Fatahi 2013 02whole
Fatahi 2013 02whole
By
Behnam Fatahi
Doctor of Philosophy
March 2013
1
CERTIFICATE OF AUTHORSHIP/ORIGINALITY
I certify that this thesis has not previously been submitted for a degree nor has it been
submitted as part of requirements for a degree except as fully acknowledge within the text.
I also certify that the thesis has been written by me. Any help that I have received in my
research work and the preparation of the thesis itself has been acknowledged. In addition, I
certify that all information sources and literature used are indicated in the thesis.
Signature of candidate
Behnam Fatahi
i
Abstract
Many closed municipal solid waste (MSW) landfills are located near urban areas, even
though originally established away from residential or commercial communities.
Construction on top of closed landfills is generally a challenging task due to complex
behaviour of creep, settlement and weak shear strength of waste materials. There is a high
prospective to reuse these sites for redevelopment in spite of potential risk for human health
and environment. The deep dynamic compaction technique is a common ground
improvement technique due to its relatively economical and easy application for landfill
sites. With deep dynamic compaction, large voids reduce and afterward other techniques
such as cement, fly ash or lime grouting can further reduce the remaining smaller voids.
Numerous studies have been conducted to treat and stabilise different types of problematic
soils using fly ash with combination of lime. However, there is no comprehensive research
on improvement of physical properties of MSW landfills using chemical admixtures such
as fly ash and lime.
This study presents the experimental and numerical results of employing fly ash (class F)
and quicklime (calcium oxide) in stabilisation of municipal solid wastes. The waste
materials, used in this study, were collected from a closed landfill in the south-west of
Sydney. The samples were prepared by integrating MSW, with a mixture of fly ash-
quicklime with a ratio of 3:1 in percentages of 5, 10, 15 and 20 of fly ash by dry weight of
the MSW. An array of experimental tests has been conducted on treated and untreated
MSW samples including sieve analysis, Atterberg limits, compaction, permeability, large
direct shear, unconfined compressive strength and consolidated-drained triaxial tests.
Results of this investigation are evidence for a significant improvement in geotechnical
properties of MSW materials, mixed with fly ash and quicklime. It has been found that the
chemical stabilisation effectively increases the maximum dry density, the compressive
strength, the shear strength parameters, the stiffness and the brittleness index, while
decreases the compressibility, the permeability coefficient and the optimum moisture
content of the MSW.
ii
It has been quantified that by increasing fly ash-quicklime admixtures from 0 to 26.7% (0
to 20% fly ash) the internal friction angle increased from 29o to 39o and the cohesion
intercept increased from 11 kPa to 30 kPa. Under an effective confining pressure of 300
kPa, the peak strength, the brittleness index and the Young’s modulus at failure increased
from 600 kPa to 1150 kPa, 0.13 to 0.35 and 5.5 MPa to 28 MPa, respectively, by addition
of 26.7% fly ash-quicklime admixture. The coefficient of permeability for untreated
specimen was 6.2×10-8 m/s and it was reduced to 3.2×10-8 m/s for specimens mixed with
26% fly ash-quicklime (under average confining pressure of 250 kPa). The compression
and the secondary compression indices decreased from 0.33 to 0.23 and 0.052 to 0.033,
respectively. Moreover, increasing the curing time enhanced the unconfined compressive
strength, the friction angle, the cohesion and the preconsolidation pressure of the treated
specimens, whereas no change in the permeability coefficient, the primary compression
index and the secondary compression index were observed. The findings of this study may
facilitate the calculations of the bearing capacity and settlement as well as the slope
stability analysis of chemically treated closed landfill sites.
A finite element program, PLAXIS version 9, has been used to evaluate the settlement of
the untreated and chemically treated landfill layers for 10 and 20 years after applying
surcharge loads such as the traffic load. The effects of depth of stabilisation and the fly ash-
quicklime content on vertical and horizontal displacements of the model have been
investigated. Treated and untreated MSW parameters, used for the model, have been
obtained from the results of the extensive laboratory program performed in this study. The
numerical results indicated that treatment of MSW with fly ash-quicklime reduced the
vertical displacement of the model under traffic load at the midpoint below the
embankment. This reduction is more pronounced with higher fly ash-quicklime contents
and deeper improvement of layers. For depths of 3m, 6m, 9m, 12m and 24m of the landfill
improved with 26.7% fly ash-quicklime, the vertical settlements at the centreline of the
embankment, 10 years after applying traffic load, were reduced by 20%, 32%, 40%, 46%
and 58%, respectively. Horizontal displacements of the landfill model also significantly
reduced in sections below the toe of the embankment, under traffic load. The reduction in
horizontal displacements is more pronounced with improvement into deeper layers.
iii
Acknowledgements
This PhD project could not have been possible without the support provided by numerous
people. In particular, I would like to express my deepest appreciation and gratitude to:
Dr Behzad Fatahi, for his unfailing assistance, guidance and support over the past few
years. His brilliant and sharp mind combined with his extensive technical knowledge,
experience and dedication contributed largely to the success of this project.
The UTS laboratory staffs, Rami Haddad, David Hooper, Antonio Reyno, David Dicker,
Peter Brown, Laurence Stonard and Richard Turnell, for their extensive assistance in
conducting the experimental works. Special thank goes to Antonio Reyno, for his
remarkable help in all technical matters conducting experimental testing in the soil
laboratory.
The administrative and the support staff at UTS Faculty of Engineering and IT, Phyllis
Agius, Craig Shuard, Van Lee and the IT support team for performing an excellent job in
keeping the show running.
Sydney councils, particularly Bankstown, Fairfeild, Hornsby shire, Lane cove and
Burwood Councils for their support in visiting the landfills and special thanks goes to
Oliver Brown for permission of sampling from former Bankstown landfill site.
iv
Dedication
I would like to dedicate this Doctoral dissertation to my family, particularly my wife
Fouzieh Lotfi for her love, understanding and the sacrifice she has had to support my study,
my father Dr Bahram Fatahi and my mother Monir Kheirandish for instilling in me the
wisdom needed to complete this PhD project.
v
List of Publications
x Khabbaz, H. & Fatahi, B. (2012), “Stabilisation of Closed Landfill Sites by Fly Ash
Using Deep Mixing Method”, Grouting and Deep Mixing 2012, Louisiana, USA,
February 2012 in Proceedings of the Fourth International Conference on Grouting
and Deep Mixing 2012, ASCE, USA, pp. 417-426.
x Fatahi, B., Khabbaz, H. and Fatahi, B. (2012). “Mechanical Characteristics of Soft Clay
Treated with Fibre and Cement”, Geosynthetics International, 19(3), 252 –262.
x Fatahi, B. and Khabbaz, H. (2013). “Influence of Fly Ash and Quicklime Addition on
Behaviour of Municipal Solid Wastes”, Journal of Soils and Sediments
(accepted/in-press) DOI 10.1007/s11368-013-0720-4
x Fatahi, B. and Khabbaz, H. (2013). “Influence of Fly Ash and Quicklime Addition on
Permeability of Municipal Solid Wastes”, Journal of Soils and Sediments
(Submitted)
x Fatahi, B., Le, T. M., Fatahi, B. Khabbaz, H. (2013). “Shrinkage of Soft Clay Treated
with Cement and Geofibres”, Geotechnical and Geological Engineering:
International Journal (accepted/in-press) DOI 10.1007/s10706-013-9666-y
x Fatahi, B., Fatahi, B. Le, T. M., Khabbaz, H. (2013). “Small-Strain Properties of Soft
Clay Treated with Fibre and Cement”, Geosynthetics International (submitted/under
review)
vi
Table of Contents
Abstract .................................................................................................................................. ii
Acknowledgements ............................................................................................................... iv
List of Publications ............................................................................................................... vi
List of Figures ..................................................................................................................... xiii
List of Tables...................................................................................................................... xxii
CHAPTER 1 ......................................................................................................................... 1
1 Introduction .............................................................................................................. 1
1.1 Introduction ........................................................................................................... 1
1.2 Statement of the problems ..................................................................................... 2
1.3 Research Background ............................................................................................ 3
1.3.1 Improvement Techniques ................................................................................. 3
1.3.2 Chemical Stabilisation ..................................................................................... 3
1.4 Research Scope and Objectives ............................................................................. 4
1.5 Outline of Thesis ................................................................................................... 6
CHAPTER 2 ......................................................................................................................... 8
2 Literature Review ..................................................................................................... 8
2.1 Waste Mechanics ................................................................................................... 8
2.2 Waste Components ................................................................................................ 9
2.3 Landfill Components ............................................................................................. 9
2.3.1 Liner System .................................................................................................... 9
2.3.2 Leachate Collection And Removal System ................................................... 10
2.3.3 Gas Collection and Control System ............................................................... 10
2.3.4 Final Cover System ........................................................................................ 10
2.3.5 Composite Liners ........................................................................................... 11
2.4 Unit Weight of Municipal Solid Waste ............................................................... 11
2.4.1 Introduction .................................................................................................... 11
2.4.2 Importance of MSW Unit Weight in Engineering Analyses ......................... 12
2.4.3 Methods Used to Estimate MSW Unit Weight .............................................. 14
2.4.4 Unit Weight Model for Municipal Solid Waste ............................................. 15
vii
2.4.5 Effect of Compaction on Unit Weight of MSW ............................................ 16
2.4.6 Effect of Depth on the Unit Weight of MSW ................................................ 18
2.4.7 Effect of Moisture Content on Unit Weight of MSW .................................... 18
2.5 Compressibility.................................................................................................... 19
2.5.1 Introduction .................................................................................................... 19
2.5.2 Mechanism of Waste Settlement.................................................................... 19
2.5.3 Primary Compression ..................................................................................... 21
2.5.4 Completion of Primary Settlement Time ...................................................... 21
2.5.5 Secondary Compression ................................................................................. 22
2.5.6 Total Compression ......................................................................................... 23
2.5.7 Influencing Factors......................................................................................... 23
2.5.8 Settlement Estimation Methods for MSW Landfills ...................................... 24
2.5.8.1 Sowers method....................................................................................... 24
2.5.8.2 Rheological model ................................................................................. 27
2.5.8.3 Power creep model ................................................................................ 27
2.5.8.4 Hyperbolic function model .................................................................... 27
2.5.9 Categories of Secondary Settlement .............................................................. 28
2.5.9.1 Settlement under self-weight: ................................................................ 28
2.5.9.2 Settlement under external loads: ............................................................ 29
2.6 Shear Strength of MSW ....................................................................................... 34
2.6.1 Introduction .................................................................................................... 34
2.6.2 Background .................................................................................................... 35
2.6.3 Effect of particles orientation ......................................................................... 44
2.6.4 Effect of Normal stress................................................................................... 45
2.6.5 Back Calculations from Field Cases .............................................................. 47
2.6.6 Limitations ..................................................................................................... 49
2.7 Hydraulic Conductivity ....................................................................................... 50
2.7.1 Introduction .................................................................................................... 50
2.7.2 Saturated Flow ............................................................................................... 51
2.7.3 Background .................................................................................................... 52
2.7.4 Influence of Effective Stress and Waste Density on Hydraulic Conductivity 54
viii
2.7.5 Effect of Waste Degradation .......................................................................... 56
2.7.6 Waste Anisotropy ........................................................................................... 57
2.8 Improvement Techniques for MSW landfills ...................................................... 59
2.9 Chemical Stabilisation ......................................................................................... 59
2.9.1 Lime Stabilisation .......................................................................................... 60
2.9.2 Fly ash Stabilisation ....................................................................................... 61
2.9.3 Lime/Fly ash Stabilisation ............................................................................. 65
2.10 Summary.............................................................................................................. 75
CHAPTER 3 ....................................................................................................................... 79
3 Geotechnical Characterisation of the Collected MSW Samples ........................ 79
3.1 Introduction ......................................................................................................... 79
3.2 Classes of Landfill ............................................................................................... 79
3.3 Number of Landfills by Type of Waste Acceptance in Australia ....................... 80
3.4 Sampling Permission ........................................................................................... 81
3.5 Sites Visits ........................................................................................................... 81
3.5.1 Blackman Park – Lane Cove Council ............................................................ 81
3.5.2 Wangal Park – Burwood Council .................................................................. 82
3.5.3 Dartford Road Landfill, Thornleigh – Hornsby Shire Council ...................... 83
3.5.4 Brenan Park – Fairfield City Council............................................................. 84
3.6 Bankstown Landfill Site and Test pits Locations ................................................ 85
3.7 In-situ MSW Characterisation in Bankstown Former Landfill ........................... 87
3.8 Primary Geotechnical MSW Characterisation .................................................... 88
3.9 Excavations and Samples Collection ................................................................... 88
3.10 Geotechnical Characterisation of MSW .............................................................. 91
3.11 Moisture Content and Organic Content of MSW – Material Loss ...................... 92
3.11.1 Background .................................................................................................... 92
3.11.2 Definitions ...................................................................................................... 92
3.11.3 Material Loss Testing Procedure ................................................................... 93
3.11.4 Results of Moisture Content and Organic Content for this Research ............ 94
3.12 Waste Composition ............................................................................................. 95
3.13 Sieve Analysis and Characterisation of Smaller Fraction ................................... 95
ix
3.14 Summary.............................................................................................................. 99
CHAPTER 4 ..................................................................................................................... 101
4 Materials, Sample Preparation and Laboratory Testing Program ................. 101
4.1 Introduction ....................................................................................................... 101
4.2 Materials ............................................................................................................ 101
4.2.1 Municipal Solid Waste ................................................................................. 101
4.2.2 Fly ash .......................................................................................................... 102
4.2.2.1 Chemical Composition of Fly Ash ...................................................... 104
4.2.3 Lime ............................................................................................................. 104
4.2.3.1 Activation of Fly ash with Lime .......................................................... 106
4.3 Compaction Tests .............................................................................................. 107
4.4 Particle Size Limitation ..................................................................................... 109
4.5 Ratio of Fly Ash and Quicklime in Soil Stabilisation ....................................... 110
4.6 Mixing of Materials ........................................................................................... 110
4.7 Sample Preparation ............................................................................................ 112
4.8 Experimental Program ....................................................................................... 113
4.8.1 Unconfined Compressive Strength Tests ..................................................... 113
4.8.2 Direct Shear Tests ........................................................................................ 117
4.8.2.1 Description of the Large Direct Shear Box Device ............................. 118
4.8.2.2 Large Direct Shear Testing Program ................................................... 120
4.8.3 Hydraulic Conductivity ................................................................................ 121
4.8.3.1 Triaxial Hydraulic Conductivity Procedures for Recompacted .......... 121
4.8.3.2 Constant Head Hydraulic Conductivity Tests ..................................... 123
4.8.4 Triaxial Test ................................................................................................. 126
4.8.4.1 Triaxial and Consolidation Specimen Preparation and Testing
Procedures ......................................................................................................... 126
4.9 Summary............................................................................................................ 131
CHAPTER 5 ..................................................................................................................... 133
5 Experimental Results and Discussion ................................................................. 133
5.1 Introduction ....................................................................................................... 133
5.2 Unconfined Compressive Tests ......................................................................... 133
x
5.2.1 Effect of the Content of Fly ash-Quicklime ................................................. 133
5.2.2 Effect of Curing Time .................................................................................. 136
5.3 Direct Shear Test ............................................................................................... 139
5.3.1 Direct Shear Tests Results ........................................................................... 139
5.3.2 Effect of Curing Time .................................................................................. 139
5.3.3 Stress-Strain Behaviour ................................................................................ 141
5.4 Permeability Test ............................................................................................... 148
5.4.1 Effect of Fly ash-Quicklime Content ........................................................... 148
5.4.2 Effect of Curing Time .................................................................................. 149
5.4.3 Permeability Results from Consolidation Tests Using Triaxial Apparatus . 150
5.4.3.1 Coefficient of Consolidation ............................................................... 150
5.4.3.2 Hydraulic Conductivity ....................................................................... 151
5.4.3.3 Permeability Change Index.................................................................. 155
5.5 Consolidated Drained Triaxial Test Results ...................................................... 158
5.5.1 Stress-Strain Behaviour ................................................................................ 158
5.5.2 Shear Strength Parameters ........................................................................... 164
5.5.3 Modulus of Elasticity ................................................................................... 166
5.5.4 Brittleness Index........................................................................................... 168
5.5.5 Primary Compression ................................................................................... 169
5.5.6 Secondary Compression ............................................................................... 176
5.6 Summary............................................................................................................ 180
CHAPTER 6 ..................................................................................................................... 182
6 Numerical Analysis to Predict the Settlement of Closed Landfills .................. 182
6.1 Introduction ....................................................................................................... 182
6.2 Finite Element Modeling ................................................................................... 183
6.3 Mesh Generation and Boundary Conditions ..................................................... 183
6.4 Adopted Material Models .................................................................................. 185
6.5 Material Parameters ........................................................................................... 191
6.6 Analysis Type .................................................................................................... 194
6.7 Results and Discussion ...................................................................................... 195
6.7.1 Vertical Settlement 10 Years after Applying Traffic Load .......................... 195
xi
6.7.2 Vertical Settlement 20 Years after Applying the Traffic Load: ................... 199
6.7.3 Horizontal Displacement 10 Years after Applying the Traffic Load ........... 204
6.8 Summary............................................................................................................ 208
CHAPTER 7 ..................................................................................................................... 209
7 Summary and Conclusions .................................................................................. 209
7.1 Summary............................................................................................................ 209
7.2 Concluding Remarks ......................................................................................... 211
7.3 Recommendation for Further Study .................................................................. 216
References ......................................................................................................................... 218
Appendix ........................................................................................................................... 231
xii
List of Figures
xiii
Figure 2.18 Direct shear strength of Tri-Cities landfill MSW: (a) curved strength envelope
for samples with varying waste composition, and (b) decrease in secant friction angle with
increasing normal stress assuming c = 15 kPa (Bray et al. 2009.) ..................................... 46
Figure 2.19 Response of MSW with 62% less than 20 mm material in direct shear testing
loaded at two displacement rates (Bray et al. 2009). .......................................................... 47
Figure 2.20 Shear strength envelope (Kavazanjian et al. 1995). ........................................ 48
Figure 2.21 Vertical hydraulic conductivity against (a) the logarithm of the vertical
effective stress in first loading; (b) the drainable porosity; and (c) density, for four waste
types (data from Beaven 2000 and Hudson et al. 2001) ...................................................... 54
Figure 2.22 Summary of relationships between vertical hydraulic conductivity and waste
dry density.(Beaven et al. 2011) ........................................................................................... 56
Figure 2.23 kh : kv versus applied stress for sample AG1. (Beaven et al. 2011) ................ 57
Figure 2.24 kh : kv versus applied stress for sample DN1. (Beaven et al. 2011) ............... 58
Figure 2.25 Compaction curves of blended CCR-stabilised clay (Horpibulsuk et al. 2012)
.............................................................................................................................................. 67
Figure 2.26 Strength development in blended CCR-stabilized clay at OWC and 5% binder
for different CCR:Fly ash ratios (Horpibulsuk et al. 2012) ................................................ 67
Figure 2.27 Strength development in blended CCR-stabilized clay at OWC and 10% binder
for different replacement ratios (Horpibulsuk et al. 2012) .................................................. 68
Figure 2.28 Unconfined compressive strength of soils stabilised with a blend of 15% coal
fly ash and 3% limestone dust relative to those of control soils (Brooks et al. 2011) ......... 69
Figure 2.29 UCS of lime and CFA mixes (Singh et al. 2010) .............................................. 69
Figure 2.30 Resilient modulus for stabilised soils (σd = 42 kPa and σ3=13 kPa) (Singh et
al. 2010) ............................................................................................................................... 70
Figure 2.31 Water content versus compressive strength (Han-bing et al. 2009). ............... 70
Figure 2.32 Water content versus deformation modulus (Han-bing et al. 2009). ............... 71
Figure 2.33 Relationships between unconfined compressive strength and fly ash content
(Shao et al. 2008) ................................................................................................................. 71
Figure 2.34 Dry density versus lime (%) at different % of fly ash (Kumar et al. 2007) ...... 72
Figure 2.35 Optimum moisture content (%) versus lime (%) at different % of fly ash
(Kumar et al. 2007) .............................................................................................................. 73
Figure 2.36 Variation of unconfined compressive strength with % of lime for different % of
fly ash (after 7 days curing) (Kumar et al. 2007) ................................................................ 74
Figure 2.37 Variation of unconfined compressive strength with % of lime for different % of
fly ash (after 14 days curing) (Kumar et al. 2007) .............................................................. 74
Figure 2.38 Variation of unconfined compressive strength with % of lime for different % of
fly ash (after 28 days curing) (Kumar et al. 2007) .............................................................. 74
xiv
Figure 3.1 Blackman park - waterlogged after rainfall....................................................... 82
Figure 3.2 Wangal park - Wetland Surrounded by Fence .................................................. 83
Figure 3.3 Dartford road landfill - golf driving range under construction ......................... 84
Figure 3.4 Brenan Park ....................................................................................................... 84
Figure 3.5 Location of the former Bankstown landfill (courtesy of Google Maps)............. 85
Figure 3.6 Plan view of the former Bankstown landfill and test pits locations ................... 86
Figure 3.7 Location of TP1 in former Bankstown landfill ................................................... 86
Figure 3.8 Digging the test pit using a backhoe .................................................................. 89
Figure 3.9 Side view of the excavated test pit ...................................................................... 89
Figure 3.10 View of the primary waste components ............................................................ 89
Figure 3.11 View of the primary waste components ............................................................ 90
Figure 3.12 Drums filled from the first borehole ................................................................. 90
Figure 3.13 Filling drums with excavated waste materials using a backhoe ...................... 90
Figure 3.14 Placement of representative samples of waste in the drums ............................ 91
Figure 3.15 Placement of the soil fraction in plastic bags and in the drum ........................ 91
Figure 3.16 View of furnace heated at 440 degrees Celsius................................................ 93
Figure 3.17 View of furnace used for estimation of organic content of waste material ...... 94
Figure 3.18 Samples of less than 19 mm material in furnace.............................................. 94
Figure 3.19 Waste material included all particle size before sieving.................................. 96
Figure 3.20 View of the larger than 19 mm waste material (retained on the sieves) .......... 96
Figure 3.21 Dry sieve analyses of finer than 19 mm fraction.............................................. 97
Figure 3.22 Sieve analysis process ...................................................................................... 97
Figure 3.23 Processing of the waste through the 19 mm sieve. ........................................... 97
Figure 3.24 Dry sieve analyses of finer than 9 mm fraction................................................ 98
Figure 3.25 Remaining fraction of processed material on different sieves ......................... 98
Figure 3.26 Small wood particles on one of the sieves ........................................................ 99
Figure 4.1 Eraring Fly Ash used as an additive to waste material ................................... 103
Figure 4.2 View of quicklime used as an additive to waste material ................................. 105
Figure 4.3 View of compaction test equipments ................................................................ 107
Figure 4.4 Filling the compaction mould with smaller than 9mm decomposed waste
material .............................................................................................................................. 108
Figure 4.5 Compacted waste material with mould after compaction completed .............. 108
Figure 4.6 Waste material included all particle size before sieving.................................. 109
xv
Figure 4.7 Smaller than 19 mm waste material before mixing with additives .................. 111
Figure 4.8 Mixture procedure of waste material with fly ash and quicklime .................... 111
Figure 4.9 View of sample of waste material after mixing with fly ash and quicklime ..... 111
Figure 4.10 Moisture content absorption ratio of treated MSW samples ......................... 112
Figure 4.11 Moisture content of treated MSW samples ..................................................... 112
Figure 4.12 The split mould used for the unconfined compressive strength test sample
preparation......................................................................................................................... 114
Figure 4.13 View of prepared samples for unconfined compressive strength test with
various fly ash – quicklime content .................................................................................... 115
Figure 4.14 View of specimen with 13% additives (10% fly ash + 3.3% quicklime) ........ 116
Figure 4.15 A specimen with 20% additives before unconfined compressive strength test
............................................................................................................................................ 116
Figure 4.16 A specimen with 13% additives after completion of unconfined compressive
strength test ........................................................................................................................ 117
Figure 4.17 A specimen with 26% additives after completion of unconfined compressive
strength test ........................................................................................................................ 117
Figure 4.18 View of the direct shear device used for the performance of the tests ........... 119
Figure 4.19 The crane used for lifting the normal force and applying to the sample ....... 119
Figure 4.20 3 LVDTs for the measurement of the horizontal and vertical displacements 119
Figure 4.21 The hammer used for the specimen compaction in shear box at the University
of Technology Sydney......................................................................................................... 120
Figure 4.22 A compacted waste material in a mould prepared for the permeability test . 124
Figure 4.23 Using high permeable fibre to prevent sample particles movement into
drainage layers .................................................................................................................. 125
Figure 4.24 Drainage layer on top (and bottom) of the sample ........................................ 125
Figure 4.25 Prepared sample for permeability tests ......................................................... 125
Figure 4.26 An experimental setup for permeability tests ................................................. 126
Figure 4.27 The triaxial sample preparation mould and the extruder .............................. 130
Figure 4.28 The automated triaxial apparatus during a CD test on a MSW sample ........ 130
Figure 4.29 A snapshot of a failed waste sample after completion of a CD Triaxial test . 131
Figure 5.1 Unconfined compressive strength of untreated and treated MSW specimen with
different fly ash-quicklime contents ................................................................................... 134
Figure 5.2 Engineering properties of organic soil–fly ash mixtures as a function of fly ash
percentage in the mixture (Tastan et al. 2011) .................................................................. 136
Figure 5.3 Effect of curing time on unconfined compressive strength of MSW specimens 137
xvi
Figure 5.4 Unconfined compressive strength of fly ash versus curing period for unsoaked
specimens with varying percentages of lime (Ghosh and Subbarao 2007) ....................... 138
Figure 5.5 Unconfined compressive strength of fly ash versus curing period for soaked and
unsoaked specimens with varying percentages of lime and (a) 0.5%; (b) 1.0% gypsum
(Ghosh and Subbarao 2007) .............................................................................................. 138
Figure 5.6 Shear strength envelope from the results of direct shear test on untreated and
treated MSW specimens with different fly ash-quicklime contents .................................... 140
Figure 5.7 Effect of curing time on shear strength envelope of treated MSW specimens with
20% fly ash-quicklime ........................................................................................................ 140
Figure 5.8 Shear strength of untreated MSW specimen in direct shear test under different
normal stresses ................................................................................................................... 142
Figure 5.9 Vertical displacement vs. horizontal displacement for untreated MSW specimen
under different normal stresses .......................................................................................... 142
Figure 5.10 Shear strength of treated MSW specimen mixed with 13.3% fly ash-quicklime
in direct shear test under different normal stresses ........................................................... 143
Figure 5.11 Vertical displacement vs. horizontal displacement for treated MSW specimen
mixed with 13.3% fly ash-quicklime under different normal stresses ................................ 143
Figure 5.12 Shear strength of treated MSW specimen mixed with 20% fly ash-quicklime in
direct shear test under different normal stresses ............................................................... 144
Figure 5.13 Vertical displacement vs. horizontal displacement for treated MSW specimen
mixed with 20% fly ash-quicklime under different normal stresses ................................... 144
Figure 5.14 Shear strength of treated MSW specimen mixed with 26.7% fly ash-quicklime
in direct shear test under different normal stresses ........................................................... 145
Figure 5.15 Vertical displacement vs. horizontal displacement for treated MSW specimen
mixed with 26.7% fly ash-quicklime under different normal stresses ................................ 145
Figure 5.16 Large-scale DS test results on MSW from Canada (Landva and Clark, 1990).
