineering in assuring the integrity of buried pipeline systems. International Conference on Sustainable Built Environment (ICSBE), Kandy, Sri Lanka.

Wijewickreme D, Weerasekara L (2010) Pipeline geotechnical engineering. Encyclopedia of Life Support Systems (EOLSS). www.eolss.net/sample-chapters. Accessed July 01, 2015.

207

Wilson G, Fredlund D, Barbour S (1997) The effect of soil suction on evaporative fluxes from soil surfaces. Canadian Geotechnical Journal 34(1):145-155.

Wilson G, Fredlund M (2000) The application of knowledgebased surface flux boundary modeling. Unsaturated soils for Asia. Toll&Leong (Eds), Proceedings of the Asian Conference on Unsaturated Soils, UNSAT-Asia 2000, Singapore.

Wilson GW (1990) Soil evaporative fluxes for geotechnical engineering problems. PhD. Dissertation, University of Saskatchewan. Saskatoon, Canada.

Wilson GW, Fredlund D, Barbour S (1994) Coupled soil-atmosphere modelling for soil evaporation. Canadian Geotechnical Journal 31(2):151-161.

Winkler E (1867) Theory of elasticity and strength. Dominicus Prague (in German). Czechoslovakia.

Wulfsohn D, Adams B, Fredlund D (1996) Application of unsaturated soil mechanics for agricultural conditions. Canadian Agricultural Engineering 38(3):173-181.

Yesiller N, Miller C, Inci G, Yaldo K (2000) Desiccation and cracking behavior of three compacted landfill liner soils. Engineering Geology 57(1):105-121.

Yong RN, Warkentin BP (1966) Introduction to soil behavior. Series in Civil Engineering. Macmillan, New York, USA. 451.

Young MH, Sisson JB (2002) Tensiometry. Methods of Soil Analysis: Part 4 Physical Methods. Soil Science Society of America, USA. 575-678.

Young W, Budynas R, Sadegh A (2011) Roark's formulas for stress and strain. McGraw Hill Professional, New York, USA.

Zarghamee MS (1986) Buried flexible pipe with nonuniform soil support. Journal of transportation engineering 112(4):400-415.

Zhan C, Rajani B (1997) Load transfer analyses of buried pipe in different backfills. Journal of transportation engineering 123(6):447-453.

Zhang L, Chen Q (2005) Predicting bimodal soil–water characteristic curves. Journal of Geotechnical and Geoenvironmental Engineering 131(5):666-670.

Zhang X (2004) Consolidation theories for saturated-unsaturated soils and numerical simulations of residential buildings on shrink-swell soils. Department of Civil Engineering, Texas A&M University. College Station, TX, USA.

Zhao JQ, Kuraoka S, Baker T, Gu P, Masson J, Boudreau S, Brousseau R (1998) Durability and Performance of Gravity Pipes: A State-of-the-Art Literature Review. Institute for Research in Construction , National Research Council of Canada, Ottawa, Canada. 60.

Zotarelli L, Dukes MD, Romero CC, Migliaccio KW, Morgan KT (2014) Step by step calculation of the Penman-Monteith Evapotranspiration (FAO-56 Method). Agricultural and Biological Engineering Department, Florida Cooperative Extension Service, Institute of Food and Agricultural Sciences, University of Florida, Florida, USA.

208

APPENDIX

A.1: Modeling Load-deformation Analysis (Selected Samples)

Figure A.1.1: Vertical displacement distribution due to surface loading of 50 kPa (Case Description: Pipe diameter = 300mm, E = Es)

Figure A.1.2: Vertical displacement distribution due to surface loading of 50 kPa (Case Description: Pipe diameter = 300mm, no trench configuration, E = 5Es)

209

Figure A.1.3: Vertical displacement distribution due to surface loading of 50 kPa (Case Description: Pipe diameter = 300 mm, with trench configuration, E = 5Es)

210

A.2: Modeling Partial Saturation Analysis (Selected Results)

A.2.1: Vertical heave distribution due to change of soil suction of around 2000 kPa (Fixed-fixed end restrains)

Figure A.2.2: Vertical heave distribution due to change of soil suction of 1000 kPa (Fixed-fixed end restrains)

211

Figure A.2.3: Vertical heave distribution due to a change in soil suction of 1000 kPa (Hinged-hinged end restrains)

Figure A.2.4: Vertical heave distribution due to a change of soil suction of 2000 kPa (Hinged-hinged end restrains)

212

A.3: Soil Column Modeling (Selected Results)

Figure A.3.1: VWC mapping at elapsed time of 30 days (I=0.01 and fmax = 50)

Figure A.3.2: VWC mapping at elapsed time of 90 days (I=0.01 and fmax = 50)

213

Figure A.3.3: VWC mapping at elapsed time of 120 days

Figure A.3.4: VWC mapping at elapsed time of 240 days

214

Figure A.3.5: Suction mapping at elapsed time of 30 days

Figure A.3.6: Suction mapping at elapsed time of 90 days

215

Figure A.3.7: Suction mapping at elapsed time of 120 days

Figure A.3.8: Suction mapping at elapsed time of 240 days

216

A.4: Field Site Modeling (Selected Results)

Figure A.4.1: VWC mapping at elapsed time of 5 days

Figure A.4.2: VWC mapping at elapsed time of 30 days

217

Figure A.4.3: VWC mapping at elapsed time of 140 days

Figure A.4.4: VWC mapping at elapsed time of 210 days

218

A.5: Field Site Location Map, Stratigraphic Cross Sections and Selected Photographs

Figure A.5.1: Site location Map

Emerald Park Road, Regina, Saskatchewan.

Project: Field Installation of an Instrumented Section of Water Main Pipe on Emerald

Park Road, October 24 to 31, 2005.

Center for Sustainable Infrastructure Research (CSIR). National Research Council

Canada.

Site Location

219

Fig

ure

A.5

.2: S

tratig

raph

ic c

ross s

ectio

n a

long

or n

ea

r 4th A

ve. (E

ast to

West) R

egin

a, S

askatc

hew

an

afte

r

(Fre

dlu

nd

, 1975)

220

Fig

ure

A.5

.3: S

tratig

raph

ic c

ross s

ectio

n a

long

or n

ea

r Alb

ert S

t. (No

rth to

So

uth

), Re

gin

a, S

askatc

hew

an a

fter

(Fre

dlu

nd

, 197

5)

221

Fig

ure

A.5

.4: S

tratig

raph

ic c

ross s

ectio

n a

long

or n

ea

r Vic

toria

Ave

. (Ea

st to

We

st), R

egin

a, S

askatc

hew

an

(Interpreted fro

m W

ater S

ecurity

Agency’s w

ate

r we

ll drille

r's re

ports

)

222

Figure A.5.5: Excavation of the pipe trench

Figure A.5.6: Old pipe configuration

223

Figure A.5.7: Placement of side support system

Figure A.5.8: Placement of the instrumented pipe

224

Figure A.5.9: Placement of soil pressure cells

Figure A.5.10: Placement of soil moisture sensors

225

Figure A.5.11: View of the site after the placement of the instrumented pipe

Figure A.5.12: View of the site showing the pavement cracks