............................................................................................................................................ 146
Figure 5.17 Recommended static shear strength of MSW based primarily on direct shear
tests and field observations of static slope stability (Bray et al. 2009). ............................ 147
Figure 5.18 Results of in situ direct shear tests on MSW (Caicedo et al. 2002a). ............ 148
Figure 5.19 Coefficient of permeability of untreated and treated MSW specimens under 7 ,
28 and 93 days curing time ................................................................................................ 149
Figure 5.20 Coefficient of consolidation of MSW specimens for different fly ash-quicklime
contents under various effective confining pressures ........................................................ 151
Figure 5.21 Coefficient of permeability of MSW specimens for different fly ash-quicklime
contents under various effective confining pressures ........................................................ 154
Figure 5.22 Coefficient of permeability of MSW specimens for different fly ash-quicklime
contents under various effective confining pressures ........................................................ 154
xvii
Figure 5.23 Void ratio-permeability relationship of untreated MSW specimen ................ 155
Figure 5.24 Void ratio-permeability relationship of treated MSW specimen with 6.7% fly
ash-quicklime ..................................................................................................................... 156
Figure 5.25 Void ratio-permeability relationship of treated MSW specimen with 13.3% fly
ash-quicklime ..................................................................................................................... 156
Figure 5.26 Void ratio-permeability relationship of treated MSW specimen with 20% fly
ash-quicklime ..................................................................................................................... 157
Figure 5.27 Void ratio-permeability relationship of treated MSW specimen with 26.7% fly
ash-quicklime ..................................................................................................................... 157
Figure 5.28 Void ratio-permeability relationship of untreated and treated MSW specimen
with different fly ash-quicklime contents............................................................................ 158
Figure 5.29 Stress-strain-volumetric response of untreated MSW specimens .................. 160
Figure 5.30 Stress-strain-volumetric response of treated MSW specimens with 6.7% fly
ash-quicklime content......................................................................................................... 160
Figure 5.31 Stress-strain-volumetric response of treated MSW specimens with 13.3% fly
ash-quicklime content......................................................................................................... 161
Figure 5.32 Stress-strain-volumetric response of treated MSW specimens with 20% fly ash-
quicklime content ............................................................................................................... 161
Figure 5.33 Stress-strain-volumetric response of treated MSW specimens with 26.7% fly
ash-quicklime content......................................................................................................... 162
Figure 5.34 Stress-strain responses of treated and untreated MSW specimens with different
percentages of fly ash-quicklime at effective confining pressure of 100 kPa .................... 162
Figure 5.35 Stress-strain responses of treated and untreated MSW specimens with different
percentages of fly ash-quicklime at effective confining pressure of 200 kPa .................... 163
Figure 5.36 Stress-strain responses of treated and untreated MSW specimens with different
percentages of fly ash-quicklime at effective confining pressure of 300 kPa .................... 163
Figure 5.37 Stress–strain response of fly ash, 7 and 28 days curing (Ghosh and Subbarao
2007) .................................................................................................................................. 164
Figure 5.38 Stress–strain response of fly ash with 10% lime and 1% gypsum, 7 days curing
(Ghosh and Subbarao 2007) .............................................................................................. 164
Figure 5.39 Peak-strength envelopes of untreated and treated MSW specimens .............. 165
Figure 5.40 Residual-strength envelopes of untreated and treated MSW specimens ........ 166
Figure 5.41 Variation of Young’s modulus at 50% failure stress for untreated and treated
MSW specimens under different effective confining pressures. ......................................... 167
Figure 5.42 Variation of Young’s modulus at failure stress for untreated and treated MSW
specimens under different effective confining pressures. ................................................... 167
Figure 5.43 Variation of stiffness ratio of untreated and treated MSW specimens under
different effective confining pressures. .............................................................................. 168
xviii
Figure 5.44 Effect of fly ash-quicklime contents on brittleness index of MSW specimens
under different effective confining pressures ..................................................................... 169
Figure 5.45 Primary compression of untreated and treated MSW specimens with different
fly ash-quicklime contents .................................................................................................. 170
Figure 5.46 Primary compression of treated MSW specimens with 20% fly ash-quicklime
for 28 and 93 days curing period ....................................................................................... 171
Figure 5.47 Effect of fly ash-quicklime contents on compression index of MSW specimens
............................................................................................................................................ 172
Figure 5.48 Void ratio versus pressure of raw and 6% stabilised Alloway Clay (Okoro et
al. 2011) ............................................................................................................................. 173
Figure 5.49 Void ratio versus pressure curves of raw and 10% CFA-stabilised Made Land
(Okoro et al. 2011) ............................................................................................................. 173
Figure 5.50 Variation of Cc for both expansive and nonexpansive clays (Phanikumar and
Sharma 2007) ..................................................................................................................... 174
Figure 5.51 e-log p curves of expansive clay specimens (Phanikumar and Sharma 2007)
............................................................................................................................................ 175
Figure 5.52 e-log p curves of nonexpansive clay (Phanikumar and Sharma 2007) .......... 175
Figure 5.53 Primary and secondary compression of MSW specimens when effective
confining pressure increased from 200 kPa to 300 kPa .................................................... 176
Figure 5.54 Volumetric strain of treated and untreated MSW specimens during primary
and secondary compression (when effective confining pressure increased from 200 kPa to
300 kPa) ............................................................................................................................. 177
Figure 5.55 Effect of fly ash-quicklime contents on the secondary compression index of
MSW specimens .................................................................................................................. 178
Figure 5.56 Effect of curing time on primary and secondary compression of MSW
specimens when effective confining pressure increased from 200 kPa to 300 kPa ........... 179
Figure 5.57 Effect of fly ash on secondary consolidation (Phanikumar and Sharma 2007)
............................................................................................................................................ 180
Figure 6.1 Cross-section of the numerical model .............................................................. 183
Figure 6.2 Dimensions of the model .................................................................................. 184
Figure 6.3 15-nodded triangle elements, used in the modeling ......................................... 184
Figure 6.4 Cross-section of generated mesh...................................................................... 184
Figure 6.5 Closer view of generated mesh......................................................................... 185
Figure 6.6 Consolidation and creep behaviour in standard Oedometer tests (Wehnert 2000)
............................................................................................................................................ 188
Figure 6.7 Logarithmic relationship between volumetric strain and mean stress including
creep (after Wehnert 2000) ................................................................................................ 189
xix
Figure 6.8 Yeild surface of the SS-model in p’-q plane (after Wehnert 2000) .................. 191
Figure 6.9 Vertical displacement for untreated landfill 10 years after applying traffic load
............................................................................................................................................ 196
Figure 6.10 Vertical displacement for 3-m treated with 26.6% fly ash-quicklime 10 years
after applying traffic load .................................................................................................. 196
Figure 6.11 Vertical displacement for 6-m treated with 26.6% fly ash-quicklime 10 years
after applying traffic load .................................................................................................. 197
Figure 6.12 Vertical displacement for 9-m treated with 26.6% fly ash-quicklime 10 years
after applying traffic load .................................................................................................. 197
Figure 6.13 Vertical displacement for 12-m treated with 26.6% fly ash-quicklime 10 years
after applying traffic load .................................................................................................. 198
Figure 6.14 Vertical displacement for 24-m treated with 26.6% fly ash-quicklime 10 years
after applying traffic load .................................................................................................. 198
Figure 6.15 Vertical settlement versus time for 3-m improved landfill with various fly ash-
quicklime contents under traffic load................................................................................. 200
Figure 6.16 Vertical settlement versus time for 6-m improved landfill with various fly ash-
quicklime contents under traffic load................................................................................. 200
Figure 6.17 Vertical settlement versus time for 9-m improved landfill with various fly ash-
quicklime contents under traffic load................................................................................. 201
Figure 6.18 Vertical settlement versus time for 12-m improved landfill with various fly ash-
quicklime contents under traffic load................................................................................. 201
Figure 6.19 Vertical settlement versus time for 24-m improved landfill with various fly ash-
quicklime contents under traffic load................................................................................. 202
Figure 6.20 Vertical settlement versus time for landfill treated with 6.7% fly ash-quicklime
content under traffic load ................................................................................................... 202
Figure 6.21 Vertical settlement versus time for landfill treated with 13.3% fly ash-
quicklime content under traffic load .................................................................................. 203
Figure 6.22 Vertical settlement versus time for landfill treated with 20% fly ash-quicklime
content under traffic load ................................................................................................... 203
Figure 6.23 Vertical settlement versus time for landfill treated with 26.7% fly ash-
quicklime content under traffic load .................................................................................. 204
Figure 6.24 Horizontal displacements for untreated landfill 10 years after applying load
............................................................................................................................................ 205
Figure 6.25 Horizontal displacement for 6-m treated with 26.6% fly ash-quicklime 10 years
after applying load ............................................................................................................. 205
Figure 6.26 Horizontal displacement for 12-m treated with 26.6% fly ash-quicklime 10
years after applying load ................................................................................................... 206
xx
Figure 6.27 Horizontal displacement for 24-m treated with 26.6% fly ash-quicklime 10
years after applying load ................................................................................................... 206
Figure 6.28 Horizontal displacement versus depth for the landfill treated with 26.7% fly
ash-quicklime content in various improvement depths at a section below the toe of
embankment........................................................................................................................ 207
Figure 6.29 Horizontal displacement versus depth for 9-m improved landfill treated with
various fly ash-quicklime contents at a section below the toe of embankment .................. 207
xxi
List of Tables
Table 1.1 Per capita waste generation for Australia, estimated, 2006/07 (Hyder Consulting
2008) ...................................................................................................................................... 2
Table 2.1 Engineering properties of MSW required for design (from Dixon and Jones,
2005). ................................................................................................................................... 12
Table 2.2 Hyperbolic parameters for different compaction effort and amount of soil cover
(Zekkos 2005) ....................................................................................................................... 16
Table 2.3 Statistical summaries of bulk unit weight data for fresh waste (after Fassett et al.
1994) .................................................................................................................................... 17
Table 2.4 Bulk unit weights from international literature (after Dixon and Jones 2005) ... 17
Table 2.5 Recommended Values of Cα Parameters (Sharma 2007) .................................... 29
Table 2.6 Model parameters for settlement prediction (Hossain and Gabr 2005) .............. 31
Table 2.7 MSW shear strength laboratory tests results from literature .............................. 36
Table 2.8 Summary of tests relating the vertical hydraulic conductivity of MSW type
materials to dry density. ....................................................................................................... 55
Table 2.9 kh :kv ratios obtained by laboratory testing of wastes. ....................................... 57
Table 3.1 Two basic landfill classes .................................................................................... 80
Table 3.2 Approximate number of landfill by class and State ............................................. 80
Table 3.3 Depth and characterisation of test pits ................................................................ 85
Table 3.4 Moisture content and organic content results of collected waste samples .......... 95
Table 4.1 Chemical composition of Eraring fly ash .......................................................... 104
Table 4.2 Compaction test results for untreated and treated MSW samples (particles
smaller than 9 mm) ............................................................................................................ 108
Table 4.3 Compaction test results for untreated and treated MSW samples (particles
smaller than 19 mm.) ......................................................................................................... 109
Table 4.4 Ranges of additives to MSW samples ................................................................. 113
Table 4.5 Summary of unconfined compressive strength specimens and tests performed. 115
Table 4.6 Summary of large direct shear specimens and tests performed at the University of
Technology Sydney ............................................................................................................. 121
Table 4.7 Summary of permeability test specimens through consolidation of specimen in
triaxial cell ......................................................................................................................... 122
Table 4.8 summary of permeability test specimens and their details ................................ 124
Table 4.9 Summary of triaxial specimens and tests performed ......................................... 128
Table 4.10 Summary of consolidation specimens and tests performed ............................. 129
xxii
Table 5.1 Results of unconfined compressive strength tests .............................................. 135
Table 5.2 Summary of large direct shear tests results ....................................................... 141
Table 5.3 Effect of fly ash on compaction Behavior, and hydraulic conductivity (Modified
after Kumar and Sharma 2004) ......................................................................................... 150
Table 5.4 Summary of permeability test results through consolidation of specimen in
triaxial cell ......................................................................................................................... 153
Table 5.5 Results of Peak and residual principal stress difference in triaxial compression
test ...................................................................................................................................... 159
Table 5.6 Peak and residual strength and elastic parameters for untreated and treated
MSW in triaxial compression test. ..................................................................................... 165
Table 5.7 Variation of primary compression index for MSW specimen treated with different
fly ash-quicklime content and curing time ......................................................................... 171
Table 5.8 A summary of swell and compression indices of raw and stabilised soils. (after
Okoro et al. 2011) .............................................................................................................. 173
Table 5.9 Variation of secondary compression index for MSW specimen treated with
different fly ash-quicklime content and curing time ........................................................... 178
Table 5.10 Effect of Fly Ash on Coefficient of Secondary Consolidation, Cα (Phanikumar
and Sharma 2007) .............................................................................................................. 180
Table 6.1 Parameters for soft soil creep model in FEM analysis ...................................... 193
Table 6.2 Model parameters used for cover layer and road embankment ........................ 194
Table 6.3 Fly ash-quicklime contents and depths of treatment for closed landfill model . 194
Table 6.4 Vertical displacement of the landfill model 10 years after applying traffic load
............................................................................................................................................ 195
Table 6.5 Vertical displacements of the landfill model, 20 years after applying the traffic
load..................................................................................................................................... 199
xxiii
CHAPTER 1
1 Introduction
1.1 Introduction
Every year in Australia, millions of tonnes of municipal solid waste (MSW) are produced,
the variation of which depends on the population size of the different states. This is
reported by presenting data on waste generation on a per capita basis. As shown in Table
1.1, ACT, South Australia and Western Australia have the highest amounts of waste
generation per capita, and Queensland has the lowest levels of total waste generation
(Hyder Consulting 2008).
Municipal solid waste is a combination of wastes that are mainly collected from residential
and commercial sources. MSW consists of food, paper products, plastics, rubber, textiles,
wood, ash, and soils. Construction on closed landfill sites requires particular planning and
design to consider the characteristics of the waste and safety issues. Several engineering
solutions have been introduced by researchers for stabilising old landfills, so the challenge
of this research is to develop an effective chemical stabilisation technique in conjunction
with mechanical stabilisation approaches.
1
Additives such as lime, cement, fly ash, and lime-fly ash admixtures are used to improve
the engineering properties of soils. The choice and effectiveness of an additive depends on
the type of soil and its field conditions, although knowledge of the mechanistic behaviour
of treated soil is as equally important as selecting the stabiliser.
Table 1.1 Per capita waste generation for Australia, estimated, 2006/07 (Hyder Consulting
2008)
Waste generated
State / territory Population Kilograms per
capita
New South West 6,888,000 1,904
Victoria 5,205,000 1,976
Queensland 4,181,000 1,834
Western Australia 2,106,000 2,492
South Australia 1,584,000 2,259
Australian Capital Territory 340,000 2,249
Tasmania 493,000 Unknown
Northern Territory 215,000 Unknown
Many closed municipal solid waste (MSW) landfill sites are near urban areas, even though
originally established away from residences or business communities (Figure 1.1). There
are high expectations of improving the geotechnical properties of these sites for
redevelopment in spite of potential risks for human health and the environment.
Construction on closed landfill sites is generally a challenging task because developers are
not interested in constructing on land containing waste materials due to their complex
behaviour, including significant creep deformation, differential settlement, and the high
amount of moisture content and weak shear strength of waste. The stress-strain response of
MSW must be studied further to have sustainability between the shear strength and
deformation along failure surfaces. Not many comprehensive investigations have been
conducted on the behaviour of waste materials, and research on the stabilisation of waste
using ground improvement techniques is limited. While it is not possible to fully
characterise the engineering properties of wastes due to their heterogeneous nature, it is
important to understand their basic behaviour and the likely ranges of their key engineering
properties.
2
Figure 1.1 Payatas landfill in Philippin (photo by: Scott Merry)
The stabilisation of landfills and waste disposal sites for structural and environmental
purposes has been performed through the application of current soil stabilisation and
ground improvement techniques. The deep dynamic compaction (DDC) technique is a
common ground improvement technique due to its relatively economical and easy
application. The deep dynamic compaction technique has been carried out with success in a
large range of soils including MSW. Zekkos et al. (2012) reviewed 64 case histories
regarding the effectiveness of the DDC on MSW landfill sites. Their results indicated that
the depth of improvement was smaller in MSW compared to cohesionless soils. It can be
noted that the settlement caused by DDC depends on the applied energy and it is in the
range of 5% to 25% of the MSW thickness.
With deep dynamic compaction large voids are reduced and afterward other techniques
such as fly ash-lime grouting can further reduce the remaining smaller voids. Blacklock
3
(1987) showed that a lime slurry pressure injection process can be utilised to control
methane gas, and a lime/fly ash slurry injection can help protect the groundwater, neutralise
leachate, and place curtain walls to prevent leachate migration. Not only that, replacing
cement with by-product materials such as fly ash can decrease the expense of stabilisation.
Many researchers reported the application of fly ash in geotechnical projects (e.g. Kawasaki
et al. 1981; Kitazume et al. 2000; Kehew 1995). According to Horpibulsuk et al. (2011),
from an economic and environmental viewpoint, rich materials in slaked lime (Ca(OH)2,
calcium hydroxide) can be treated together with pozzolanic materials, such as fly ash, to
develop a cementitious material. The experimental program conducted in this study was
inspired by previous investigations into the properties of MSW outlined in Chapter 2. A
thorough laboratory test on aged MSW using chemical agents has been carried out to
quantify the behaviour of stabilised MSW in this research.
The specific objectives of the experimental component of this study are as follows:
x Investigating several landfills in New South Wales to get permission to take samples
from the field for conducting laboratory tests.
x Modifying of existing equipment rigs at University Technology Sydney for testing waste
materials with specific maximum particles size.
4
x Developing appropriate sample preparation methods and testing procedure for municipal
solid waste with different particle sizes to use in extensive experimental program.
x Studying the effect of fly ash-quicklime contents on revised and normal maximum dry
density and optimum moisture content of municipal solid waste.
x Finding out the shear strength parameters of municipal solid wastes before and after
chemical stabilisation with fly ash-quicklime and also investigating the effect of curing
time on those parameters using the large shear box and triaxial apparatus.
x Studying the compressibility and the permeability of municipal solid waste before and
after treatment with different fly ash and quicklime contents using triaxial and
permeability apparatus and also considering effect of curing time and confining pressure
on those parameters.
x Investigating the stiffness and the brittleness characteristic of untreated and fly ash-
quicklime treated municipal solid waste with different contents and also determining the
effect of confining pressure on those parameters.
x Determining the optimum percentage of chemical agents (fly ash and quicklime) for use
in a high performance stabilisation of municipal solid waste.
x Developing a numerical model using PLAXIS to predict the vertical and horizontal
displacements of closed landfill sites.
x Investigating the effects of various fly ash-quicklime contents on displacement of closed
landfill sites under traffic load, long time after closure.
5
x Investigating the effects of depth of improvement, drainage layer, the top cap cover, and
groundwater position on vertical and horizontal displacements of closed landfill sites
while incorporating in the numerical model.
x Investigating the effect of time on displacement of treated and untreated closed landfill
site after applying traffic load for different depths of improvement.
This thesis reports the results and findings of the field work, the waste characterisation, and
the laboratory tests performed at the University of Technology Sydney. MSW properties
investigated through the course of this research are presented in this thesis. This thesis
consists of 7 chapters which are organised as follows:
Chapter 3 describes the excavations and procedures for collecting samples, and
recommended procedures for the characterisation and classification of collected MSW. It
contains a description of the sites visited for sampling purposes, and investigations into the
in-situ MSW characterisations in a former landfill at Bankstown, including its composition,
age, level of moisture content and rate of waste degradation. These data facilitated in
selecting an appropriate sampling procedure that was used in the experimental program of
this study. Material loss tests were carried out to estimate the moisture content and organic
content of the waste material.
6
Chapter 4 describes the materials, sample preparation, laboratory tests and procedures, and
experimental program used in the current research. It also describes the testing apparatus
used for this research at the University of Technology Sydney.
Chapter 5 presents the results of the geotechnical laboratory program to estimate how the
use of fly ash and quicklime could improve the geotechnical properties of MSW. An array
of experimental tests has been conducted on treated and untreated MSW samples, including
sieve analysis, Atterberg limits, compaction, unconfined compressive strength, direct shear,
permeability and consolidated-drained triaxial tests. The samples were prepared by
integrating MSW, with a mixture of fly ash-quicklime at a ratio of 3:1 in percentages of 5,
10, 15 and 20 of fly ash by dry weight of the MSW.
Chapter 6 briefs numerical analysis to predict the vertical and horizontal displacement of
closed landfills. The finite element program PLAXIS version 9 has been used to evaluate
the settlement of stabilised landfill. The soft soil creep model was used for this analysis.
Untreated and treated MSW parameters used in FEM analysis have been estimated from
extensive laboratory tests in this study explained in Chapter 5.
Chapter 7 presents a summary and conclusions drawn from this study. Several
recommendations for further research are also outlined in this chapter.
7
CHAPTER 2
2 Literature Review
8
clear instructions for choosing appropriate values of the engineering parameters to be used
in design and assessing the behaviour of waste.
Dixon and Jones (2005) reported that the constituents of waste have a controlling effect on
the average unit weight of waste mass. The individual components of waste have a wide
range of particle unit weights and these can change over time. Components may have voids
within and between components, which means that a significant percentage of waste
particles behave differently to soil particles due to their high compressibility. Degradation
of components with organic content will result in a loss of mass, changes in size, and
alteration of the mechanical properties (i.e. compressibility and shear strength). It will also
change the density of the component. Dixon and Jones (2005) also noted that as a waste
body degrades, the void ratio reduces and hence a volume reduction occurs. Although there
are few field measurements in degraded waste, it generally shows that degradation results
in an increase in its unit weight.
Municipal solid wastes (MSW) are generally composed of the following main components:
The liner system acts as a barrier to the transport of leachate solutes. The liner system is
9
located on the lateral and bottom surface of a landfill. Its major purpose is to isolate solid
waste and prevent contamination of the surrounding soil and groundwater. The barrier can
consist of a compacted clay layer, a geomembrane, a geosynthetic clay layer, or a
combination of both.
Leachate is the portable part of the solid waste in landfill. Leachate is generated from liquid
compressed out of the waste itself; water that permeates into the landfill is called primary
leachate and that which percolates through the waste is called secondary leachate. Leachate
consists of solvent and solutes. A leachate collection system is used to collect the leachate
generated in a landfill, to prevent the gradual increase of a leachate head on the liner, and
then to drain leachate to a wastewater treatment plant by a sanitary sewer line or a leachate
storage tank for treatment and disposal.
Municipal solid waste can produce large volume of gas during decomposition, the primary
components of which are methane (CH4) and carbon dioxide (CO2). A gas collection and
control system is used to gather gas from the landfill as the organic constituents of the solid
waste decompose. Landfill gas can be used to generate energy of flared under controlled
conditions.
Drainage and barrier layers constitute the final covering system whose major purpose is to
reduce water permeation into the landfill to decrease the amount of leachate produced after
closure. A layer of soil protects the underlying layers against intrusion, damage, and the
effects of frost. Unlike a liner, a cover is only used as a hydraulic barrier against permeation
from outside water. It is not used as a barrier against the leakage of leachate solutes under
combined advection and diffusion. It should be noted that a covering system should be
designed with the same degree of consideration and care that is applied to the liner.
10
2.3.5 Composite Liners
It is important that the liner system should work properly to maintain the efficiency of the
landfill. The first liners consisted of a single layer of clay or a synthetic polymeric
membrane, but by the time the tendency towards using composite liner systems had
increased. Composite liners consist of both clay and synthetic geomembranes together with
the drainage layers.
2.4.1 Introduction
Fassett et al. (1994) summarised the common difficulties in evaluating the MSW unit
weight as follows:
x Separation of the contribution of daily soil cover
x Time and depth effects on assessing the changes in unit weight
x The majority of reported values are related to waste at or near the surface
x Obtaining data on the moisture content of the waste.
Fassett et al. (1994) suggested that the following factors should be recorded along with
measured unit weights:
x MSW compositions, including the daily cover and moisture content
x Methods and degree of compaction
x The depth at which the unit weight is measured
x The age of the waste
11
Values can be given as the dry unit weight (sample could have been artificially dried), the
bulk unit weight (some moistures are present but the waste is not saturated) and the
saturated unit weight. In most studies, it is the bulk unit weight that is measured and
reported.
Referring to Table 2.1, Dixon and Jones (2005) observed that the MSW unit weight is the
only material characteristic which has a significant effect on almost all components of
landfill design and analysis.
Table 2.1 Engineering properties of MSW required for design (from Dixon and Jones,
2005).
Horizontal
Unit Vertical Shear Hydraulic
Design Case In-Situ
Weight Compressibility Strength Conductivity
Stress
Subgrade stability ● ● ●
Subgrade integrity ● ● ●
Waste slope stability ● ● ● ●
Shallow slope liner stability ● ● ● ●
Shallow slope liner integrity ● ● ● ●
Steep slope liner stability ● ● ● ●
Steep slope liner integrity ● ● ● ●
Cover system integrity ● ● ●
Drainage system integrity ● ●
Leachate/gas well integrity ● ● ● ● ●
Choosing a typical profile of MSW unit weight has a significant effect on the determination
of many engineering properties. Several researchers reported different numbers and
methods for estimating MSW unit weight, as follows:
x Matasovic and Kavazanjian (1998) noted that a 60 m high landfill with a constant
unit weight profile of 10 kN/m3 has an evaluated waste capacity of 40% less than a
12
landfill with the same unit weight profile as was measured in-situ at the OII landfill
unit, and it corresponds to a constant unit weight of 15 kN/m3.
x Wiemer (1982) suggested a unit weight of 6 kN/m3 near the surface for a
compaction lift thickness of 2 m and a highest value of 10 kN/m3 at a depth of 20 m,
whereas waste is placed in layers that range from 0.2 - 0.5 m thick.
x Henke (1985) evaluated a unit weight of 15.8 kN/m3 for a combination of MSW and
industrial waste.
x Landva and Clark (1990) distinguished between the unit weight of the soil cover,
the unit weight of refuse, and the total unit weight, and noted a reduction in the unit
weight of the soil cover with an increase in water content. The dry unit weight of
the soil cover varies from 22 kN/m3 for low water contents, to 10 kN/m3 for higher
water contents.
x Richardson and Reynolds (1991) in their investigations of the stability of a landfill
in central Maine, noted the significance of the unit weight of the MSW. They
showed that the landfill was 12 m high, and based on the weight and volume data,
the unit weight was12.5 kN/m3. In addition they reported average unit weight of
15.34 kN/m3 based on the results of large-scale density tests which were executed
by excavating test pits of 8 m3 of waste.
x Sanchez-Alciturri et al. (1993) reported a unit weight of 12 kN/m3 for Meruello
landfill in Spain.
x Fasset et al. (1994) illustrated that the unit weight ranges between 3-9 kN/m3 for
poor compaction, 5-8 kN/m3 for moderate compaction, and 9-10.5 kN/m3 for fills
with good compaction.
x Kavazanjian et al. (1995) suggested a unit weight profile with a minimum value of 6
kN/m3 at the surface and 13 kN/m3 at depths greater than 50 m.
x Withiam et al. (1995) found out the average unit weight to be between 11 and 13
kN/m3on the Dekorte Park landfill, New Jersey, USA. The evaluation was based on
two test pits, application of a down hole nuclear depth density gauge and the
investigation of historical records of weight of waste and volume of landfill.
x Manassero et al. (1996) presented the unit weights from Dendermonde landfill and
pointed out that freshly deposited paper waste had a value of 5 kN/m3, compacted
13
milled paper waste a value of 8 kN/m3 and for compressed blocks of milled paper
waste a value of 8 kN/m3.
x Shimizu (1997) proposed a total unit weight of 7-10 kN/m3 for Port Harbor landfill
in Japan.
x Oweis and Khera (1998) indicated in-situ unit weight estimation by applying a 300
mm diameter bucket auger. Based on these results the unit weights increased with
depth and varied from 9.5 to 17.5 kN/m3. In addition they reported that at the same
depth the total unit weight of fresh landfill was slightly more than old landfills due
to the higher moisture content. However, dry unit weights did not change much.
x Gomes et al. (2002) calculated the unit weight of fresh waste trench (0.6 m high, 1.2
m wide and 2m long) and also a borehole (0.2 m diameter, 11 m height) in an area
with 2 to 3 year old waste. The unit weights measured 11.3kN/m3 and 11.6 kN/m3
respectively.
x Geosyntec (2003) measured the in-situ unit weight of MSW. The volume of the
borehole was evaluated using two methods. Firstly, assuming that the diameter of
the borehole was the same as the diameter of the bucket, and secondly by assuming
a 16% overbore on the borehole diameter. The Author assumed the first method as
an upper boundary value of the waste unit weight while the second method is a
lower boundary value. At depths of 0-3 m, 3-6 m, and greater than 6 m, the average
unit weight was designated as 8 kN/m3, 10 kN/m3 and 11.3 kN/m3 respectively.
Zekkos (2005) reported using 3 methods to evaluate the MSW unit weight:
1. Surveys and records: This method allows the weight of waste and soil buried in landfill
to be measured and based on topographic surveys, the volume of waste and soil can be
measured. By this information the average in situ unit weight can be measured.
2. Undisturbed samples unit weight: The unit weight can be estimated if representative
undisturbed samples can be taken. However this method is not suggested because large
14
particles exist in landfills and are probably not taken sufficiently in samples, so it can cause
significant errors.
3. In-situ large-scale methods: This method is simulated from the sand cone density test but
is conducted on a larger scale. Large test pits are excavated and the weight of excavated
waste material is measured. The volume of the test pit is evaluated either by surveying or
by replacing the waste material with calibrated material of known unit weight, such as
uniform gravel. Zekkos (2005) mentioned that the in-situ large scale method is the most
reliable approach for estimating the in-situ unit weight of MSW because it includes large
volumes of material with large particles.
Zekkos (2005) presented a hyperbolic model for the MSW unit weight according to the
experimental test results. Equation (2.1) illustrates the model recommended to calculate the
unit weight based on the average confining stress. He also proposed Equation (2.2) which is
a more useful equation for MSW unit weight based on depth. This type of relationship is
more practical because field data usually includes unit weights at specific depths.
Vv
J Ji (2.1)
D v E v .V v
Z
J Ji (2.2)
D E .Z
where J i is the unit weight at the surface, which can be estimated by executing test pits, and
D v and E v are modelling parameters in the units of m4/kN and m3/kN, respectively. V v and z
are the vertical overburden stress and depth (m) at the point where the MSW unit weight is
to be evaluated respectively. The J i in Equation (2.1) shows the influence of primary
compaction, while the second term represents the influence of confining stress on the MSW
unit weight.
The physical definition of the α and β parameters, and the hyperbolic parameters for
different compaction efforts are presented in Figure 2.1 and Table 2.2. Zekkos (2005) stated
15
that parameter β is a function of the difference in the unit weight between the surface and at
a lower depth where the unit weight is constant. The parameter α is a function of the growth
in unit weight near the surface.
Figure 2.1 Physical meaning of the hyperbolic parameters α and β (Zekkos 2005)
Table 2.2 Hyperbolic parameters for different compaction effort and amount of soil cover
(after Zekkos 2005)
Compaction effort Ji β α
MSW consists of different materials where a large part of the constituents have a high void
ratio and a high compressibility. By using the advantages of compaction, the voids within a
single component and the voids between different components can be reduced. Dixon and
Jones (2005) cited that the unit weight of compacted waste would depend on the following:
x the waste components,
x the thickness of layer,
x the weight and type of compaction plant, and
x the number of times the equipment passes over the waste.
16
A thickness of 0.5 to 1.0 m for each layer will promote good compaction and therefore high
unit weights. In addition, it is typical for waste to be placed in 2 to 3 m thick layers. These
results are associated with poor to moderate compaction. Fassett et al. (1994) conducted a
comprehensive survey of the information available from literature on bulk unit weight. The
degree of compaction was estimated from the practices at each site. Table 2.3 shows a
statistical summary of bulk unit weight data for fresh waste, with the outcomes showing a
large reduction in the unit weight when compaction was either small or non-existent. Dixon
and Jones (2005) also prepared a summary of bulk unit weights from literature, which is
presented in Table 2.4.
Table 2.3 Statistical summaries of bulk unit weight data for fresh waste (after Fassett et al.
1994)
Poor Moderate Good
Compaction Compaction Compaction
Range (kN/m3) 3.0 to 9.0 5.0 to 7.8 8.8 to 10.5
Average (kN/m3) 5.3 7.0 9.6
Standard Deviation (kN/m3) 2.5 0.5 0.8
Coefficient of Variation (%) 48 8 8
Table 2.4 Bulk unit weights from international literature (after Dixon and Jones 2005)
Measured Bulk
Country Unit Weights Comments References
(kN/m3)
6 x Compacted in 2 m lifts using steel
United wheeled 21 tonne compactor Watts and
Kingdom 8 x 0.6 m lifts using same compactor as Charles (1990)
above
5 to 10 x Common compaction practice Manassero et
Belgium
al. (1996)
7 x Upper layers of fresh (non-degraded) Gourc et al.
France
MSW (2001)
6 to 7 x Fresh MSW after initial placement
Kavazanjian
USA 14 to 20 x Degraded waste with high % of soil like (2001)
material
17
2.4.6 Effect of Depth on the Unit Weight of MSW
The unit weight of waste changes with effective stress is a function of depth. Figure 2.2,
which is present by Powrie and Beaven (1999), shows the changes in dry density, saturated
density, and the density at field capacity, with the vertical effective stress. The data were
collected by compressing samples of waste in a large diameter cylindrical test chamber.
Powrie et al. (1998) noted that compaction can have a similar influence on the burial of the
waste by several metres of overburden.
Figure 2.2 Relationships between density and average vertical stress. Trend lines shown
are based on average measured values (Powrie and Beaven 1999, taken from Dixon and
Jones 2004)
Dixon and Jones (2005) stated that the moisture content of waste depends on:
2.5 Compressibility
2.5.1 Introduction
Bjarngard and Edgers (1990) mentioned that the total settlement in municipal solid waste
(MSW) landfills can be evaluated as between 25% and 50% of the thickness of the waste.
Settlement occurs partially during the filling phase, but a huge amount of long term
settlement continues. Landfill settlement is the result of load and bio-degradation. Current
settlement models can consider the effects of load but the numbers of models that are able
to evaluate bio-degradation are limited. An accurate prediction of the compressibility of
landfill is a challenging task because a large number of variables are involved in the
settlement process. The following factors have significant effects on analysing waste
settlement:
The process of waste settlement can be divided into the following major procedures
(Sowers 1973; Edil et al. 1990; Sharma and Reddy 2004):
19
x Physical and mechanical processes: These consist of the reorientation of grains, the
movement of fine materials into larger voids, and a breakdown of the void gap.
x Physico-chemical process: This includes corrosion, combustion, and oxidation.
x Dissolution process: This includes dissolving soluble substances by permeating
liquids and then forming leachate.
x Biological decomposition: The organic materials in the waste will decompose
during time, affected by temperature, humidity, and organic contents in the refuse.
Yen and Scanlon (1975) pointed out five mechanisms for long term settlement, including:
x Movement of fines into large voids
x Loss of strength due to chemical and biological reactions
x Loss of material due to bio-degradation and production of methane
x Creep processes
x Consolidation processes
The mechanisms of settlement throughout the life of a landfill site are presented by
McDougall (2011) in Figure 2.3. The settlement due to load, which may be the result of self
weight or surcharge will be completed within a month of application. In addition,
mechanical creep is happening at a decreasing rate throughout the life of the site.
20
The bottom half of Figure 2.3 shows the classification of settlement as suggested by several
researchers. Sowers (1973), Morris and Woods (1990), and Wall and Zeiss (1995) noted the
effects of load as initial and primary settlement happening over the first month. After one
month the effects of mechanical creep, physico-chemical/corrosion and bio-degradation
(Secondary or Long-term phase) continue for 30 or more years. Bjarngard and Edgers
(1990) and Park and Lee (1997) use various time frame but their interpretations are
consistent with others.
It is obvious that only creep and consolidation would be identified as current settlement
mechanisms. Wall and Zeiss, (1995) noted that the waste is less than fully saturated,
especially at the time of filling, and particularly even if saturated, has a high permeability to
permit any excess pore water pressure.
Manasseroet et al. (1996) have listed the main points of the mechanisms resulting in the
compression of waste. Based on Equation (2.3), it can be assumed that total settlement
consists of two main constituents; primary compression (δp) and secondary compression
(δs)
G G p Gs (2.3)
Primary settlement is due to physical and mechanical mechanisms. Secondary or long term
settlement, primarily due to physico-chemical and bio-chemical decay, occurs under
constant load after primary settlement is complete. The primary settlement of municipal
21
solid waste occurs within the first four months after the load has been placed (Sowers 1973;
Bjarngard and Edgers 1990; Sharma 2000; NAVFAC 1983).
Dixon and Jones (2005) stated that the secondary compression of waste contains all the
effects of creep and degradation (chemical and biological). Creep effects include time
dependent particle distortion and reorientation. Degradation contains the collapse of waste
due to corrosion and degradation of organic compounds. They noted that Bio-degradation is
the most significant constituent of secondary compression in MSW landfills. Many
methods have been suggested to characterise and estimate secondary compression, whereas
in reality the degradation process is affected by a range of factors which change within a
landfill and with time, such as the moisture content, temperature, and level of stress.
According to the extensive field data shown in Figure 2.4, Bjarngard and Edgers (1990)
identified two different phases of secondary settlement. In the first phase, settlement
happens at a relatively low rate and is controlled by mechanical processes, while the second
phase occurs at a higher rate of settlement that can be related to decomposition.
Figure 2.4 Landfill settlement vs. log time from field case histories (Bjarngard and Edgers
1990)
22
2.5.6 Total Compression
Dixon and Jones (2005) stated that to estimate the total settlement of waste, the following
considerations are necessary:
x Settlement of waste caused by self-weight (primary compression)
x Settlement caused by the weight of each subsequent layer (primary compression)
x Settlement caused by secondary compression, taking into account that Cα is likely to
decrease with age
x Settlement of the mineral basal liner caused by primary and secondary compression
and settlement of any compressible sub-grade
Due to the heterogeneous nature of landfill components and varied rates of decomposition,
differential settlement often happens, but this issue is more complicated because filling
often takes place on top of older waste deposits at different times. Differential settlement is
important because it can endanger the stability of the final cap and the integrity of the
geosynthetic layer and compacted clay barriers.
Figure 2.5, modified after McDougall (2011), shows a range of factors influencing the
primary and secondary settlement mechanisms of landfill. McDougall outlined the
influencing factors and explained the condition of the waste within a landfill cell, and the
initial conditions which affect the internal conditions. However, some factors affecting bio-
degradation such as moisture, leachate, and the composition of gas are clearly not included
in the settlement models. The effect of these factors can vary by time, although Landva et
al. (1984) found that the effects of decomposition are too complex to be predicted.
23
Figure 2.5 Mechanisms and factors influencing landfill settlement. (Modified after
McDougall 2011)
The total settlement (ΔH) is divided into primary settlement (ΔHp) and secondary
settlement (ΔHs). The following equations present the Sowers (1973) method of estimating
settlement.
P0 'P
'H p HC ce log (2.4)
P0
t2
'H S H 1CD log (2.5)
t1
where:
24
Δp = incremental pressure;
t1 = starting time for secondary compression;
t2 = ending time for secondary compression
Oweis and Khera (1998) reported values of Cα for different ages of waste materials. Table
2.5 presents the values chosen from their summary and presents the difficulties of trying to
use one Cα value for the entire period of secondary compression. Since the rate of
degradation is not constant with time, it is not unexpected that Cα is not a constant number.
McDougall and Pyrah (2001) mentioned that methods of estimating settlement based on
modelling the bio-degradation procedure are growing.
Sharma (2000) noted that estimating settlement requires calculation of ΔHs, which is made
after about 15 to 20 years of waste placement. Where t2 is approximately 20 years and the
value of ΔHs is not very sensitive to the choice of t1 between three or 4 months. Table 2.6
shows summary of MSW Cce and Cα Parameters collected by Sharma (2007) from several
previously published data.
Table 2.5 Secondary compression parameters for MSW material (After Oweis and Khera
1998)
Material Cα
Ten year old landfill 0.02
Fifteen year old landfill 0.24
Fifteen to twenty year old landfill 0.02
Old landfill 0.04
Old landfill with high soil content 0.001 to 0.005
25
Table 2.6 Summary of MSW Cce and Cα Parameters from Literature (Collected by Sharma
2007)
Reference Primary Cce Secondary Cα
a
Sowers (1973) (for e0=3) 0.1–0.41 0.02–0.07
a
Zoino (1974) 0.15–0.33 0.013–0.03
a
Converse (1975) 0.25–0.3 0.07
a
Rao et al. (1977) 0.16–0.235 0.012–0.046
a
Oweis and Khera (1986) 0.08–0.217 -
a
Landva et al. (1984) 0.2–0.5 0.0005–0.029
a
Bjarngard and Edgers (1990) - 0.004–0.04
a
Wall and Zeiss (1995) 0.21–0.25 0.033–0.056
a
Gabr and Valero (1995) 0.2–0.23 0.015–0.023
a
Boutwell and Fiore (1995) 0.09–0.19 0.006–0.012
a
Stulgis et al. (1995) 0.16 0.02
a
Green and Jamenjad (1997) - 0.01–0.08
a
Landva et al. (2000) 0.17–0.24 0.01–0.016
Zaminskie et al. (1994) 0.01–0.04 0.001–0.006
b
El-Fadel and Al-Rashed (1998) - 0.1–0.32 with leachate recirculation
b
Earth Tech (2001) - 0.18–0.26 with leachate recirculation
0.014–0.063 for “fresh waste”
Park et al. (2002) - 0.087–0.34 for waste undergoing active
decomposition
Marques et al. (2003) (field monitoring) 0.073–0.132 -
0.015–0.03 for creep
Hossain et al. (2003) (laboratory testing 0.16–0.37 0.19 for waste undergoing active
decomposition
Anderson et al. (2004) (field monitoring) 0.17–0.23 0.024–0.030
Cα(EL)c : 0.02
Calculated by Sharma (2000) -
Cα(SW)d: 0.19–0.28
Cα(SW): 0.04 for “fresh waste”
Yolo County case history -
Cα(SW): 0.16 for waste in bioreactor
Lewis et al. (2004) (field monitoring) for Cα(EL)0.014 for waste treated with DDCe
-
external load Cα(EL): 0.045 for waste treated with surcharge
a
Data originally presented by Landva et al. (2000).
b
Data originally presented by Hossain et al. (2003).
c
EL=external load.
d
SW=self-weight.
e
DDC=deep dynamic compaction.
26
2.5.8.2 Rheological model
Based on Gibson and Lo’s (1961) rheological model, Edil et al. (1990) suggested a method
for estimating total settlement (ΔH) based on the following equation:
Where
H = initial height of waste;
Δp = change in pressure;
α = primary compression parameter;
b = secondary compression parameter;
λ/b = rate of secondary compression,
t = time since load application.
where
ΔH = total settlement;
H = initial height of waste;
Δp = change in pressure;
m = reference compressibility;
n = compression rate;
tr= reference time introduced into equation to make time dimensionless;
t = time since load application.
Ling et al. (1998) used the following hyperbolic function to estimate long term settlement.
27
t
S
1 / p0 t / S ult (2.8)
where
The values of p0 and Sult can be found through a regression analysis conducted on the t/Sult
versus t relationship. Ling et al. (1998) found that the hyperbolic function method provided
a valid estimation of long term settlement, in contrast to the logarithmic and power function
methods.
Based on the type of loading applied on MSW landfills, Sharma (2007) showed that
secondary settlement can be classified into the following two divisions:
This type of settlement is induced by the load imposed as a result of the weight of waste
material on the underlying layers of waste. Long term settlement due to self weight can be
estimated with the following equation:
t2
'H S 'H ( SW ) CD ( SW ) H 1 log (2.9)
t1
where
Time dependent secondary settlement happens over a long period of time in response to
externally applied loads. This type of settlement can be estimated by using the following
equation
t2
'H S 'H ( EL) CD ( EL) H 1 log (2.10)
t1
where
Table 2.7 summarises the recommended Cα values for various conditions recommended by
Sharma (2007). External loads are applied during or after the landfill has been closed, and
therefore data for Cα(EL) are available to waste undergoing active decomposition.
29
Bjarngardand and Edgers (1990) collected MSW landfill data and suggested the following
model for estimating total settlement:
'H P0 'P t t
CR log CD 1 log 2 CD 2 log 3 (2.11)
H P0 t1 t2
where:
ΔH = settlement
H = initial thickness of waste layer;
ΔH/H = vertical strain (normalized settlement);
P0 = Initial average vertical effective stress;
ΔP = average induced vertical stress increment;
t1 = time (days) for completion of “initial” compression;
t2 = time (days) for completion of “intermediate” compression;
t3 = period of time (days) for prediction of settlement;
CR = compression ratio;
Cα1 = intermediate secondary compression index;
Cα2 = long term secondary compression index.
Hossain and Gabr (2005) showed that the settlement model (Figure 2.6) that considers both
creep and biological components can be presented as:
§t · §t · §t ·
H (t ) CDi log ¨¨ 2 ¸¸ C E log ¨¨ 3 ¸¸ CDf log ¨¨ 4 ¸¸ (2.12)
© t1 ¹ © t2 ¹ © t3 ¹
30
By using this model, they compared the predicted MSW settlement with the monitored
field settlement from six different bio-reactor landfills. The settlement data were gathered
from the Mountain View Control Landfill Project, CA, Yolo county landfill, CA and
Spruce Ridge Landfill, MN, Richmond Landfill, VA, Atlantic Landfill, VA, AND Mid-
Peninsula Landfill, VA (Earth Tech Consultants, 2001).The model parameters used to
estimate the settlement are given in Table 2.8.
Table 2.8 Model parameters for settlement prediction (Hossain and Gabr 2005)
1
Landfills Cαi Cβ Cαf t1 (days) t2 (days) t3 (days) t4 (days)
Figure 2.6 Time dependent secondary settlement model extending to three stages as
proposed by Hossain and Gabr (2005)
Hossain and Gabr (2005) showed a time dependent secondary settlement model extending
to three stages, as presented in Figure 2.6. They assumed the time factor t2 as a value of
100, 200, and 500 days for the purpose of estimation. The evaluated settlement data shown
in Figure 2.7 reveals that in most cases the time factor t2 is between 100 and 500 days. They
reported that the laboratory creep index (0.03) and biological degradation index (0.19)
31
fitted well with all the field rates presented in Figure 2.7. As monitored for most cases,
accelerated degradation occurred within first 100 to 500 days.
Figure 2.7 Comparison of predicted and observed field settlement for bioreactor landfills,
(Hossain and Gabr 2005)
Chen et al (2010) presented the results of highly degraded specimens which performed
under optimum bio-degradation conditions and experienced further primary compression
after a 235 day test. Another compression test was also performed on fresh samples of
MSW, where the results for strain are shown as H 235 for decomposed waste, and H 3 for the
fresh MSW in Figure 2.8. The composition, unit weight, and the initial void ratio of the
prepared specimens were the same as for the highly degraded specimens. The duration of
each load step was 3 days.
32
Figure 2.8 Measured and predicted strains with stress applied to fresh waste specimen and
decomposed waste specimen (Chen et al 2010).
Chen et al (2010) reported the result of compression tests on the specimens with two
different unit weights (6.3 and12.1 kN/m3). The applied stress was 150 kPa and the
duration of the test was 45 days. Variations of strain with time for the specimens are
presented in Figure 2.9. Based on their results the compression rate reduced rapidly during
the first 24 hours. They recommended that 24 to 72 hours would be needed to complete the
primary compression of the tested MSW specimens. In addition, the variation in the rate of
MSW settlement with the logarithm of time is shown in Figure 2.10 as an approximately
linear trend.
33
Figure 2.10 Variation of strain with time (Chen et al 2010)
2.6.1 Introduction
Currently there are limited studies on in situ MSW shear behaviour. The shear strength of
MSW is generally determined using the Coulomb failure criterion. Great attention should
be taken while using the experience of shearing in soils to study MSW. Dixon and Jones
(2005) reported that waste includes compressible particles such as plastic that can support
large tensile strains and can vary with time through degradation. The results of using the
Coulomb criterion show that the shear strength increases with an increasing level of stress.
This is consistent with waste being considered as a frictional material. While the
engineering information of the response of municipal solid waste to monotonic and cyclic
loadings has improved, various important issues remained unresolved. Characterising the
shear strength of MSW is an important step in presenting reliable static and seismic
stability analyses of landfills, and the stress-strain response of MSW is essential to give
compatibility between the mobilised shear strength and deformation along potential failure
surfaces.
34
2.6.2 Background
Landva and Clark (1986) conducted a numbers of large-scale direct shear (DS) tests on
specimens of waste from different Canadian landfills. The DS device had plan dimensions
of 434 by 287 mm, and the specimens were sheared at about 1.5 mm/minute. The
interpreted value of cohesion (c) was between 10 and 23 kPa and the measured friction
angle (φ) varied between 24 and 42 degrees. When specimens of old waste from
Blackfoot/Burbank were tested for the first time, the cohesion and friction angle were
estimated to be 16 to 19 kPa and 38 to 42 degrees, respectively. The material was then
stored in plastic containers, and when tested a year later the shear strength had decreased
to16 kPa and 33 degrees, respectively. Initially, the authors stated that this could be a result
of decomposition. Later, Landva and Clark (1990) reported that the reduction in strength
was within the variability of the shear strength values and should not necessarily be
attributed to decomposition effects. Landva and Clark (1986) also reported data from
Edmonton, which included large amounts of plastic sheets. The relatively low friction angle
of 24 degrees was attributed to the fact that sliding largely occurred between plastic sheets
that were oriented parallel to the direction of the imposed shearing. They also reported that
the cohesion of the material should be attributed to the particles interlocking.
The results from the large scale CU Triaxial tests (diameter=300mm and height=600 mm),
conducted by Caicedo et al (2002b), are presented in Figure 2.11. The material did not
provide a distinct peak strength with an axial strain of nearly 15%. A pressure-phicometer
device was used for an in situ evaluation of the shear strength of MSW. This procedure
includes a probe which was placed in a borehole to the desired depth and then inflated
horizontal flaps were pushed into the waste using internal pressure. The strength of the
waste was estimated after an axial force was applied to the probe. The results of the in situ
pressure-phicometer tests were consistent with those from the in situ large scale direct shear
tests. Table 2.9 presents the results of MSW shear strength laboratory tests from literature.
35
Table 2.9 MSW shear strength laboratory tests results from literature
Initial unit
weight, Strength Shape of stress-
Type of Dimensions Failure criteria Reference
kN/m3 parameters strain plot
test (mm)
CU- C=16.8 kPa, Gabr and
71 7.4-8.2 20% strain Failure not reached
Triaxial φ=34º Valero (1995)
Drained C=0-27 kPa, 5% and 10% Continued strength Gabr and
63.5 10-12.1
direct shear φ=20.5º-39º strain increase Valero (1995)
Direct D=800, C=43 kPa, Peak shear Roughly Mazzucato et
7
36
38
Figure 2.12 Representative results from consolidated drained triaxial tests on partially
saturated MSW with unit weight of 12 kN/m3 and water content of 67% (Vilar and
Carvalho, 2002).
39
Figure 2.13 Representative results from consolidated drained triaxial tests on saturated
MSW with unit weight of 12 kN/m3 (Vilar and Carvalho, 2002).
40
Figure 2.14 Shear strength envelopes for triaxial specimens on MSW. Unit weight 12kN/m3
both saturated and unsaturated (Vilar and Carvalho, 2002).
41
Gomes et al. (2002) carried out 2 triaxial tests under a confining pressure of 100 kPa on
samples collected from the Santo Tirso landfill. The stress-strain results are presented in
Figure 2.15 and it shows a strain hardening response for strains up to 30% to 40%.The
authors did not report the unit weight or the size of the specimen. Mahler and De Lamare
Netto (2003) executed large scale direct shear tests (400 × 250 × 100 mm) on waste from
an area in Rio de Janeiro in Brazil, and particles greater than 20 mm were removed
beforehand. The results showed that the stress-strain curves were hyperbolic in shape and
demonstrated a post-peak drop in strength. Towhata et al. (2004) carried out triaxial tests (d
= 100mm and h = 200mm) with particles smaller than 10mm. The authors tested organic
waste from Germany and incinerated waste from Tokyo, Japan. The unit weight of the
specimens determined between 7.4 and 7.7 kN/m3. The organic waste did not reach peak
strength during tested deformation. The results revealed a linear stress-strain behaviour of
up to 20% axial strains. The authors attributed this behaviour to the stretching of plastic
sheets in the horizontal direction. To confirm this concept, the authors performed tests on
the same organic waste without any plastic sheets. Those specimens yielded at axial strains
of about 15%.
Figure 2.15 Stress-strain relationships from triaxial tests performed by Gomes et al. (2002)
Reddy et al. (2009) reported the results of small scale laboratory tests carried out on fresh
MSW collected from Orchard Hills Landfill in Illinois, USA. Based on 63 mm diameter
42
direct tests, the drained cohesion fluctuated from 31 to 64 kPa and the drained friction angle
varied from 26 to 30 degrees. In addition, the small scale CU triaxial tests (d=70 mm)
conducted on saturated MSW presented total strength parameters of c = 32 kPa and φ=12
degrees, and the effective strength parameters of 38 kPa and 16 degrees for a failure
criterion of 15% axial strain. The stress-strain response did not demonstrate a clear upward
curve as typically observed in large scale testing of MSW.
Bray et al. (2009) presented the results of extensive large scale laboratory testing of MSW
collected from Tri-Cities landfill in the San Francisco Bay Area of California. The samples
were reconstituted and subjected to monotonic loading in three different large scale testing
devices. The direct shear box was 300 mm by 300 mm (h = 180 mm), the triaxial device
was 300 mm in diameter (h = 610mm), and the simple shear device was 400 mm by 300
mm (h = 150 mm). A total of 23 direct shear tests, 27 triaxial, and 3 simple shear large
scale monotonic loading tests were carried out. Additional testing was conducted in a 71
mm diameter triaxial device. The influence of the test device, waste constitution, fibrous,
particle orientation, confining stress, rate of loading, stress path, stress-strain compatibility,
and unit weight on the shear strength of MSW were evaluated. Additional details are
presented in Zekkos (2005).
Triaxial compression tests on MSW samples with and without fibrous waste larger than 20
mm was conducted by Bray et al. (2009). As shown in Figure 2.16, the stress-strain
responses from triaxial compression tests displayed upward curvature only for specimens
that included fibrous waste larger than 20 mm. For these specimens the shearing surfaces
cut across fibres to consider their reinforcing effect. The failure surface in a typical triaxial
compression test on MSW is oriented at an angle of about 60 to 65 degrees from the
horizontal. Additionally, MSW specimens compress significantly during axial loading and
greater than 10 to 20% axial strains are required to mobilise friction angles of 30 degrees.
43
Figure 2.16 Responses of MSW in monotonic triaxial compression testing for specimens
with varying waste compositions (Bray et al. 2009).
Bray et al. (2009) showed from the results of direct shear tests on specimens that fibrous
waste was oriented at different angles to the shear plane, which demonstrated the
anisotropic nature of MSW. As presented in Figure 2.17, when particles of fibrous waste
were oriented parallel to the shearing surface they had an insignificant effect on the direct
shear response of the waste and any contribution from the fibrous waste materials was very
small. However, when shearing is constrained to cut across long, fibrous waste particles,
the shear resistance of MSW increases significantly and a strain hardening response is
observed as an upward curvature of the stress-displacement response. These observations
signified that the upward curvature in triaxial tests is related to a gradual contribution by
the fibrous materials in MSW when the shear plane cuts across their long axis.
44
Figure 2.17 Stress-displacement response for MSW specimens with plastic reinforcement
oriented at different angles at a normal stress of 50 kPa (Bray et al. 2009).
Bray et al. (2009) presented a non-linear shear strength envelope shown in Figure 2.18.a
which was independent of the amount of fibrous material in MSW and shows shearing
parallel to the preferred orientation of the larger fibrous particles within MSW. In addition,
Figure 2.18.b shows that the secant friction angle reduces with increasing confining stress.
The direct shear strength of the Tri-Cities landfill waste materials from this initial test series
was defined by:
W c V n . tan(MV ) (2.13)
where W is the direct shear strength of Tri-Cities MSW, c is the cohesion intercept, V n is the
total normal stress, and MV is a normal stress dependent friction angle given by
Vn
MV M 0 'M . log (2.14)
pa
where M 0 is the friction angle measured at a normal stress of one atmosphere, 'M is the
change in the friction angle over one log-cycle change of normal stress, and pa is the
atmospheric pressure (i.e., 101 kPa). The Tri-Cities landfill MSW direct shear test results
are best characterised by c = 15 kPa, M 0 = 41o, and 'M = 12o.
45
(a)
(b)
Figure 2.18 Direct shear strength of Tri-Cities landfill MSW: (a) curved strength envelope
for samples with varying waste composition, and (b) decrease in secant friction angle with
increasing normal stress assuming c = 15 kPa (Bray et al. 2009.)
Bray et al. (2009) also reported a strain-rate influence by shearing MSW specimens at
various rates of strain. As presented in Figure 2.19, the shear stress mobilised at a higher
strain rate was greater than that mobilised at the lower strain rate applied in this direct shear
test on MSW. The mobilised shear stress increased as the loading rate increased, by
approximately 10-15% per log cycle of strain rate. The effects of the strain rate were
greater for specimens with higher amounts of fibrous waste material.
46
Figure 2.19 Response of MSW with 62% less than 20 mm material in direct shear testing
loaded at two displacement rates (Bray et al. 2009).
Howland and Landva (1992) reported evaluations of MSW shear strength from a landfill
failure in New Jersey that happened after nearly 50 feet (15.24 m) of new waste was placed
over 15-year old waste in a 4 to 5 month period. The failure began in the underlying deposit
and developed up through waste. High amount of rain was reported over last three days
before the failure. The authors also back-calculated MSW shear strength based on a plate
load test that was executed on the OII landfill in southern California. The authors compared
the back-calculated strength estimates with large-scale laboratory direct shear tests
conducted by Landva and Clark (1990) and in situ direct shear tests carried out by
Richardson and Reynolds (1991). They reported that the back-calculated shear strength of
MSW from field data is significantly lower than the strength measured in the direct shear
tests. The reasons could be due to scale effects, uncertainties in the soil properties assumed,
and the assumption of a factor of safety of one. From these data, Howland and Landva
(1992) developed a lower bound estimate of c = 10 kPa and M = 23o.
Kavazanjian et al. (1995) presented several case histories, including the New Jersey landfill
failure and the plate load test at the OII landfill. The authors stated that the back-calculation
of the New Jersey landfill failure showed by Howland and Landva (1992) underestimates
the shear strength of MSW due to only a portion of the shear strength of the MSW was
47
mobilized and failure happened at small displacements. The authors also noted that the
assumption of FS = 1 was conservative for interpreting the OII landfill plate load test,
because the test was ended before the waste failed.
Kavazanjian et al. (1995) displayed shear strengths back-calculated results from case
histories and in situ direct shear testing results in Figure 2.20. They suggested a bilinear
shear strength envelope for MSW with parameters of c= 24 kPa and M = 0o for normal
stresses less than 30 kPa and c = 0 kPa and M = 33o for normal stresses greater than 30
kPa.
Augello et al. (1995) presented the results of an investigation into the seismic performance
of solid waste landfills during the 1994 North ridge earthquake. They estimated the
dynamic shear strength of MSW by assuming the Kavazanjian et al. (1995) unit weight and
shear wave velocity profiles for MSW. The strain-dependent shear modulus reduction and
material damping curves recommended by Kavazanjian and Matasovic (1995) for PI=200
were used. By assuming that FS = 1.2 and c = 0, Augello et al. (1995) back calculated the
static friction angles between 19 and 35 degrees from their limited equilibrium slope
stability analyses. Their back calculated dynamic friction angles were between 30 and 40
48
degrees from which they concluded that the dynamic shear strength of the MSW was higher
than its static shear strength. Augello et al. (1998) developed improved estimates of the
MSW properties using updated waste characterisations. The MSW shear wave velocity and
unit weight profiles of MSW recommended by Kavazanjian et al. (1996) were used.
Assuming that FS = 1.3 and c = 0, the back calculated static friction angles now ranged
from 25 to 41 degrees. Their back calculated dynamic friction angles at three unlined
landfills ranged from 27 to 45 degrees, which again indicated that the dynamic strength of
MSW was higher than its static strength. The authors recommended that the mid-range
dynamic friction angles of 35 to 38 degrees were most appropriate for use in engineering
practice, with M = 35o as their best recommended estimate.
2.6.6 Limitations
There are several limitations to fully understanding the shear strength of MSW, these
include:
x The influence of the moisture content of MSW has not been investigated completely
in the existing studies, particularly when the waste material is saturated. Excess pore
pressures can be generated throughout shearing because of the highly compressible
nature of waste material when it is almost saturated.
x The generation of landfill gas in saturated MSW can intensify the development of
excess pore water pressures (Merry et al. 2006) which can cause a significant
reduction in the shear strength. The ability to reduce the effective stress as a result
of generating excess pore pressure in saturated waste has not been studied
completely.
x Literature indicate that the water content does not have a significant effect on the
shear strength in the range of 10% to 25%, which is less than the field capacity of
most MSW materials.
x The influence of waste degradation on its shear strength has not been studied
sufficiently. The degree of degradation in waste is a difficult to estimate, and the
age, waste constitution, climate, water content, and landfill construction procedures
are also likely to be important.
49
x Researchers utilised various specimen sizes and maximum particle sizes and
different methods for preparing specimens, so a clear understanding of the stress-
strain response of MSW is slowed down by the lack of a consistent preparation and
testing methods for testing MSW.
x In many cases researchers have not reported all of the information required
interpreting the test data and as a consequence, there are significant inconsistencies
in the findings of many studies.
2.7.1 Introduction
The flow of water through waste in a landfill is important for two main reasons. First, water
has the potential to pollute the surrounding environment because liquids travel through the
waste and appear as leachate that dissolves and carries substances in suspension that could
cause contamination by interacting with the natural groundwater. Barriers help to prevent
this and by controlling leachate pressures, differences in head acting on the boundary of
barriers will be minimised. Changes in the distribution of waste permeability occur due to
anisotropy, heterogeneity, partial saturation, and variations in waste density. Secondly,
water flow has a significant effect on bringing landfill to a stable, non-polluting condition
in a controlled way. Municipal solid wastes have a large amount of bio-degradable
components and water is necessary for the biochemical decomposition of organic
substances (Pohland 1975; Leckie et al 1979; Klink and Ham 1982). The content of water
has increased and leachate flows are controlled in order to accelerate the degradation and
biological stabilisation of organic matter, releasing increased quantities of commercially
exploitable landfill gas (Tchobanoglous et al. 1977; Leckie et al, 1979).
According to Beaven et al. (2011) the flow and distribution of water in landfills are
complicated because:
x Waste is a heterogeneous material;
x There is a wide variety of particle and pore size and shapes, and types of solid
materials,
50
x Waste and landfill are anisotropic;
x The hydraulic properties of waste vary with stress (depth) and potentially over time
(through compaction and degradation);
x The presence of landfill gas within the waste interacts with and affects the flow of
water such that both saturated and unsaturated flow may occur.
Some of these complexities are found in other hydrogeological systems, while others are
unique to landfills. A combination of all these processes in one system makes
understanding, quantifying, and predicting the flow of liquid in waste landfills a difficult
and challenging subject area. Nonetheless, it is one that landfill operators need to
understand, especially as landfills become more actively managed to prevent pollution and
accelerate stabilisation.
In landfilled waste, the distribution and pore size of particles, and the heterogenic
composition and leachate chemistry of the waste complicate any determination of its
moisture retention and hydraulic properties. Predicting the distribution of moisture,
generation and flow of leachate in landfilled waste using unsaturated flow theory depends
on both the validity of the flow theory and the hydraulic properties of the landfilled waste.
Unsaturated flow has been extensively researched in the fields of soil physics, hydrology,
and geotechnical engineering. Although the unsaturated hydraulic properties of
conventional soils have been extensively investigated, the evaluation of landfilled waste is
severely limited (Korfiatis et al. 1984; McDougall et al. 1996).
Saturated waste hydraulics is important because it is related to all leachate extraction and
control measures at a landfill. Flow in saturated waste is almost characterised in terms of
Darcy’s law,
q k .i. A (2.15)
where q is the flow rate (m3/s), A is the cross-sectional area of flow (m2), i is the hydraulic
gradient and k is the hydraulic conductivity. k is related to the properties of the porous
medium and the permeant by
51
Jf
k K (2.16)
Kf
where K is the intrinsic permeability of the porous medium (m2), J f is the unit weight of
2.7.3 Background
Clearly stating a range of hydraulic conductivities of waste is not appropriate due to the
very wide range and it also covers factors such as the composition and density of waste
which affect the hydraulic conductivity in a predictable way. In addition, laboratory tests
have estimated the hydraulic conductivity in vertical flow, whereas in a landfill the
anisotropy caused by the deposition and layering of the waste, results in horizontal flow. It
will be much more complicated when the effects of channeling, gas generation, random
heterogeneity, and waste decomposition are taken into account.
Jain et al. (2006) collected a large data set from different studies on the hydraulic
conductivity of MSW that combine the results of field and laboratory tests. The hydraulic
conductivities of MSW presented in literature fluctuated between nearly 1×10-3 m/s and
1×10-9 m/s, although more typical values were in the range of 10-5 to 10-6 m/s.
One of the most systematic series of large scale tests on the hydraulic conductivity of MSW
has been reported by Beaven and Powrie (1995), Powrie and Beaven (1999) and Beaven
(2000), carried out in the Pitsea compression cell. The Pitsea cell accommodates a waste
sample 2 m in diameter and up to 2.5 m high. Overburden pressures are simulated by
applying a vertical stress via hydraulic rams acting on top of the waste. Typically, the
applied stress is increased in five or six stages to a maximum of 600 kPa, representing
landfill depths of up to 60 m. At the end of each compression stage, the bulk density,
drainable porosity and saturated hydraulic conductivity of the waste estimated. Tests were
performed on four different samples of domestic waste (DM3, PV1, DN1 and AG1) to
evaluate the effects of particle size reduction, degradation, and compression on the bulk
vertical hydraulic conductivity. DM3 was fresh, unprocessed waste; PV1 was fresh waste
that had been pulverised; DN1 was fresh waste that had been partly sorted and tumbled in a
52
drum using the Dano system; and AG1, which was a 25 year old partly degraded waste
containing a mixture of soil, crude waste and pulverised waste that had been recovered
from a depth of less than 5 metres from a landfill site. Further tests were carried out on a
sample of DN1 to investigate the effects of partial saturation and gassing. Full
characterisation analyses are given by Powrie and Beaven (1999) for sample DM3, by
Hudson et al. (2001) for waste DN1, and by Beaven (2000) for wastes PV1 and AG1. Raw
data of hydraulic conductivity, drainable porosity, and density at various vertical effective
stresses were given for samples DM3, PV1, and AG1 by Beaven (2000) and for sample
DN1 in conditions of high and low accumulation of gas, and with high and low pore water
pressures by Hudson et al. (2001). Figure 2.21 shows the permeability of all four wastes in
saturated conditions, plotted as functions of: (a) the vertical effective stress in the first
loading, (b) the drainable porosity, and (c) the dry density. Referring to Figure 2.21, the
following observations were reported by Beaven et al. (2011).
1. There is a single correlation for all samples between the logarithm of the vertical
hydraulic conductivity and vertical effective stress in the first loading. Differences in
hydraulic conductivity resulting from a reduction in particle size and degradation of the
waste are essentially second order, but appear to become more significant at higher vertical
effective stresses.
2. There are individual correlations between the logarithm of the vertical hydraulic
conductivity and density for each type of waste, with an essentially linear relationship
between the logarithm of the vertical hydraulic conductivity and the dry density.
3. There is a single correlation between the logarithm of vertical hydraulic conductivity and
the drainable porosity of the waste. Drainable porosity represents a measure of the size and
degree of connectivity of the voids, both of which will have a major influence on the bulk
hydraulic conductivity.
It is important to emphasise that this correlation is with the vertical effective stress in the
first loading step; if the waste is compacted prior to landfilling then an equivalent pre-
compaction stress must be estimated.
53
Figure 2.21 Vertical hydraulic conductivity against (a) the logarithm of the vertical
effective stress in first loading; (b) the drainable porosity; and (c) density, for four waste
types (data from Beaven 2000 and Hudson et al. 2001)
As can be seen in Figure 2.21c, the relationship between the hydraulic conductivity and the
dry density has been defined but they are different for each type of waste. Many researchers
have presented data relating the hydraulic conductivity to the waste density, but the effects
54
of waste composition and other variables makes any evaluation of the relationship difficult
for different types of wastes. Landva and Clark (1986, 1990), performed large scale
permeability tests at the surface of different landfills in Canada. The hydraulic conductivity
was estimated based on the rate of fall of the water level and flow nets appropriate to each
stage. Hydraulic conductivities ranged between 4×10-4m/s and 1×10-5 m/s. The unit weights
of the waste excavated from the pits was mostly between 10-14 kN/m3 but the correlation
between hydraulic conductivity and refuse density was poor due to variations in the
composition of materials between the different sites.
Table 2.10 Summary of tests relating the vertical hydraulic conductivity of MSW type
materials to dry density
Author(s) Apparatus Waste tested
Chen and Chynoweth (1995) 30 cm diameter rigid different types of processed
walled permeameter MSW, including a mix that
combined paper, plastics and
yard waste.
Bleiker et al. (1995) falling head tests in a cored samples recovered
flexible walled from depths of 32 - 36 m in
permeameter the Keele Valley landfill by
drilling using a 100mm
diameter auger
Staub et al. (2009) 20 cm diameter rigid samples of shredded MSW
wall permeameter. recovered from two French
landfill sites
Beaven (2000); Powrie and 2 m diameter rigid Various
walled compression
cell (Pitsea cell)
55
Figure 2.22 Summary of relationships between vertical hydraulic conductivity and waste
dry density (Beaven et al. 2011)
There are few studies in literature on the influence that waste degradation has on hydraulic
conductivity because obtaining a direct comparison between the fresh and degraded
conditions of a waste is difficult. However, Beaven et al. (2011) noted that a comparison of
the curves for the aged waste sample AG1 with the other data given in Figure 2.21
proposed that any effect of degradation on hydraulic conductivity is small in comparison
with the effect of an increase in stress. The graph of hydraulic conductivity against vertical
stress in the first loading lies within the range given by the tests on fresh wastes, while the
relationship between hydraulic conductivity and drainable porosity is similar. Figure 2.21c
shows that the dry densities of the aged waste are generally greater than the fresh waste
samples, which could be caused by degradation. The relationship between hydraulic
conductivity and dry density is different for each type of waste. Gas generation is a
consequence of waste degradation, and this will have its own effect on hydraulic
conductivity.
56
2.7.6 Waste Anisotropy
Hudson et al. (2009) investigated the effect of anisotropy on the hydraulic conductivity of
MSW type materials. A review of literature in Table 2.11 shows there is a very wide
variation in the ratio of horizontal to vertical hydraulic conductivities of MSW.
Data reported by Hudson et al. (2009) from tests on samples of a 20 year old waste
recovered from a landfill (AG1) and a fresh, Dano-processed waste (DN1), carried out in
the Pitsea cell, are reproduced in Figures 2.23 and 2.24 by Beaven et al. (2011) in terms of
the ratio kh:kv as a function of the applied stress. Determining kh from flow tests was
analytically quite complex and involved numerical modelling using MODFLOW, as
described by Hudson et al. (2009).
Figure 2.23 kh : kv versus applied stress for sample AG1. (Beaven et al. 2011)
57
Figure 2.24 kh : kv versus applied stress for sample DN1. (Beaven et al. 2011)
(unshaded markers indicate tests run in gas accumulated conditions)
All the tests results showed higher horizontal than vertical hydraulic conductivities, and
there was a general trend of an increase in the kh:kv ratios with an increase in stress. The
ratio was between 5 and 7 at a low stress (40 kPa) and turned to 10 at higher stresses (300
to 500 kPa). Sample DN1 displayed slightly greater kh:kv ratios than sample AG1. In Figure
2.24, data from tests performed in gas accumulated conditions are indicated by unshaded
markers. Although both vertical and horizontal hydraulic conductivity were considerably
lower in gas accumulated conditions than in saturated conditions, the kh:kv ratios for both
conditions are not clearly dissimilar. This finding recommends that the accumulation of gas
has no significant influence on the kh:kv ratio.
Buchanan and Clark (1997, 2001) showed that with increasing stress there was a slight
reduction in anisotropy, although the tests were restricted to the finer fraction of the waste
and low densities. They noted that highly treated wastes to smaller particle size may
demonstrate a smaller degree of anisotropy. The slightly higher k h:kv ratios achieved for the
fresh (DN1) waste than for the aged sample (AG1) may indicate a slight breakdown in the
anisotropic structure of waste with decomposition.
58
2.8 Improvement Techniques for MSW landfills
The stabilisation of landfills and waste disposal sites for structural and environmental
purposes has been performed through the application of current soil stabilisation and
ground improvement techniques. The deep dynamic compaction (DDC) technique is a
common ground improvement technique due to its relatively economical and easy
application. The deep dynamic compaction technique has been carried out with success in a
large range of soils including MSW. Zekkos et al. (2012) reviewed 64 case histories
regarding the effectiveness of the DDC on MSW landfill sites. Their results indicated that
the depth of improvement is smaller in MSW compared to cohesionless soils. In addition,
the settlement caused by DDC depends on the applied energy and it is in the range of 5% to
25% of the MSW thickness. With deep dynamic compaction large voids reduce and
afterward other techniques such as fly ash-lime grouting can further reduce the remaining
smaller voids. Blacklock (1987) showed that the lime slurry pressure injection can be
utilized for the control of methane gas. Moreover, the lime/fly ash slurry injection has
significant effects on protecting groundwater, neutralising leachate, and for placing curtain
walls to prevent leachate migration. Replacement of cement with by-product materials such
as fly ash can decrease the stabilisation expenses. Many researchers reported the
application of fly ash in geotechnical projects (e.g. Kawasaki et al. 1981; Kitazume et al.
2000; Kehew 1995). According to Horpibulsuk et al. (2012), from an economic and
environmental viewpoint, rich materials in slaked lime (Ca(OH)2, calcium hydroxide), can
be treated together with pozzolanic materials, such as fly ash, to develop a cementitious
material.
As explained earlier in improvement techniques for MSW landfills, the deep dynamic
compaction (DDC) technique is a common ground improvement technique due to its
relatively economical and easy application. With deep dynamic compaction large voids
reduce and afterward other technique such as fly ash-lime grouting can further reduce the
remaining smaller voids. There are no research conducted on treatment and improvement of
MSW landfills with chemical agents such as fly ash and lime. However, several studies
59
were conducted to treat and stabilise different types of the problematic soils using lime and
fly ash combination. Understanding the behavior of fly ash and lime on stabilisation of
different types of soils can provide better outlook for their application on waste materials.
Following sections briefs the application of fly ash and lime in several types of soils.
Several studies were performed to evaluate the soil stabilisation process using lime (either
CaO or Ca(OH)2) (Prakash and Sridhran1989; Wild et al. 1993; Bell 1996; Rajasekan et al.
1997; Rajasekan and Rao 1998, Burkart et al. 1999, Qubain et al 2000, Weber 2001, Yusuf
et al. 2001, Ismail 2004; and Ampera and Aydogmus 2005).
Qubain et al. (2000) mentioned the benefits of sub-grade lime stabilisation in the design of
a highway pavement in Pennsylvania. The plan included widening and completely
reconstructing 21 km of Highway. Field investigations showed that the sub-grade was
clearly homogeneous and composed of medium to stiff clayey soils. To protect against
softening as a result of rain, modifying the soil with lime has traditionally been used
because it improves the strength of clay by hydration, flocculation, and cementation. The
first and second mechanisms occur immediately after lime is added, while cementation
occurs over a long period of time. Qubain et al. (2000) examined the first and second
mechanisms. Experimental programs were carried out to demonstrate the immediate
advantage of lime stabilisation. Both treated and natural clayey samples were under the
resilient modulus and California bearing ratio tests, and the specimens treated with lime
were not allowed to cure. The results showed a significant increase in strength, a result
which can lead to a reduction in the thickness of the pavement and a subsequent saving in
costs. Weber (2001) investigated the influence of curing and degree of compaction on loss
loam stabilised by applying various additives. He achieved the best results under condition
of moist, atmospheric storage. In this condition, tempering of the stabilised specimens was
delayed because the pH-value in the pore water changed. He also noticed that the change of
the degree of compaction of the stabilised specimens affected their behaviour and
compaction caused brittle failure.
60
Ismail (2004) treated materials related to road construction by applying lime (10%), cement
(10%), and lime/cement (2.5%/7.5%). He estimated the consistency limits, compaction
properties, and shear- and uniaxial strength, and concluded that the cohesion and the
friction angle of the treated soil increased. With the materials treated with lime, cohesion
decreased with the curing time. He also reported that the loss of weight during the freezing
and thawing test was low and depended on the type of material.
Ampera and Aydogmust (2005) treated Chemnitz clayey soil (A-7-6 Group - AASHTO)
using lime (2, 4 and 6%) and cement (3, 6 and 9%). They carried out compaction,
unconfined compressive strength, and direct shear tests on untreated and treated specimens,
and remarked that the strength of soil treated with cement was generally greater than soil
treated with lime. They also reported that lime stabilisation is better suitable to clayey soils.
Fly ash by itself has a slightly cementatious characteristic but it reacts chemically with
moisture to form a cementatious mixture used to improve the strength and compressibility
of soil. It has a long history as an engineering material and has been successfully utilized in
geotechnical applications. Several studies were performed to evaluate the soil stabilisation
procedure using a by-product material such as fly ash class F or class C (Lee and Fishman
1992, Turner 1997, Sahu 2001, Acosta et al. 2002, Edil et al. 2002, Şenol et al. 2002, and
Thomas and White 2003).
Edil et al. (2002) carried out a field investigation using options such as fly ash, bottom ash,
foundry slag, and foundry sand that could be used for construction over soft sub-grade
soils. A class C fly ash was used for one test section, and an unconfined compression test
revealed that 10% fly ash (on the basis of dry weight) was enough to provide appropriate
strength for construction on the sub-grade. The results obtained before and after placing fly
ash, by testing undisturbed samples in the laboratory using a soil stiffness gauge (SSG), and
a dynamic cone penetrometer (DCP) test in the field. The unconfined compressive strength,
soil stiffness, and dynamic cone penetration of the soil before the placement of fly ash
ranged between 100 - 150 kPa, 4 - 8 MN/m², and 30 -90 mm/blow, respectively. After the
addition of fly ash the unconfined compressive strength was as high as 540 kPa, the
61
stiffness ranged from 10 to 18 MN/m2, and the Dynamic Penetration Index (DPI) was less
variable and ranged between 10 and 20 mm/blow. A CBR of 32% was reported for the
stabilised sub-grade, which is rated as “good” for sub-base highway construction. The CBR
of untreated sub-grade was 3%, which is rated as “very poor” according to Bowles, 1992.
Acosta et al. (2002) evaluated self cementing fly ash as a sub-grade stabiliser for Wisconsin
soils. A laboratory testing program was carried out to estimate the mechanical properties of
fly ash and to evaluate how it can improve the engineering properties of a range of soft,
sub-grade soil. Seven soils and four types of fly ash were considered for the study. Samples
of soil were prepared with different contents of fly ash (i.e., 0, 10, 18, and 30%), and
compacted with different water contents (optimum water content, 7% wet of optimum
water content, and 9 to 18% wet of optimum water content). Three types of tests were
performed, a California bearing ratio test, a resilient modulus test, and an unconfined
compressive strength test. The soils selected represented poor sub-grade conditions with a
CBR ranging between 0 and 5 in their natural condition. A substantial increase in the CBR
was achieved when fly ash was mixed with the soils. Specimens prepared with 18% fly ash
and compacted at the optimum water content improved the best, with a CBR ranging from
20 to 56. Specimens prepared with 18% fly ash and compacted at 7% wet of optimum
water content showed a significant improvement compared to untreated soils, with a CBR
ranging from 15 to 31 (an approximately average CBR gain of 8 times). On the other hand
there was less improvement when the specimens were prepared with 18% fly ash and
compacted in a very wet condition (CBR ranging from 8 to 15). Mixtures of soil and fly ash
with 18% fly ash and compacted at 7% wet side of the optimum water content had a similar
or higher modulus than untreated specimens compacted at the optimum water content. The
resilient modulus of specimens compacted in significantly wet conditions had a generally
lower module than the specimen compacted at optimum water content. The resilient
modulus increased as the curing time increased. The resilient modulus of specimens 18%
fly ash and compacted at 7% wet of optimum water content was 10 to 40% higher after 28
days of curing than those after 14 days of curing. The unconfined compressive strength of
the soil and fly ash mixtures increased with the volume of fly ash. Specimens of soil with
10% and 18% of fly ash and compacted at 7% wet side of the optimum water content had
62
an unconfined compressive strength 3 and 4 times higher than the original untreated
specimen of soil compacted at 7% wet side of the optimum water content. The CBR and
resilient modulus data were used in the design of a flexible pavement. Data developed
from stabilised soils showed that a reduction of approximately 40% in the base thickness
could be achieved when 18% fly ash was used to stabilise a soft sub-grade.
Şenol et al. (2002) studied the use of self cementing class C fly ash for stabilising the soft
sub-grade of a city street in cross plains, Wisconsin, USA. Both the strength and modulus
based approaches were applied to estimate the optimum mix and to determine the thickness
of the stabilised layer. Stabilised samples were prepared by mixing fly ash at three different
contents (12, 16 and 20%) with varying water contents. The samples were subjected to
unconfined compression tests after 7 days of curing to develop the water content strength
relationship. The study showed that engineering properties such as the unconfined
compressive strength, CBR, and the resilient modulus increased substantially after
stabilisation with fly ash. This stabilisation process was sensitive and required strict control
of the moisture content. The impact of delayed compaction that commonly occurs in field
construction was evaluated by Şenol et al. (2002). One set of the samples was compacted
immediately after being mixed with water while the other set was compacted two hours
later. The results showed that the loss of strength due to delayed compaction was
significant and therefore it must be considered in the design and construction. CBR and
resilient modulus tests were conducted and used to determine the thickness of the stabilised
layer in the design of this pavement.
Thomas and White (2003) used self cementing fly ash (from eight different sources) to treat
and stabilise five different types of soil (ranging from ML to CH) in Iowa for road
construction applications. They investigated various geotechnical properties (under
different curing conditions) such as compaction, unconfined compressive strength, wet/dry
and freeze/thaw durability, the effect of curing time, etc. They reported that self cementing
Iowa fly ash can effectively stabilise Iowan soil. Indeed the unconfined compressive
strength, gain in strength, and the CBR value of stabilised soils increased with curing time.
A mixture of soil and fly ash secured under freezing conditions and soaked in slaked water
63
were unable to be tested for strength. They also noticed that stabilised paleosol exhibited an
increase in its freeze/thaw durability, whereas stabilised Turin loess failed the test.
The effect of fly ash on expansive soil was studied by Erdal Cokca (2001). Expansive soils
can potentially be stabilised efficiently by cation exchange using fly ash. Erdal Cokca
performed investigations using Soma Fly ash and Tuncbilek fly ash and mixed it with
expansive soil at 0-25%. Specimens with fly ash were cured for 7 days and 28 days, after
which they were subjected to Oedometer free swell tests. The experimental results showed
that the plasticity index, activity, and swelling potential of the samples reduced with an
increasing amount of stabiliser and curing time. In addition, the optimum amount of fly ash
required to lower the swell potential was found to be 20%. He also reported that variations
in the physical characteristics and swelling potential can be due to additional silt size
particles, chemical reactions that cause immediate flocculation of the clay particles, and the
time dependent pozzolanic and self hardening properties of fly ash. Furthermore, he
concluded that both high calcium and low calcium class C fly ash can be recommended as
effective stabilising agents for improving expansive soils.
A similar study was conducted by Phanikumar and Sharma (2004), who studied the effect
of fly ash on the engineering properties of expansive soil in an experimental program. The
influence on parameters such as the free swell index (FSI), swell potential, swelling
pressure, plasticity, compaction, strength and hydraulic conductivity of expansive soil was
studied. Expansive soil was mixed with 0, 5, 10, 15, and 20% of fly ash on a dry weight
basis. They concluded that increasing the volume of fly ash decreased the plasticity and
reduced the FSI by about 50% when 20% of fly ash was added. They also observed that the
hydraulic conductivity of expansive soils mixed with fly ash decreased as the volume of fly
ash increased due to an increase in the maximum dry unit weight with an increase in the
content of fly ash. They reported that when the content of fly ash increased there was a
decrease in the optimum moisture content and an increase in the maximum dry unit weight,
while the undrained shear strength of the expansive soil blended with fly ash increased with
an increase in the amount of fly ash.
64
2.9.3 Lime/Fly ash Stabilisation
Several studies were conducted to treat and stabilise different types of problematic soils
using a combination of lime and fly ash (Nicholson and Kashyap 1993; Nicholson et al.
1994; Indraratna et al. 1995; Virendra and Narendra 1997; Shirazi 1999; Muntuhar and
Hantoro 2000; Lav A. and Lav M. 2000; Cokca 2001; Consoli et al. 2001, Nalbantoglu
2001, Nalbantoglu and Tuncer 2001, Yesiller et al. 2001, Nalbantoglu and Gucbilmez
2002, Zhang and Cao 2002, Beeghly 2003, and Parson and Milburn 2003).
Zhang and Cao (2002) performed an experimental program to investigate the individual and
mixed effects of lime and fly ash on the geotechnical properties of expansive soil. Lime and
fly ash were added to samples of expansive soil at 4-6% and 40-50% by dry weight of soil,
respectively. The testing specimens were estimated and evaluated for chemical
composition, distribution of grain size, consistency limits, compaction, CBR, free swell,
and swell capacity. The influence of adding lime and fly ash on reducing the swelling
potential of the texture of an expansive soil was reported. In fact the texture of expansive
soil changed dramatically when lime and fly ash were mixed in. Moreover the plastic limit
increased by mixing lime and the liquid limit decreased by mixing fly ash, and caused a
reduction in the plasticity index. As the amount of lime and fly ash was increased there was
an apparent reduction in the maximum dry density, free swell, and the swelling capacity
under 50 kPa of pressure. There was also a corresponding increase in the percentage of
65
coarse particles, optimum moisture content, and in the CBR value. They concluded that
expansive soil can be successfully stabilised with lime and fly ash.
Beeghly (2003) evaluated the use of lime and fly ash together in stabilising subgrade soil
(silty and clayey soils) and the coarse granular aggregate base beneath a flexible layer of
asphalt or rigid layer of concrete. He reported that lime by itself stabilised clay soils but a
combination of lime and fly ash lowered the plasticity of soils. He also reported that the
unconfined compressive strength and CBR values of soils treated with lime and fly ash
together are higher than when only treated with lime.
Beeghly (2003) also concluded that soaking the treated specimens caused a 15-25%
reduction in unconfined compressive strength, and the mixture of lime/fly ash resulted in
50% savings on the cost of materials compared to using Portland cement for stabilisation.
Parson and Milburn (2003) conducted a series of tests to evaluate the stabilisation process
of seven different soils using lime, cement, class C fly ash, and an enzymatic stabiliser.
They estimated the Atterberg limits and unconfined compressive strengths of the stabilised
soils before and after conducting durability tests (freeze/thaw, wet/dry, and leach testing),
and reported that soils stabilised with lime and cement improved more than soils treated
with fly ash. In addition, the enzymatic stabiliser did not improve the soils very much
compared to other stabilising agents such as cement, lime, and fly ash.
Horpibulsuk et al. (2012) reported that Calcium Carbide Residue (CCR) and fly ash are
both waste products and the mixture produces a cementitious material because CCR
contains a lot of Ca(OH)2, while fly ash is a pozzolanic material. They investigated the
possibility of using this cementitious material (a mixture of CCR and Fly ash) to improve
the strength of problematic silty clay in northeast Thailand. They observed that the
maximum strength of the stabilised silty clay occurred at approximately the optimum water
content for different amounts of binder, ratios of CCR:Fly ash, and the curing times.
Figure 2.25 shows that replacing CCR with less than 30% fly ash increased the dry unit
weight of the stabilised samples. Moreover, Figures 2.26 and 2.27 show the improvement
66
in strength in blended CCR-stabilised clay at the optimum water content for 5% and 10%
binder at different ratios of CCR:Fly ash respectively.
Figure 2.25 Compaction curves of blended CCR-stabilised clay (Horpibulsuk et al. 2012)
Figure 2.26 Strength development in blended CCR-stabilized clay at OWC and 5% binder
for different CCR:Fly ash ratios (Horpibulsuk et al. 2012)
67
Figure 2.27 Strength development in blended CCR-stabilized clay at OWC and 10% binder
for different replacement ratios (Horpibulsuk et al. 2012)
Brooks et al. (2011) stabilised two weak soils from southeastern Pennsylvania using coal
fly ash and limestone dust. The results of the study revealed that the properties of these
soils were improved significantly by stabilisation with additives, apart from the UCS of one
soil. The improvement consisted of a reduction in plasticity and an increase in the
California bearing ratio. The results also showed that the maximum dry density of the soil-
additive mixture decreased with an increase in amount of additive, whereas the optimum
moisture content of the mixture increased with an increase in amount of additive. The
authors attributed the reduction in the maximum dry density to a cation exchange reaction
in the mixture which caused flocculation and agglomeration of the clayey particles of the
soil, resulting in larger particles with larger voids in between particles which reduced the
weight-volume ratio of the mixture. Their results are presented in Figure 2.28 and show the
UCS of the soils with the blended admixtures compared to the untreated soils. Generally, it
was observed that the UCS values for all the soil-additive mixtures were higher than the
control group at both early and latter ages.
68
Figure 2.28 Unconfined compressive strength of soils stabilised with a blend of 15% coal
fly ash and 3% limestone dust relative to those of control soils (Brooks et al. 2011)
Singh et al. (2010) conducted a laboratory study to evaluate the effectiveness of lime and
Class C fly ash (CFA) as soil stabilising agents for a sulfate containing A-7-6 type CL
Oklahoma soil. The effect of different percentages of lime and CFA on the engineering
properties of stabilised soil was examined. Using different percentages and 28 day
unconfined compressive strength tests (UCS) and the pH value, the optimum amount of
lime (5%) and CFA (16%) were determined. Cylindrical samples were compacted at the
optimum moisture content (OMC) and cured for 28 days for the resilient modulus and UCS
tests. Figures 2.29 and 2.30 show that the UCS and resilient modulus of the soil increased
substantially with the addition of both lime and CFA compared to those of raw soil at
OMC. They reported that the resilient modulus increased from 50 MPa of raw soil to 250
MPa of its stabilised counterpart, and the UCS improved from approximately 100 kPa to
1000 kPa.
Figure 2.29 UCS of lime and CFA mixes (Singh et al. 2010)
69
Figure 2.30 Resilient modulus for stabilised soils (σd = 42 kPa and σ3=13 kPa) (Singh et
al. 2010)
Han-bing et al. (2009) studied the mechanical characteristics of fly ash soil, fly ash, and
silty clay that had been blended in proportions of 1:4, 1:2, and 1:1 by dry weight
respectively, and with water contents of 15%, 17%, and 19%, to test the static characteristic
of fly ash soil. The test results shown in Figure 2.31 indicate that the compressive strength
of fly ash soil with three different ratios of lime soil decreased when the water content
increased, but when the ratio of lime and soil was1:2 the compressive strength was greater
than the others. The compressive strength of fly ash soil (1:2) slowly decreased when the
water content increased, and was more than 150 MPa when the water content ranged from
15% to 23.5%. Furthermore, based on Figure 2.32, they reported that the deformation
modulus of fly ash soil (1:2) decreased with an increase in the water content, but it was still
more than 3 MPa.
Figure 2.31 Water content versus compressive strength (Han-bing et al. 2009)
70
Figure 2.32 Water content versus deformation modulus (Han-bing et al. 2009)
Shao et al. (2008) illustrated the influence of adding fly ash to ordinary Portland cement
using the cement mixed method. As shown in Figure 2.33, they reported that the
unconfined compression strength increased slowly when the fly ash was less than 8%, but it
increased considerably when fly ash content was reached 12%. However, when the fly ash
content exceeded 12%, the unconfined compressive strength decreased as fly ash content
increased until the curing time reached 60 days. They concluded from the unconfined
compression test that the optimal fly ash content for soft lagoonal clay was 12%.
Figure 2.33 Relationships between unconfined compressive strength and fly ash content
(Shao et al. 2008)
71
Kumar et al. (2007) conducted an experimental program to study how effectively lime
could stabilise the geotechnical characteristics of mixtures of fly ash and soil. An Indian
fly ash was mixed with expansive soil in different proportions and then the geotechnical
characteristics of specimens of fly ash and soil, and specimens of lime and soil were
investigated. Lime and fly ash were added to an expansive soil at ranges between1–10%
and 1–20%, respectively. The test specimens were subjected to compaction tests and
unconfined compression tests and then cured for 7, 14, and 28 days after which they were
subjected to further unconfined compression tests. With reference to Figure 2.34, they
reported that with an increase in the content of lime the maximum dry density of soil-lime
mixes decreased and the optimum moisture content increased. This reduction in density
was greater with lower percentages of lime. Furthermore, based on the results shown in
Figure 2.35, they showed that when fly ash was added to the mixture of soil and lime the
maximum dry density decreased further and the optimum moisture content increased.
Figure 2.34 Dry density versus lime (%) at different % of fly ash (Kumar et al. 2007)
72
Figure 2.35 Optimum moisture content (%) versus lime (%) at different % of fly ash
(Kumar et al. 2007)
The results from Kumar et al. (2007) regarding the effects of the contents of lime and fly
ash on the soaked and unconfined compressive strength of specimens of soil and lime after
7,14, and 28 days of curing, are shown in Figures 2.36 to 2.38. It is noted that the curing
time did not produce much of an increase in strength with up to 4% of lime. They thought
this could be because almost all the lime is taken up by the clay fraction in the soil at the
early stages, leaving very little to react with silica and alumina to produce pozzolanic
reactions. The increase in strength with the curing period after 6% of lime was added
indicates that some of the lime is available for pozzolanic reactions. They also noted that
expansive mixtures of soil and lime containing more than 8% lime decreased in strength,
which could be attributed to the platy shapes of the unreacted particles of lime in the soil
mass. They also reported that the strength of the mixtures of fly ash-soil-lime increased
with an increasing curing time and the unconfined compressive strengths of fly ash-soil
lime mixtures after 7, 14, and 28 days of curing were higher than those with respective soil-
lime samples. The optimum value of fly ash and lime may be taken as 15% and 8%,
respectively.
73
Figure 2.36 Variation of unconfined compressive strength with % of lime for different % of
fly ash (after 7 days curing) (Kumar et al. 2007)
Figure 2.37 Variation of unconfined compressive strength with % of lime for different % of
fly ash (after 14 days curing) (Kumar et al. 2007)
Figure 2.38 Variation of unconfined compressive strength with % of lime for different % of
fly ash (after 28 days curing) (Kumar et al. 2007)
74
2.10 Summary
In order to improve the geotechnical properties of closed landfill sites for redevelopment
purposes, understanding the characteristic, behaviours and properties of MSW are essential.
Summary of some important engineering properties of MSW such as unit weight, shear
strength, compressibility and permeability, based on current literature review, are presented
in this section. Finally current improvement techniques for stabilising weak soils and closed
landfill sites using fly ash and lime are briefly explained. The deep dynamic compaction
(DDC) technique is a common ground improvement technique for stabilising landfill sites.
With deep dynamic compaction large voids reduces and afterward other technique such as
chemical stabilisation with fly ash-lime require reducing further the remaining smaller
voids. There are no research conducted on treatment and improvement of MSW landfills
with chemical agents such as fly ash and lime. However, several studies were conducted to
treat and stabilise different types of the problematic soils using lime and fly ash
combination.
x Unit Weight
Evaluating the engineering properties and behaviour of MSW is very difficult due to
variety of combined materials and the influence of waste structure. Estimating the MSW
unit weight has a significant effect on landfill engineering performance. Based on analysis
of available laboratory and field data, the MSW unit weight profile presented a hyperbolic
equation for individual landfills. Using field data and large-scale laboratory data, the factors
that affect the unit weight of MSW have been studied. The available data indicates that the
MSW unit weight is affected by the waste composition and compaction effort applied when
first placed and the effective confining stress acting on the waste, which may be
represented by a depth term. Time under confinement also influences the MSW unit
weight. A hyperbolic model that captures the key factors in determining the unit weight of
MSW at a particular landfill has been developed and calibrated with the available field
data. To establish the unit weight profile of MSW for a specific landfill with confidence, in-
situ unit weight data is required. Reliable in-situ large-scale unit weight testing can be
executed through the use of shallow test pits to establish the near-surface initial unit weight
75
of the MSW and through the use of large-diameter borings to establish MSW unit weight at
depth.
x Shear Strength
Researchers believe that the Mohr-Coulomb strength criterion can be used to describe the
shear strength of MSW. The shear strength of MSW is initially stress dependent (i.e.,
frictional), specifically at higher confining pressures, however it has significant strength at
low confining pressures (i.e., apparent cohesion) and it attributes mainly to the fibrous
component of the waste. The shear strength used to characterise MSW depends on the
specimen preparation procedures, testing conditions, and strength criterion used. Large-
scale direct shear tests (300 mm × 300 mm) yield a logical estimate of MSW shear strength.
Undisturbed and reconstituted large-scale direct shear tests on MSW present similar shear
strength results. The monitored differences in the stress-strain response of MSW in triaxial
compression tests and in direct shear tests can be described by the strain-hardening effect of
fibrous waste particles, while shearing modes involves larger particles. MSW acts as a
reinforced material. Most of research that performed with large scale direct shear boxes on
MSW estimates the shear strength parameters of c and M between 0 and 50 kPa and
between 27o and 41o, respectively. The most recent shear strength parameters estimation by
Bray et al. (2009) suggests that c = 15kPa, M = 36o and 'M = 5o (explained in section
2.6.5). This MSW shear strength parameters is suitable for shearing parallel to the long axis
of the fibrous material within the waste and for waste with moisture content less than field.
A large number of triaxial compression tests results present that the MSW shear strength
mobilised at 5% strain can be characterised by a secant friction angle between 34o and 44o
for confining pressures up to 400 kPa. This mode of shearing cuts across the fibrous waste
particles, therefore, it is higher than direct shear test results when shearing did not include
the fibrous waste particles.
x Compressibility
Mechanisms resulting in settlement of waste include physical compression and creep, and
decomposition due to biodegradation of organic components. For simplicity, the total
76
settlement of a MSW landfill can be taken as a combination of primary and secondary
compressions. The primary compression includes the physical compression of components
and consolidation. The secondary compression includes all creep effects and those relating
to degradation. Researchers have proposed various settlement prediction models to be
applied on MSW landfills. Most settlement models can be classified into four broad
categories: soil-mechanics based models, rheological models, empirical models, and
biodegradation-induced settlement models. For each category, one appropriate model has
been adopted and explained in this chapter. The selected models are: (i) from the
rheological model category, the Gibson and Lo model (Edil et al. 1990); (ii) from the
empirical model category, the hyperbolic function model (Ling et al. 1998) and the power
creep law model (Edil et al. 1990); (iii) from the soil mechanics-based model category, the
Bjarngard and Edgers (1990) Model and (iv) from the biodegradation-induced model
category, the biological model (Hossein and Gabr 2005).
x Permeability
There has been a reasonable amount of research and understanding of saturated waste
hydraulics. Darcy’s law is usually applied and appears satisfactory for most cases. Changes
in the effective stress on the first loading appears to be the dominant influence on
hydrogeological properties, especially the hydraulic conductivity, with secondary
influences coming from the landfill gas, differences in waste composition and possibly
degradation. The effect of degradation on the hydraulic conductivity is small in comparison
with the effect of an increase in the stress. Waste anisotropy and heterogeneity are also
other important considerations. With increasing the stress, slightly reduction in anisotropy
can be observed. A review of literature shows a very wide variation in the ratio of
horizontal to vertical hydraulic conductivities of MSW. There is a single correlation for all
samples between the logarithm of the vertical hydraulic conductivity and the vertical
effective stress in the first loading, the density of each waste type and the drainable porosity
of the waste. Hydraulic conductivities of MSW presented in literature fluctuate between
nearly 1×10-3 m/s and 1×10-9 m/s, although more typically values are in the range of 10-5 to
10-6 m/s.
77
x Chemical Stabilisation with Fly ash and Lime
The stabilisation of landfills and waste disposal sites for structural and environmental
purposes has been performed through the application of current soil stabilisation and
ground improvement techniques. The deep dynamic compaction (DDC) technique is a
common ground improvement technique due to its relatively economical and easy
application. The deep dynamic compaction technique has been applied with success on a
large range of soils including MSW. Upon deep dynamic compaction, large voids reduce
and afterward other technique such as fly ash-lime grouting can further reduce the
remaining smaller voids. Replacement of cement with by-product materials such as fly ash
can decrease the stabilisation expenses. Many researchers reported the application of fly
ash in geotechnical projects (e.g. Kawasaki et al. 1981; Kitazume et al. 2000; Kehew 1995).
From an economic and environmental viewpoint, rich materials in slaked lime (Ca(OH)2,
calcium hydroxide), can be treated together with pozzolanic materials, such as fly ash, to
develop a cementitious material. There is not any thorough research, conducted on
treatment and improvement of MSW landfills with chemical agents such as fly ash and
lime. However, several studies have been conducted to treat and stabilise different types of
the problematic soils using lime and fly ash combination. Understanding the behaviour of
fly ash and lime on stabilisation of different types of soils can provide better outlook for
their application on waste materials.
Reported laboratory test data, using lime and fly ash as a combination or separately,
showed that the compression index of both expansive and non-expansive clays decreased
and the coefficient of secondary consolidation also decreased. Class C fly ash and lime
stabilisation reduced the compressibility of soil due to new cementing products in the
stabilised specimens. The unconfined compressive strength of organic soils increased using
fly ash, but the amount of increase depends on the type of soil and characteristics of the fly
ash. Furthermore, resilient modules of organic soils significantly improved. The increases
in strength and stiffness are resulted from pozzolanic reactions. The reduction in water
content, caused by the addition of dry fly ash, also contributes to the strength gain.
78
CHAPTER 3
3 Geotechnical Characterisation of the Collected
MSW Samples
3.1 Introduction
The characterisation and classification of municipal solid waste (MSW) for geotechnical
engineering purposes is a challenging task. There are no standard procedures, limited
experience throughout the world, and limited interaction among practitioners that could
lead to improved characterisation procedures. In this chapter, recommended measures for
characterisation and classification of MSW are discussed. The field sampling procedures
and the basic properties of the collected MSW samples are also described.
Definitions of landfill vary greatly among the states in Australia. Two waste acceptance
classes of solid waste landfill have been adopted in order to permit reasonable comparison
of requirements and practice among states. These are summarised at Table 3.1 below.
79
Table 3.1 Two basic landfill classes
Based on the national landfill survey conducted by the Waste Management Association of
Australia (WMAA, 2007), the approximate number of licensed landfills in each of the two
landfill waste-acceptance classes is presented in Table 3.2. It shows that Western Australia
and Queensland have the highest number of landfills in Australia. New South Wales, South
Australia and Victoria are the next, based on available number of landfills respectively.
Source: Compiled by WCS from data in the WMAA National Landfill Survey (2007)
80
3.4 Sampling Permission
In current research, it was necessary to take the samples from the field in order to conduct
laboratory tests. Therefore, permission was sought from several councils to access their
landfill sites. Due to a high concentration of gases and the unpredictability of the contents
residing within the landfill and safety issues, excavation in a landfill site is generally
considered dangerous so most requests were not approved by the councils due to safety
issues. Many of these sites were located in urban areas and digging a test pit could cause
environmental problems in residential areas. Finally, after numerous contacts were made,
and after conducting a risk assessment and addressing safety issues, the permission for
sampling from Bankstown landfill was received. A short history of some of the landfill
sites that were visited is reported below.
Figure 3.1 displays Blackman Park is located in Lloyd Rees Drive Lane Cove West.
Tipping of garbage was started on 7 February 1952 and by March 1954 had been extended
almost to the tidal flats of Stoney Creek. By the early 1970’s, a playing field and dressing
room had been established on part of the site. However, by the 1973 when additional areas
for tipping were required, council removed the playing field to allow an increase in the
depth and extent of fill, and the tipping was then continuing for a proposed four to five
years. Based on the information received from lane cove council, timeline of events
happened as following:
x 1977 initial planning to close the tip and create playing fields started.
x The tips use was scaled back and restricted to residents use on weekends in 1979,
and closing the tip permanently on 28 January 1980.
x The park is named after John Harold Blackman, Mayor of Lane Cove in 1952
x The playing fields and surrounding of Blackman Park were built in the 1980’s on a
former Council waste tip.
81
Figure 3.1 Blackman park - waterlogged after rainfall
Figure 3.2 shows Wangal Park, which is located on Cheltenham Road, Burwood, NSW. It
has limited environmental impacts on the community. According to the Burwood Council
Report (2010), this landfill site was capped in 2004. This landfill site was originally a brick
pit, which means limited organic materials were dumped in the landfill and most of the
material were construction and demolition wastes. Accordingly, there has been a very
limited gas emission and leachate from the site. Burwood Council built some gas trenches
around the site to detect the carbon dioxide (CO2) and methane (CH4) emission. The gas
trench is consisted of a plank with membrane, filled with gravel inside this system. There is
also a pipe which collects the gas from the landfill. The council continuously monitors the
gas problem till no more gas comes out from the site. Burwood Council also monitors the
environment water quality regularly.
82
Figure 3.2 Wangal park - Wetland Surrounded by Fence
Figure 3.3 presents Dartford Road landfill, which is located in Dartford Road, Thornleigh.
According to Hornsby Shire Council report (2004), Dartford Road landfill was originally
used as an extractive industry involving the quarrying of clay and shale in 1901. In 1979,
the extractive materials were basically empty, followed by a decision which was made by
the Metropolitan Waste Disposal Authority to use the abandoned pits for the disposal of
putrescible waste. The waste disposal facility officially was launched in June, 1980 as a
part of a long term plan that would enable the site to be developed for playing fields when
the pits were filled to a satisfactory level. The pit was used as a regional storage for
domestic waste until 1986, and the site contains approximately 1.1 million cubic metres of
waste. Additionally, it is noted that 500,000 tonnes of waste were deposited at the Dartford
road landfill site between 1980 and 1986. Finally, the landfill site is covered with between
0.25m and 2.1m of silty clay with occasional gravel horizons.
83
Figure 3.3 Dartford road landfill - golf driving range under construction
Figure 3.4 demonstrates Brenan Park, which is located in Smithfield suburb. According to
Fairfield city council report (2000), Brenan Park was used as a garbage disposal depot until
1980. At the end of this time a heavy mantle of impervious clay was then spread over the
surface. The land was declared as unhealthy building land in 1981 and developed into a
reserve. The park includes six soccer fields and tennis courts and some informal open
spaces. It has built up the surface levels above surrounding landforms.
The decomposed municipal solid waste samples used in this study, were collected from the
former Bankstown landfill, which is a closed landfill site in south-west of Sydney. It was
closed 30 years ago. Two test pits were excavated using a back hoe. Test pit 1 (TP1) was
4.6 m deep and test pit 2 (TP2) reached 4 m. The locations of the landfill, a plan view of the
former Bankstown landfill and the locations of test pits are shown in Figures 3.5 to 3.7,
respectively. In addition, Table 3.3 presents the depth and characterisation of the test pits.
Figure 3.5 Location of the former Bankstown landfill (courtesy of Google Maps)
85
Figure 3.6 Plan view of the former Bankstown landfill and test pits locations
86
3.7 In-situ MSW Characterisation in Bankstown Former Landfill
Essential information that could help characterise the waste materials in the landfill was
collected. Site characterisation was mainly based on the collection of information relevant
to the operation of the landfill. The main in-situ activities included excavating and bulk
sampling of waste materials. Dynamic Cone Penetrometer (DCP) test, which measures a
material’s in-situ resistance to penetration, was conducted by repeatedly dropping a metal
cone with a specific weight, into the waste material from a specific height. The penetration
of the cone is measured and recorded after each blow to continuously measure the shearing
resistance up to the required distance below the surface. Due to the existence of large
materials such as bricks, the results were inconsistent and therefore cannot be reported. The
materials were described visually during the process of excavation. Information regarding
the degradation and age of the waste materials asked from Bankstown council staff and
then confirmed from the dates of several newspapers, magazines, and milk cartons found
within the waste mass.
x Composition: Visual description of the material that observed at each depth, the
existence of soil material and the general description of the waste materials (e.g.
paper, plastic, wood, brick pieces and household garbage) were useful. This
information gave a general idea of the waste materials placed in the landfill, and
along with information collected by the landfill operators could provide a general
description of the materials.
x Age: Attempts were made to estimate the age of the waste. Magazines, newspapers
or other documents provided information on the age of the waste.
x Samples: The boring logs also indicated the locations at which samples were
collected
x Moisture content level: At each depth, the moisture level of the waste visually
characterised.
x Degradation: Although the waste degradation rate was an important feature, it was
difficult to quantify the state of MSW biodegradation.
87
3.8 Primary Geotechnical MSW Characterisation
The aim of this stage of characterisation is to provide some additional data on the samples
collected on the site. The materials should be divided between those that were larger or
smaller than 19 mm. According to Zekkos (2005), the basis of this division is that all soil
particles pass through a 19 mm sieve while particles greater than 19 mm in MSW are not
generally considered to be decomposed waste materials. The smaller than 19 mm fraction
can also be characterised with typical geotechnical characterisation practices, such as
moisture content, organic content, sieve analyses, and so on. The materials less than 19
mm, which are composed of fine gravel and smaller size particles, can be tested in the
laboratory with conventional testing equipment. The weights of the materials larger or less
than 19 mm should be measured and the relative volumes of the two fractions should be
visually estimated. The material loss tests were carried out to estimate the moisture content
and the organic content. Sieve analysis was performed on the less than 19 mm materials.
This data facilitated in selecting an appropriate sampling procedure which was then used in
the experimental program of this study.
Figure 3.8 shows the process of excavation using a backhoe-loader. After completing
excavation, the materials were visually characterised and a representative sample was
collected and placed into large drums, inside of tight plastic bags, for testing in the
laboratory. Figure 3.9 gives a side view of the excavated test pit 2. Figures 3.10 and 3.11
also show a view of the primary waste components. 12 drums were filled with waste
materials from different depths. The material was then sealed, and the date, test pits number
and depth of collection were marked on the drums. In addition, a smaller sample was
gathered from each drum for prompt determination of the moisture content. Figures 3.12 to
3.15 show the placement process of representative samples of waste in the drums. The
samples were transported to the soil laboratory at University of Technology Sydney and
stored for primary geotechnical MSW characterisation.
88
Figure 3.8 Digging the test pit using a backhoe
Figure 3.13 Filling drums with excavated waste materials using a backhoe
90
Figure 3.14 Placement of representative samples of waste in the drums
Figure 3.15 Placement of the soil fraction in plastic bags and in the drum
At this stage of the waste characterisation, the larger than 19 mm fraction was divided into
the various different waste components. According to general observation from all the
sample groups characterised, the predominant components were:
• Papers, cardboards, cartons
• Plastic bags and plastic fragments
• Wood and timber fragments
• Gravel (including ceramics and brick chips)
Other components, such as textiles, construction debris, metals, glass etc. typically
represented a small percentage of the overall mass of the waste and for that reason, were
typically not used in subsequent laboratory investigations.
91
3.11 Moisture Content and Organic Content of MSW – Material Loss
3.11.1 Background
Siegel et al. (1990) presented field moisture content for the OII landfill ranging from 10%
to 45% by drying small samples at 60ºC. Gabr and Valero (1995) dried material at 60ºC but
included only the smaller fraction of the waste materials. They showed a moisture content
profile increased with depth from about 30% near the surface to over 130% at high depths.
In addition, the organic content was estimated to be about 33%. Sanchez-Alciturri et al.
(1993) reported a moisture content of 48% for Meruello landfill in Spain, Gomes et al.
(2002) noted a moisture content ranging between 61-96% for recent waste near the surface
and about 117% for 2 to 3 year old waste. A sample collected during a rainy season had a
moisture content of about 150%. They also reported that the organic content, which is
estimated by heating the material at 450ºC varies between 43% and 63%.
3.11.2 Definitions
Material Loss Fraction (MLF) is determined as the ratio of the weight of material that is
lost during heating at a specific temperature (Wl) divided by the weight of the dry material
that remained (Wd). It is defined by the equation (3.1):
Wl
MLF
Wd (3.1)
The particle sizes (e.g. less than 19 mm) need to be stated as well as the composition of the
material and the heating temperature. The moisture content and organic content are defined
as follows:
x Moisture content is defined as the ratio of the weight of the less than 19 mm material
loss to the weight of the material remained during heating at a temperature of 55 degrees
Celsius.
x Organic content is defined as the ratio of the weight of the less than 19 mm material loss
to the weight of the material remained during heating at a temperature of 440 degrees
Celsius according to the ASTM D2974-87 (1995) procedures.
92
3.11.3 Material Loss Testing Procedure
A representative material was prepared and separated into the smaller and larger than 19
mm fraction, before heating using a proper sieve. The initial mass of each fraction was also
recorded. The two fractions were placed in the oven in separate pans and were heated at a
temperature of 55qC until the waste was dried and reached to a constant mass. The dry
mass of each fraction was recorded. The Material Loss Fraction at this stage was
determined as moisture content. In the next step the oven was heated to a temperature of
105qC until the waste was dried and reached to a constant mass (The sample weighted
minimum twice and interval between readings was at least 6 hours). Once again the dry
mass of each fraction was recorded. Two representative samples of the less than 19 mm
materials were placed in a muffle furnace, as presented in Figure 3.16. They were heated at
440qC, until the material was reached to a constant mass. Figure 3.17 displays a view of the
furnace used for the estimation of the organic content of the waste material. The Material
Loss Fraction of the less than 19 mm, after the completion of this stage was determined as
the organic content of the material. Figure 3.18 shows the samples of less than 19 mm
material in the furnace.
93
Figure 3.17 View of furnace used for estimation of organic content of waste material
3.11.4 Results of Moisture Content and Organic Content for this Research
The field moisture contents (Material Loss Fraction at 55qC for the less than 19 mm
material) for samples from test pit 1 and test pit 2 were between 19.8% and 23.3%. The
organic contents of samples from test pit 1 and test pit 2 (Material Loss Fraction at 440qC
for the less than 19 mm material) was determined by executing tests on specimens that
weighed about 50 g each. The measured organic contents of samples were between 20%
and 28%. Table 3.4 summarises the results of the moisture and organic contents of the
collected waste samples.
94
Table 3.4 Moisture content and organic content results of collected waste samples
The smaller than 19 mm fraction was composed of 50%-60% by the mass of the sample
composition. In addition to the smaller than 19 mm fraction, papers, soft plastics, wood
chips composed a significant amount of the material by mass and volume. Metals and stiff
plastics consisted of a small percentage by mass and volume.
Dry sieve analyses of the smaller than 19 mm fraction and smaller than 9 mm fraction for
the both test pits were performed. Figures 3.19 and 3.20 display waste materials included
95
all particle sizes before sieving process and a view of the larger than 19 mm waste materials
(retained on the sieves), respectively.
Figure 3.19 Waste material included all particle size before sieving
Figure 3.20 View of the larger than 19 mm waste material (retained on the sieves)
Smaller than 19 mm fraction: The decomposed MSW samples were sieved carefully and
particles smaller than 19 mm were used for a number of tests. Figure 3.21 shows the
particle size distribution of materials finer than 19 mm fraction. Based on the size of the
particles, the samples taken from the collected MSW, with a maximum size of 19 mm, were
classified as silty sandy gravel (GM) according to the Unified Soil Classification System.
Sieve analysis shows that 42% of the particles were in the range of gravel size, 38% in the
range of sand size and 20% in the range of fine grained soils. Figures 3.22 and 3.23 present
the sieve analysis process and the processing of the waste through the 19 mm sieve,
respectively.
96
100
90
80
Percentage passing 70
60
50
40
30
20
10
0
0.001 0.010 0.100 1.000 10.000
Particle size (mm)
Figure 3.21 Dry sieve analyses of finer than 19 mm fraction
100.0
90.0
80.0
Percentage passing
70.0
60.0
50.0
40.0
30.0
20.0
10.0
0.0
0.001 0.010 0.100 1.000
Particle size (mm)
Figure 3.24 Dry sieve analyses of finer than 9 mm fraction
98
Figure 3.26 Small wood particles on one of the sieves
3.14 Summary
Many councils in New South Wales, Australia were contacted to get permission to visit
their landfill sites, and also request for taking samples from these sites. Excavation in a
landfill site is considered dangerous so most requests to take samples from the landfill sites
were rejected by the councils, mainly due to safety issues. Blackman Park (Lane Cove
Council), Wangal Park (Burwood Council), Dartford Road Landfill (Hornsby Shire
Council) and Brenan Park (Fairfield City Council) among several landfill sites, were visited
for sampling purposes. Finally, after numerous contacts, conducting risk assessments and
completing the required safety paperwork, the decomposed municipal solid waste samples
used in this study, were collected from the former Bankstown landfill site, which is a closed
landfill site in south-west of Sydney. This site had been closed 30 years ago. The main in-
situ activities were included excavating and bulk sampling of waste materials. Dynamic
Cone Penetrometer (DCP) tests were also performed. Due to the presence of construction
rubbles and large size particles such as brick pieces the results were not consistent.
Accordingly the results have not been reported in this chapter. Information regarding the
composition, degradation and the age of the wastes were gathered. These collected data
facilitated in selecting an appropriate sampling procedure, employed in the experimental
program of this study. After excavation was completed, materials were characterised
visually and a representative sample of the material was collected. 12 drums (70 Liter per
99
one) were filled with waste materials from different depths. The materials were divided into
two groups: (a) the materials with particle size of smaller than 19 mm and (b) materials
larger than 19 mm. Particles greater than 19 mm in MSW are generally not considered as
decomposed waste materials. Those materials smaller than 19 mm were examined by
typical geotechnical characterisation practices such as the moisture content, the organic
content and the sieve analysis. The material loss tests were carried out to estimate the
moisture content and the organic content. Sieve analysis was performed on the less than 19
mm materials. The field moisture contents for samples were between 19.8% and 23.3%.
The measured organic contents of samples were between 20% and 28%. Based on the size
of the particles, the samples taken from the collected MSW, with a maximum size of 19
mm and 9 mm, classified as silty sandy gravel (GM) and silty sand (SM), respectively
according to the Unified Soil Classification System.
100
CHAPTER 4
4 Materials, Sample Preparation and Laboratory
Testing Program
4.1 Introduction
This chapter describes details of the materials, sample preparation, laboratory tests and
procedures and experimental program employed in the current research. In addition, the
testing set ups and apparatuses, used for this research at the University of Technology
Sydney are described.
4.2 Materials
As explained in Chapter 3, the decomposed municipal solid waste samples used in this
study, were collected from the former Bankstown landfill site in south-west of Sydney,
which was closed 30 years ago (1983). There were some difficulties in recovering and
testing the representative waste samples due to the large size of some waste components.
101
Depending on the size of the prepared specimen, there were limitations on the maximum
advisable particle size (as explained in Section 4.4). Therefore, the decomposed MSW
samples were sieved carefully and only those particles which were smaller than 19 mm or 9
mm were used for tests. The decomposed waste materials used for this research are soil like
material that includes mostly soil, degraded organic waste such as cardboards, wood and
food. Details of sieve analyses and characterisation of material is explained in Section 3.12.
ASTM C 618 (2012) defines two types of fly ashes which are classified as either Class C or
Class F, depending on its chemical constitution. Class C ashes are mostly come from sub-
bituminous coals and include mainly calcium alumino-sulfate glass, quartz, tri-calcium
aluminate, and free lime (CaO). Class C ash is categorised as high calcium fly ash, having
cementitious characteristics due to containing more than 20 percent free lime. It has the
ability to harden and gain strength in the presence of water. Class F ashes are obtained from
bituminous coals composing of alumino-silicate glass, with quartz, mullite, and magnetite.
ASTM D 5239 (2004) classifies fly ashes into three categories under their soil stabilisation
performances:
1. Non Self-Cementing (Class F) Fly Ash Stabilisation: by itself, has little effect on soil
stabilisation. It is a poor source of calcium and magnesium ions. The particle size of fly ash
may exceed that of the voids in fine-grained soils, excluding its use as a filler material.
However, in poorly graded sandy soils it can be an appropriate filler material helping in
compaction, increasing density and decreasing permeability.
102
2. Non Self-Cementing (Class F) Fly Ash Mixed with Cement or Lime: Some fine-grained
soils are pozzolanic and only need lime or cement to begin the pozzolanic reaction. The use
of Class F fly ash mixed with cement or lime in some clay improves pozzolanic properties
and soil texture.
3. Self-Cementing (Class C) Fly Ash Stabilisation: It is a rich source of calcium and
magnesium ions. Self cementing properties originate from amounts of free lime that can
provide cation exchange and ion crowding to fine grained soils when used in significant
amounts. This type of fly ash has been practiced to control swell potential of expansive
soils and also to stabilise coarse-grained soils.
Some efforts have already been made to achieve innovative approaches to take advantage
of fly ash in construction. Toth et al. (1988) reported that fly ash has been demonstrated
applicable as fill materials for highway embankments, road bases and grout mixes. Fly ash
mostly contains pozzolans, which are siliceous or aluminum siliceous substances, which are
not cementitious materials; however react with slaked lime to form a mixture with proper
cementitious characteristics. In Australia, fly ash only in class F is abundant. Therefore, a
mixture of fly ash and quicklime with an optimum ratio are used as the main chemical
stabiliser. Figure 4.1 shows Eraring fly ash used as an additive to waste material for
stabilisation.
103
4.2.2.1 Chemical Composition of Fly Ash
Table 4.1 provides a general chemical composition of Eraring fly ash, which has been used
for the current project. Eraring fly ash is a natural pozzolan. It is a fine cream/grey powder,
which is low in lime content. By itself it possesses little or no cementitious properties.
However, in its finely divided form and in the presence of moisture, it will react chemically
with calcium hydroxide (e.g. from lime or cement hydration) at ambient temperatures to
form insoluble compounds possessing cementitious properties.
Material Al2O3 CaO Fe2O3 K 2O MgO MnO Na2O P2O5 SiO2 TiO2
Percent 25.50 2.27 3.92 1.24 0.69 0.08 0.52 0.33 64.20 0.97
4.2.3 Lime
Lime (calcium oxide) is usually made by the thermal decomposition of materials such as
limestone, that contain calcium carbonate (CaCO3; mineral calcite) in a lime kiln. This is
performed by heating the material to above 825°C. A process called calcination or lime-
burning, to liberate a molecule of carbon dioxide (CO2); leaving quicklime. The quicklime
is not stable and, when cooled, will spontaneously react with CO2 from the air, and within a
proper time span, it will completely be converted back to calcium carbonate.
Lime mixing can be practiced to modify some of soil physical properties and improve the
ground into a stabilised mass, which increases its strength and durability. The amount of
lime additive depends on the type of the soil to be stabilised. Lime stabilisation is applied in
road construction to improve the subbase and the sub-grade, in railroads and airports
construction, for embankments, for soil exchange in unstable slopes, for backfill, for bridge
abutments and retaining walls, for canal linings, for improvement of soil beneath
foundation slabs, and for lime piles (Anon 1985 and 1990).
The improvement of the geotechnical properties of soil and the chemical stabilisation
process using lime occur through short term and long term chemical reactions. Short term
104
reactions include cation exchange and flocculation. Lime is a strong alkaline base, which
reacts chemically with clays causing a base exchange. Calcium ions displace sodium,
potassium, and hydrogen cations and change the electrical charge density around the clay
particles. This causes an increase in the interparticle attraction causing flocculation and
aggregation with a subsequent decrease in the plasticity of the soils. On the other hand,
long-term reaction including pozzolanic reaction, where calcium from the lime reacts with
the soluble alumina and silica from the clay in the presence of water to produce stable
calcium silicate hydrates (CSH), calcium aluminate hydrates (CAH), and calcium
aluminosilicate hydrates (CASH), which generate long-term strength gain and improve the
geotechnical properties of the soil. These hydrates were observed by many researchers
(Diamond et al. 1964; Sloane 1965; Ormsby and Kinter, 1973; and Choquette et al. 1987).
The use of lime for soil stabilisation is either in the form of quicklime (CaO) or hydrated
lime Ca(OH)2. Hydrated lime is used in most of the lime stabilisation applications.
Quicklime represents approximately 10% of the lime used in lime stabilisation process. The
Addition of the hydrated lime Ca(OH)2, in situ or in laboratory, is either as slurry formed
by the slaking of quicklime, or as dry form. In general, all lime treated fine-grained soils
exhibit decreased plasticity, improved workability and reduced volume change
characteristics. However, not all soils exhibit improved strength characteristics. In the
present study quicklime was used. Figure 4.2 demonstrates view of quicklime used as an
additive to waste material.
105
4.2.3.1 Activation of Fly ash with Lime
Fly ash mainly consists of oxides of silicon, aluminum, iron and calcium. Based on ASTM
C618 fly ash is classified as either Class C or Class F, depending on its chemical
constituents. Class C ash is categorised as high calcium fly ash, having cementitious
characteristics due to containing more than 20 percent free lime. Quicklime, CaO, is
produced by calcination of limestone, CaCO3, at 1100°C. High amount of CO2 (nearly
40%) are emitted during the calcination process. After quicklime is exposed to water it
reacts and forms hydrated lime, Ca(OH)2.
The reaction of fly ash with lime results in an immediate improvement of the soil’s
mechanical properties during stabilsation. It reduces the moisture content of the soil, which
has a fast stabilising effect, increases soil pH, preparing a condition for secondary
pozzolanic reactions and also it produces heat, which accelerates the chemical reactions
(Janz and Johansson 2002). Many activation methods have been suggested by researchers
to treat unstable soil with fly ash. The main purpose of these efforts was to improve the
reactivity of the pozzolan, in order to enhance the mechanical and stability characteristics
of the mixed product. Extended grinding proposed by Bouzoubaa et al. (1997), curing at
high temperatures suggested by Shi and Day (1993), alkali activation considered by Palomo
et al. (1999) and Xie and Xi (2001) and chemical activation procedures presented by Shi
and Day (1995) are some approaches, which have been practiced to attain that goal.
Comprehensive investigations have been conducted in the past regarding fly ash activation
by chemical methods, specifically with utilisation of lime. Huang and Cheng (1986)
examined the fly ash–hydrated lime combination to quantify the optimum mixture of
hydrated lime for accelerating the reaction of fly ash at different ages. They found that the
acceleration benefits would be insignificant, unless curing occurred at higher temperatures.
Gray and Lin (1972) presented that the strength gain ratio of lime-treated fly ash are
extremely dependent on the curing temperature. Aimin and Sarkar (1991) mixed gypsum
with cement and Class F fly ash and found a clear increase in the strength. Shi (1999)
compared the results of hydrated lime as well as quicklime mixed with pozzolan/cement.
He reported that the cement mixed with quicklime reveals considerably higher strength than
the cement mixed with hydrated lime. Brooks et al. (2011) showed that fly ash and lime
106
addition to clay soils decrease plasticity characteristics, swell potential and enhance
compressive strength of the soils. Phanikumar and Sharma (2004) showed that by mixing
fly ash with expansive soil the optimum moisture content of mixes decreases and the
maximum dry density increases.
The maximum dry density and the optimum water content of the untreated MSW samples
with particles smaller than 9 mm and 19 mm were measured by using a 2.7 kg hammer and
25 blows per layer (AS1289.5.1.1-2003). Figure 4.3 shows compaction test equipments
used. The treated MSW samples were then prepared by mixing MSW materials and fly ash-
quicklime in various ranges shown in Table 4.2. Once again the optimum water content and
the maximum dry density of the treated MSW samples were determined. Figure 4.4
demonstrates filling process of the compaction mould with smaller than 9 mm decomposed
waste material. Figure 4.5 presents the compacted waste material and the mould after
compaction process has been completed. Because the MSW samples in landfills are not as
compacted as layers of soil, a revised compaction test with 15 blows per layer was carried
out and for sample preparation this method was used. Tables 4.2 and 4.3 list the compaction
test results for untreated and treated MSW samples with various fly ash–quicklime
contents. Each test was repeated three times and the average data were reported.
Figure 4.5 Compacted waste material with mould after compaction completed
Table 4.2 Compaction test results for untreated and treated MSW samples (particles
smaller than 9 mm)
108
Table 4.3 Compaction test results for untreated and treated MSW samples (particles
smaller than 19 mm.)
There are some limitations on the maximum particle size, which can be used for specimen
preparation. Typically the largest particle size should be 1/6 of the specimen diameter. For
the triaxial specimens (diameter of 50 mm), material sieved and particles smaller than 9
mm used for specimen preparation. Because the size of some material such as wood fibres
was typically larger in one direction, the material selection needed some visual judgment.
This process ensured that for the waste particles, the maximum particle size was 1/6 of the
specimen diameter. In addition, due to this limitation the maximum particle size for
compaction tests, the direct shear tests and the permeability tests was 19 mm and the
sieving process was performed by passing the material through the 19 mm sieve whereas
for triaxial tests including consolidation and permeability in the triaxial cell was 9 mm.
Figure 4.6 displays waste material included all particle sizes before sieving.
Figure 4.6 Waste material included all particle size before sieving
109
4.5 Ratio of Fly Ash and Quicklime in Soil Stabilisation
Several investigations have been conducted to find out the optimum ratio of fly ash to
quicklime for soil stabilisation. Conner (1990) suggested that a typical free lime-to-fly ash
ratio for soil stabilisation should be between 0.15 and 0.6, and higher ratios do not
generally contribute to strength. In addition, Fang (1990) noted that lime to fly ash ratios of
1:2 to 1:7 have been tested and accepted. However ratios of 1:3 and 1:4 are used typically
for economy and quality purposes. Typical proportions of lime and fly ash range from 2.5
to 4 percent for lime and 10 to 15 percent for fly ash. Furthermore, Wesche (1991)
mentioned that the optimum lime/fly ash ratio is 1:4 and the best compressive strength is
reached when the mixture is compacted with humidity slightly below the optimum moisture
content.
Two groups of MSW samples, which were smaller than 19 mm (Figure 4.7) and smaller
than 9 mm materials, were prepared by mixing quicklime and fly ash with a ratio of 1:3 in
the range shown in Table 4.4. Following preparation, the specimens were accurately mixed
thoroughly. A mechanical mixer, shown in Figure 4.8, was used for mixing the
decomposed waste materials with fly ash and quicklime. Due to high amount of moisture
content of MSW samples, which was on the wet side of the optimum moisture content, a
dry mixing method (no water added to MSW samples during mixing process) was adopted.
Figure 4.9 presents a sample of waste materials mixed with fly ash and quicklime. Figure
4.10 shows the moisture content absorption ratio of the treated MSW samples using
different amounts of fly ash-quicklime mixture. Meanwhile, Figure 4.11 presents the
moisture content of treated MSW samples with various amounts of fly ash-quicklime.
110
Figure 4.7 Smaller than 19 mm waste material before mixing with additives
Figure 4.8 Mixture procedure of waste material with fly ash and quicklime
Figure 4.9 View of sample of waste material after mixing with fly ash and quicklime
111
0.4
0.35
Moisture content
absorbtion ratio
0.3
0.25
0.2
0.15
0.1
0.05
0
6.7 13.3 20.0 26.7
24
Moisture content of mixed
20
16
MSW sample
12
8
4
0
0.0 6.7 13.3 20.0 26.7
According to the fact that MSW samples in landfills are not compacted as much as the
normal layers of soil, a revised compaction method with 15 blows per layer was applied for
sample preparation. Based on revised compaction results all the treated and untreated
samples in this experimental program were compacted in 3 layers with their natural
moisture contents. As the moisture contents of the waste material after treatment was
reduced to values close to the optimum moisture content of materials, the dry density of
compacted material was close the maximum dry density shown in Tables 4.2 and 4.3. After
the compaction, each sample was securely wrapped to prevent moisture loss and cured in a
112
constant temperature-humidity chamber (at 25°C ± 2°C and at relative humidity ≥98%) for
28 days. Some of the quicklime/fly ash mixtures were cured for 7 and 93 days to estimate
the influence of curing time factor on the geotechnical properties and on the quicklime/fly
ash-stabilisation process. The quicklime/fly ash-soil mixtures remained securely wrapped
until testing procedures were to be conducted. It should be noted that each test was repeated
three times and the average data were reported in the thesis. After specimens were
prepared, the conventional geotechnical experiments, including compaction tests, uniaxial
compressive tests, permeability tests, triaxial tests and consolidation tests were carried out.
These experimental tests were performed according to the procedures expressed by the
Australian Standard, AS 1289 (2003).
Sample group Fly ash content Quicklime Total addatives Curing period
(%) content (%) (%) (days)
A 5 1.7 6.7 7, 28, 93
B 10 3.3 13.3 7, 28, 93
C 15 5 20 7, 28, 93
D 20 6.7 26.7 7, 28, 93
Unconfined compressive strength tests were carried out in accordance with AS1289, using
an equipment complying with the requirements of AS1289, Grade B. These requirements
included the upper block of the machine having a spherical seat. The spherical seat
facilitated to ensure the load was evenly distributed in cases that the capping had not
entirely corrected the roughness on the top surface of the specimen. The compressive
testing machine was set at a rate of 1 mm per minute for all specimens. The dimensions of
the tested specimens were 200 mm in height and 100 mm in diameter. The treated and
untreated waste materials with particles smaller than 19 mm were compacted in the mould
in 3 layers. Table 4.5 demonstrates a summary of the unconfined compressive strength
specimens and the tests performed. Mixture procedure and compaction method were
113
explained in Sections 4.5 and 4.6. Figure 4.12 shows a view of the split mould used for the
unconfined compressive strength test sample preparation. It should be noted that the
moulds were cleaned and lightly oiled prior to commencing the mixing process. Figures
4.13 and 4.14 show two snapshots of the prepared samples for unconfined compressive
strength tests with various fly ash – quicklime contents. Each test was performed in a
strain-controlled mode. The compressive testing equipment can be seen in Figure 4.15. A
20 kN ‘S’ type load cell was used for the testing. For strain measurement, an LVDT with a
range of 10 mm was used for the testing of the treated and untreated waste samples. A data
tracker was used to convert the electronic signals received from the load cell, the LVDT
and the strain gauges to data output on a nearby computer. The unconfined compressive
strength (UCS) of each specimen was measured from the testing process, and the stress-
strain graphs were plotted from the data recorded. The results of the compressive testing are
provided in Chapter 5 of this dissertation. Generally, the testing procedure for each
specimen was terminated once the load had dropped by at least 30% from the maximum
recorded value (i.e. the UCS). However, to fully observe the behaviour of the specimen
during failure, the test was occasionally carried out until full collapse occurred. Each test
produced load versus displacement data until specimen failure was observed. The raw data
were then converted into stress versus strain plots, with unconfined compression strength
(undrained shear strength) and the strain at failure. Figure 4.15 presents a specimen with
20% additives before unconfined compressive strength test. In addition, Figures 4.16 and
4.17 show the specimens with 13% and 26% additives after completion of the unconfined
compressive strength test, respectively.
Figure 4.12 The split mould used for the unconfined compressive strength test sample
preparation
114
Table 4.5 Summary of unconfined compressive strength specimens and tests performed
Figure 4.13 View of prepared samples for unconfined compressive strength test with
various fly ash – quicklime content
115
Figure 4.14 View of specimen with 13% additives (10% fly ash + 3.3% quicklime)
Figure 4.15 A specimen with 20% additives before unconfined compressive strength test
116
Figure 4.16 A specimen with 13% additives after completion of unconfined compressive
strength test
Figure 4.17 A specimen with 26% additives after completion of unconfined compressive
strength test
The direct shear test is performed in accordance with AS1289.6.2.2 (1998) for the
estimation of shear strength parameters such as the cohesion and the friction angle. The
direct shear test has several shortcomings such as the non-uniform stress distribution along
117
the failure plane and the existence of a forced horizontal failure surface. However, the
direct shear test is a conventional test for the estimation of MSW shear strength parameters.
Direct shear tests usually seems to yield lower estimates of the shear strength than the
triaxial tests. Because the failure surface is necessarily horizontal, specimens prepared with
different particle orientations with respect to the shearing surface can provide significant
insight regarding the mechanics of waste response and the anisotropic nature of strength in
MSW.
A Large Shear Box (300 mm × 300 mm) was used for performing the direct shear tests at
the soils laboratory at the University of Technology Sydney (UTS), as shown in Figure
4.18. The device consists of several transducers to measure horizontal and vertical
displacements, and a gear box for specifying and controlling the rate of the horizontal
shearing (mm/min) and the direction of movement (forward vs. backward). In addition, this
large shear box has a special system for applying substantial vertical loads. The motor
applies the horizontal displacement to the upper part of the shear box at a constant rate. The
power supply for the motors is controlled by a switch located on the equipment. The
specimens tested in this experimental program were sheared at a displacement rate of 1 mm
per minute. The shear box, in which the specimen is placed, consists of an upper and a
lower part. The lower part remains fixed and the upper part is connected to a 20 kN load
cell that measures the shear resistance between the two boxes. The shear box is surrounded
by a steel tank which (if desired) can be filled with water. Initially, the upper part of the box
sits on lower part using a pair of locating pins. Following the specimen’s preparation, the
pins are removed. The height of specimens in this large shear box is 200 mm. A steel
loading plate with slightly smaller dimensions than the size of the shear box (to ensure that
there will be no contact between the loading plate and the walls of the upper part of the
shear box) is placed at the top surface of the specimen. The vertical load is applied by dead
loads (Figure 4.19). The shear displacement of the specimen is measured by an LVDT
placed against the wall of the steel tank. Another LVDT is placed at the top of the loading
system and measures the vertical deformation during the compression/consolidation phase
as well as during shearing as shown in Figure 4.20.
118
Figure 4.18 View of the direct shear device used for the performance of the tests
Figure 4.19 The crane used for lifting the normal force and applying to the sample
Figure 4.20 3 LVDTs for the measurement of the horizontal and vertical displacements
119
4.8.2.2 Large Direct Shear Testing Program
Many treated and untreated specimens were prepared and tested in the large direct shear
box as part of the laboratory investigation program. Specimens included those that were
100% smaller than 19 mm materials only. A compaction hammer with a square base plate
was manufactured from steel (Figure 4.21). The materials were mixed and compacted in the
shear box with the hammer, in 3 equal layers, until they reached the desired dry density.
The direct shear testing program is summarised in Table 4.6. As expected, the area of the
shearing specimen was reduced during the direct shear test, with the result that, although
the initially applied vertical load remained constant during shearing, the normal stress
increased throughout the shearing process as the load was distributed over a smaller
shearing area. During the shearing tests, the cap of the specimen progressively tilted, but
during testing the cap was restrained from touching the sides of the shear box. In fact the
tests were stopped earlier to ensure the cap made no contact. This phenomenon was a result
of the shearing process and became more pronounced as the horizontal displacement
increased.
Figure 4.21 The hammer used for the specimen compaction in shear box at the University
of Technology Sydney
120
Table 4.6 Summary of large direct shear specimens and tests performed at the University of
Technology Sydney
Specimen no Normal Stress Fly ash – Dry unit weight Curing time
(kPa) quicklime upon preparation (Days)
content (kN/m3)
DS1-1 50 0-0 11.3 28
DS1-2 100 0-0 11.3 28
DS1-3 200 0-0 11.3 28
DS3-1 50 10 - 3.3 11.6 28
DS3-2 100 10 - 3.3 11.6 28
DS3-3 200 10 - 3.3 11.6 28
DS4-1 50 15 - 5 11.7 28
DS4-2 100 15 - 5 11.7 28
DS4-3 200 15 - 5 11.7 28
DS5-1 50 20 - 6.7 11.8 28
DS5-2 100 20 - 6.7 11.8 28
DS5-3 200 20 - 6.7 11.8 28
DS7-1 50 15 - 5 11.7 93
DS7-2 100 15 - 5 11.7 93
DS7-3 200 15 - 5 11.7 93
122
4.8.3.2 Constant Head Hydraulic Conductivity Tests
Constant head hydraulic conductivity tests were conducted following the procedures in
laboratory. Due to the lack of standard test procedures for determining the permeability
coefficient of municipal solid waste, the standard test procedure for determining the
permeability of the granular soils (AS1289.6.7.1-2001) was adopted for the current study.
The test setup consisted of a cylinder with an inside diameter of 105 mm and a height of
115.5 mm, porous stones, a stand, clamps, silicon grease and a tank maintaining a constant
hydraulic head. The untreated and treated waste material (prepared as explained in mixing
of material and compaction section) with particles smaller than 19 mm was compacted
inside the cylinder in three uniformly distributed layers to achieve the required uniform
density. Table 4.8 presents the summary of permeability test specimens and their details. In
addition, Figure 4.22 shows a view of compacted waste material in mould for permeability
test. The required densities were obtained by changing the compacting of sample by
maintaining the constant height of the sample. The weight of the sample is determined
based on the volume of the sample and required density. Figures 4.23 to 4.25, demonstrate
the use of a high permeable fibre to prevent the sample particles movement into drainage
layers, drainage layer on top and bottom of the sample and a view of the prepared sample
for permeability test, respectively. The setup is connected to the water line making sure that
there are no air bubbles along the water line. The head is measured 0.5 m and the flow in
the sample was allowed to stabilise until no air bubbles appeared in the outlet pipe. When
the flow rate was constant the permeability, k, was determined from the total amount of
flow, Q, in a given time, t, as equation (4-1)
QL
k
hAt (4-1)
where h is the hydraulic head, A denotes the cross sectional area of the specimen, and L is
the height of the sample. Figure 4.26 displays an experimental setup for permeability test.
123
Table 4.8 Summary of permeability test specimens and their details
Figure 4.22 A compacted waste material in a mould prepared for the permeability test
124
Figure 4.23 Using high permeable fibre to prevent sample particles movement into
drainage layers
The triaxial tests were performed in accordance with AS1289.6.4.2 (1998). The triaxial
compression tests were used to measure the shear strength of MSW samples under
controlled drainage conditions. The strain rate adopted was 0.1 mm/min, and the samples
were tested at three different confinements (i.e. 100, 200, and 300 kPa), the confinements
were applied until the samples were completely consolidated, and then the loading was
applied. The triaxial system has the capability of measurements of load and displacement
using load cells and LVDTs with varying sensitivities. Triaxial consolidated drained (CD)
tests were performed using the compacted specimens at their natural moisture contents after
mixing process with fly ash and quicklime.
126
50 mm diameter and 100 mm height treated and untreated triaxial and consolidation
specimens were prepared in 3 layers applying the tamping technique to reach to the desired
dry density. Figure 4.27 illustrates a view of the triaxial sample preparation mould and the
extruder, used for this project. In addition, Tables 4.9 and 4.10 present the summary of the
triaxial specimens and the consolidation specimen and the tests performed, respectively.
After preparation and curing period, the specimen was placed in the device and back
pressure was applied to saturate the sample. Each specimen was fully enveloped with a thin
rubber membrane. Both ends of the membrane were sealed using rubber O-rings. Test cell
assembled by placing the plexiglass cylinder cell wall and the top cap assembly. Flexible
tubings coming from the panel to the base assembly ports attached. The space between the
specimen and the cell wall filled with the de-aired water by applying a positive pressure to
the water in the water tank. This was done to remove air out of the specimen during the
specimen saturation stage. Saturation process started after applying the back pressure to the
top and bottom of the specimen. The saturation of the sample was checked by the common
B-test, suggested by Skempton (1954). The specimen subjected to the back pressure for a
period of time until a B-value of 0.95 or larger was reached. This procedure was performed
by monitoring the pore water pressure reading frequently. B-value was calculated by
dividing the change in the pore water pressure reading by the change in the cell pressure.
Following the saturation of the specimen, the consolidation process commenced. The
confining pressure increased so that the difference between the confining pressure and back
pressure matched the desired effective consolidation stress. Consolidation test performed
under different effective stresses (i.e. 50, 100, 150, 200 and 300 kPa). The pore water
pressure readings were recorded at specified times. These data used to verify the
completion of the consolidation process and to determine the loading rate for the triaxial
test. After completion of primary consolidation when the excess pore pressure reached to
zero or less than 1 kPa, for the sample with the effective stress of 300 kPa the test
continued to investigate the effect of creep with different fly ash-quicklime contents. On the
other hand, after consolidating the soil specimen under three different confinements (i.e.
100, 200, and 300 kPa), deviator loading was applied. The loading piston moved in contact
with the top platen. Using computer software the loading rate set to the specified value and
the loading process started. Figure 4.28 shows a photograph of the triaxial apparatus during
127
a CD test on a municipal solid waste sample. During the test, the computer automatically
recorded all of the sensor readings frequently. The actual test duration depended on the
loading rate and the behaviour of the soil specimen. The test was terminated at an
additional strain beyond the deviatoric stress peak. Figure 4.29 presents a snapshot of a
failed waste sample after completion of CD Triaxial test. Shortly after the triaxial test, the
water from the test cell was drained. The cell dismantled and the soil specimen removed.
The final moisture content of the soil was determined by placing the entire specimen in the
laboratory oven. It should be noted that an appropriate axial strain rate was 0.5 mm/min to
induce minimum excess pore pressure during shearing.
128
Table 4.10 Summary of consolidation specimens and tests performed
129
Figure 4.27 The triaxial sample preparation mould and the extruder
Figure 4.28 The automated triaxial apparatus during a CD test on a MSW sample
130
Figure 4.29 A snapshot of a failed waste sample after completion of a CD Triaxial test
4.9 Summary
The materials, sample preparation, laboratory tests and procedures and the details of the
experimental program, used in this research, have been explained in this chapter. Eraring
fly ash and quicklime were used as additives to waste materials for stabilisation. Eraring fly
ash is a natural pozzolan. By itself it possesses little or no cementitious properties.
However, in its finely divided form and in the presence of moisture, it will react chemically
with calcium hydroxide. The improvement of the geotechnical properties of soil and the
chemical stabilisation process using lime occur through short term and long term chemical
reactions. The reaction of fly ash with lime results in an immediate improvement of the
soil’s mechanical properties during stabilsation. It reduces the moisture content of the soil,
which has a fast stabilising effect, increases soil pH, preparing a condition for secondary
pozzolanic reactions and also it produces heat, which accelerates chemical reactions. The
maximum dry density and the optimum water content of the untreated and treated MSW
samples with particles smaller than 9 mm and 19 mm were measured. According to the fact
that MSW samples in landfills are less compacted than normal soil layers, a revised
compaction test with 15 blows per layer was carried out and all the treated and untreated
131
samples in this experimental program compacted in 3 layers with their natural moisture
contents. Following preparation, the specimens were accurately mixed thoroughly. A
mechanical mixer was used for mixing the decomposed waste materials with fly ash and
quicklime. Due to the high amount of moisture content of MSW samples, which was on the
wet side of the optimum moisture content, the dry mixing method (i.e. no water added to
MSW samples during mixing process) was adopted. As the moisture contents of the waste
material after treatment reduced to values close to the optimum moisture content of
materials, the dry density of compacted material was close to the maximum dry density. By
mixing fly ash-quicklime with MSW samples, the maximum dry density increased and the
optimum moisture content decreased. The direct shear test was performed using a large
shear box (300 mm × 300 mm) to estimate the shear strength parameters of treated and
untreated MSW samples. Description of the large direct shear box, testing program, effect
of particles orientation have been presented in this chapter. In addition, permeability test
procedures for recompacted samples have been reported. The hydraulic conductivity
(permeability coefficient) of recompacted samples was measured in a triaxial cell while the
sample was consolidating. Moreover, the specimen preparation for unconfined compressive
strength, triaxial and consolidation tests and the associated testing procedures have been
explained in detail.
132
CHAPTER 5
5 Experimental Results and Discussion
5.1 Introduction
This chapter presents the results of a geotechnical laboratory program used to estimate how
the use of fly ash and quicklime could improve the geotechnical properties of municipal
solid waste (MSW). An array of experimental tests were conducted on treated and untreated
MSW samples, including sieve analysis, Atterberg limits, compaction, unconfined
compressive strength, direct shear, permeability and consolidated-drained triaxial tests. As
explained in Chapter 3, the waste materials were collected from former landfill at
Bankstown in the southwest of Sydney. The samples were prepared by integrating MSW
with a mixture of fly ash-quicklime at a ratio of 3:1 in percentages of 5, 10, 15, and 20 of
fly ash by dry weight of the MSW.
The unconfined compressive tests were carried out on MSW specimens mixed with
different contents of fly ash-quicklime. After curing for 7, 28, and 93 days unconfined
compressive strength tests were performed. Table 5.1 summarises the unconfined
compressive strength (UCS) values for specimens with different contents of fly ash-
quicklime. It should be noted that the results were based on an average of three identical
133
tests. The overall behaviour of MSW was significantly influenced by the addition of fly
ash-quicklime. Figure 5.1 shows the axial strain versus compression strength of the MSW
specimens treated with various percentages of fly ash-quicklime. Based on this figure,
mixing fly ash-quicklime with the MSW specimens increased the unconfined compressive
stress (UCS) of the specimens quite significantly; indeed the unconfined compressive
strength of MSW specimens increased 150% from 102 kPa for an untreated specimen to
256 kPa for a 7-day cured specimen treated with 20% fly ash-quicklime. Figure 5.1 shows
that the optimum content of fly ash-quicklime increased the compressive strength of MSW
specimen by around 20% of MSW’s dry weight, but the compressive strength of the MSW
was reduced when the content of fly ash-quicklime was more than 20% (15% fly ash plus
5% quicklime). Furthermore the elasticity modulus at the first loading (Esecant), failure axial
strain (εf), and brittleness changed as a consequence of adding fly ash-quicklime. The
addition of fly ash-quicklime led to an increment of the Esecant and a decrement of the
failure axial strain (εf) for the specimens.
300
treated with 15% fly ash-5% quicklime
treated with 20% fly ash-6.7% quicklime
Unconfined Compressive Stress (kPa)
150
100
50
0
0 0.01 0.02 0.03 0.04 0.05 0.06
Axial Strain (%)
Figure 5.1 Unconfined compressive stress-strain of untreated and treated MSW specimen
with different fly ash-quicklime contents
134
Table 5.1 Results of unconfined compressive strength tests
The results of the UCS tests conducted for this study were in the same pattern presented by
Tastan et al. (2011). They reported the results of the tests performed with three organic
soils and six fly ashes. Fly ashes were blended with soils at three different percentages and
different water contents, including the optimum water content (OWC) and at the wet side of
the OWC. Referring to Figure 5.2, they reported that the unconfined compressive strength
of organic soils increased when using fly ash, but the amount of increase depended on the
type of soil and characteristics of the fly ash. Furthermore, the resilient moduli of organic
soils were significantly improved. The authors attributed the increases in the strength and
stiffness to cementing caused by pozzolanic reactions, although the reduction in the water
content caused by the addition of dry fly ash also contributed to the gain in strength.
Previous studies generally showed that the pozzolonic effect reduces as the water content
135
decreases. The significant characteristics of fly ash that affect the increase in the
unconfined compressive strength and the resilient modulus include the content of CaO and
the ratio of CaO/SiO2.
Figure 5.2 Engineering properties of organic soil–fly ash mixtures as a function of fly ash
percentage in the mixture (Tastan et al. 2011)
MSW specimens mixed with different contents of fly ash-quicklime were prepared and
cured for periods of 7, 28, and 93 days to estimate the influence of curing time on the
unconfined compressive strength of the treated specimens. The unconfined compressive
strength of specimens treated with fly ash-quicklime increased significantly with curing
time compared to the untreated MSW specimens. Figure 5.3 indicates the effect of curing
time on the unconfined compressive strength of MSW specimens mixed with various
contents of fly ash-quicklime. According to this figure the increase in curing time from 7
days to 93 days, enhances the unconfined compressive strength of the treated specimens.
136
For example, the unconfined compressive strength of the MSW specimens mixed with 20%
fly ash-quicklime increased by 91% (i.e. increased from 256 kPa to 490 kPa).
700
Untreated specimen
Mixed with 5% fly ash+ 1.7% quicklime
Unconfined Compressive Strength (kPa)
400
300
200
100
0
7 28 93
Curing Time (Days)
Figure 5.3 Effect of curing time on unconfined compressive strength of MSW specimens
The results reported by Ghosh and Subbarao (2007) showed an increase in the unconfined
compressive strength of specimens when curing time increased from 7 days to 90 days.
They displayed the shear strength characteristics of class F fly ash modified with lime or
lime combined with gypsum. A number of unconfined compression tests were performed
for unsoaked and soaked specimens that had been cured for up to 90 days. They reported
that the addition of a small percentage of gypsum (0.5 and 1.0%) along with lime (4–10%)
improved the shear strength of modified fly ash within short curing periods (7 and 28 days).
The authors observed that the addition of lime increased the shear strength of the stabilised
mixes due to an increase in the availability of lime for a pozzolanic reaction, as shown in
Figure 5.4. The rate of gain in shear strength was high with the higher content of lime.
They noted that fly ash mix stabilised with 10% lime attained 5,901 kPa of unsoaked
unconfined compressive strength (qu) after 90 days of curing time, whereas the values were
172, 1,200, and 3,130 kPa for mixes stabilised with 0, 4, and 6% of lime, respectively.
Moreover, as Figure 5.5 shows, Ghosh and Subbarao (2007) concluded that the
contribution of gypsum at early stages of curing was significant at curing periods up to 45
137
days and increased when the content of gypsum increased from 0.5 to 1.0%. However, at a
higher curing period of 90 days the contribution of gypsum was relatively less. The
addition of gypsum to lime stabilised fly ash decreased the reduction of the shear strength
due to soaking. They noted that the addition of gypsum increased the bond strength
between the particles by accelerating the formation of pozzolanic reaction products.
Figure 5.4 Unconfined compressive strength of fly ash versus curing period for unsoaked
specimens with varying percentages of lime (Ghosh and Subbarao 2007)
Figure 5.5 Unconfined compressive strength of fly ash versus curing period for soaked and
unsoaked specimens with varying percentages of lime and (a) 0.5%; (b) 1.0% gypsum
(Ghosh and Subbarao 2007)
138
5.3 Direct Shear Test
A series of direct shear tests were carried out on untreated and treated MSW specimens
with different contents of fly ash-quicklime. These tests were conducted at three different
normal stresses of 55 kPa, 110 kPa, and 220 kPa. A detailed description of the large direct
shear apparatus at the University of Technology Sydney and the method of preparing the
specimens have been given in Chapter 4. A summary of the results of the large direct shear
tests are presented in Table 5.2, and show that by increasing the content of fly ash-
quicklime the peak shear stress of treated specimens reached a higher value. Figure 5.6
shows the shear strength envelope from the results of the direct shear tests on untreated and
treated MSW specimens mixed with different contents of fly ash-quicklime. It reveals that
the optimum content of fly ash-quicklime needed to increase the shear stress of MSW
specimen is estimated to be 20% of the MSW’s dry weight. There was a slight reduction in
the shear stress of the MSW specimen when the content of fly ash-quicklime was more than
20%. These results revealed that the cohesion and friction angle values for untreated MSW
specimens were estimated as 17 kPa and 31º, respectively. Figure 5.6 also illustrates that by
mixing 20% fly ash-quicklime to the MSW specimen, the friction angle and cohesion
values increased from 31º to 39º and from 17 kPa to 33 kPa, respectively.
MSW specimens mixed with different amounts of fly ash-quicklime were prepared and
cured for 28 days and 93 days to estimate how the curing time affects the shear strength
parameters of the treated specimens. The shear strength parameters of specimens treated
with fly ash-quicklime increased slightly with a longer curing time compared to shear
strength parameters of untreated MSW specimens. Figure 5.7 indicates the effect of curing
time on the shear strength parameters of MSW specimens mixed with 20% fly ash-
quicklime. According to these results, by increasing the curing time of specimens treated
with 20% fly ash-quicklime from 28 days to 93 days, the friction angle increased from 39º
to 40º and the cohesion value increased from 33 kPa to 40 kPa.
139
300
untreated specimens
specimens mixed with 10% fly ash - 3.3% quicklime
250 specimens mixed with 20% fly ash- 6.7% quicklime
specimens mixed with 15% fly ash - 5% quicklime
Shear Stress (kPa)
Φ = 39º , C= 44 kPa
200
Φ = 38º , C= 32 kPa
Φ = 36º , C= 22 kPa
150
Φ = 31º , C= 16 kPa
100
50
0
0 50 100 150 200 250 300
Normal Stress (kPa)
Figure 5.6 Shear strength envelope from the results of direct shear test on untreated and
treated MSW specimens with different fly ash-quicklime contents
300
treated with 15% fly ash + 5% quicklime
250
Φ = 40º , C= 40 kPa
Shear Stress (kPa)
200
150
Φ = 39º , C= 34 kPa
100
Figure 5.7 Effect of curing time on shear strength envelope of treated MSW specimens with
20% fly ash-quicklime
140
Table 5.2 Summary of large direct shear tests results
Specimen Fly ash – Dry unit Curing Normal Peak shear c Friction
no quicklime weight time stress stress (kPa) angle, φ
content (%) (kN/m3) (days) (kPa) (kPa) (º)
DS1-1 0-0 11.3 28 50 47
DS1-2 0-0 11.3 28 100 85 17 31
DS1-3 0-0 11.3 28 200 146
DS3-1 10 - 3.3 11.6 28 50 61
DS3-2 10 - 3.3 11.6 28 100 103 22 36
DS3-3 10 - 3.3 11.6 28 200 181
DS4-1 15 - 5 11.7 28 50 72
DS4-2 15 - 5 11.7 28 100 135 34 39
DS4-3 15 - 5 11.7 28 200 211
DS5-1 20 - 6.7 11.8 28 50 71
DS5-2 20 - 6.7 11.8 28 100 127 33 38
DS5-3 20 - 6.7 11.8 28 200 204
DS7-1 15 - 5 11.7 93 50 79
DS7-2 15 - 5 11.7 93 100 143 40 40
DS7-3 15 - 5 11.7 93 200 220
Figures 5.8 to 5.15 demonstrate the shear stress versus the axial strain and vertical
displacement (all specimen showed dilation) versus horizontal displacement of untreated
and treated MSW specimen according to the results of direct shear tests. It was observed
that an increase in the normal stress intensified the peak shear stress. Moreover, the
maximum vertical displacement for the specimen tested under a higher normal stress had a
lower value. For example, for the MSW specimens mixed with 15% fly ash-quicklime, the
maximum vertical displacement were 4.7 mm and 2.6 mm under 55 kPa and 220 kPa
normal stresses, respectively. Furthermore, it can be observed that the peak vertical
displacement had a higher value for treated MSW specimen with the higher amount of fly
ash-quicklime.
141
160
Normal stress
140 220 kPa
Shear Stress (kPa) 120 110 kPa
55 kPa
100
80
60
40
20
0
0 0.02 0.04 0.06 0.08 0.1 0.12
Axial Strain (%)
Figure 5.8 Shear strength of untreated MSW specimen in direct shear test under different
normal stresses
3
Normal stress
Vertical Displacement (mm)
2.5 55 kPa
110 kPa
2 220 kPa
1.5
0.5
0
0 5 10 15 20 25 30 35
Horizontal Displacement (mm)
Figure 5.9 Vertical displacement vs. horizontal displacement for untreated MSW specimen
under different normal stresses
142
200
100
Normal stress
50
220 kPa
110 kPa
55 kPa
0
0 0.02 0.04 0.06 0.08 0.1
Axial Strain (%)
Figure 5.10 Shear strength of treated MSW specimen mixed with 13.3% fly ash-quicklime
in direct shear test under different normal stresses
3.0
Vertical Displacement (mm)
2.0
Normal stress
1.0
55 kPa
110 kPa
220 kPa
0.0
0 5 10 15 20 25 30
Horizontal Displacement (mm)
Figure 5.11 Vertical displacement vs. horizontal displacement for treated MSW specimen
mixed with 13.3% fly ash-quicklime under different normal stresses
143
250
Normal stress
220 kPa
200 110 kPa
Shear Stress (kPa) 55 kPa
150
100
50
0
0 0.02 0.04 0.06 0.08
Axial Strain (%)
Figure 5.12 Shear strength of treated MSW specimen mixed with 20% fly ash-quicklime in
direct shear test under different normal stresses
4.0
Normal stress
Vertical Displacement (mm)
55 kPa
3.0 110 kPa
220 kPa
2.0
1.0
0.0
0 5 10 15 20
Horizontal Displacement (mm)
Figure 5.13 Vertical displacement vs. horizontal displacement for treated MSW specimen
mixed with 20% fly ash-quicklime under different normal stresses
144
250
Normal stress
200 220 kPa
Shear Stress (kPa) 110 kPa
55 kPa
150
100
50
0
0 0.01 0.02 0.03 0.04 0.05
Axial Strain (%)
Figure 5.14 Shear strength of treated MSW specimen mixed with 26.7% fly ash-quicklime
in direct shear test under different normal stresses
5.0
Normal stress
Vertical Displacement (mm)
4.0 55 kPa
110 kPa
220 kPa
3.0
2.0
1.0
0.0
0 5 10 15
Horizontal Displacement (mm)
Figure 5.15 Vertical displacement vs. horizontal displacement for treated MSW specimen
mixed with 26.7% fly ash-quicklime under different normal stresses
145
Landva and Clark (1990) conducted several direct shear tests using a large scale direct
shear device. The results reported by them for old refuse were in good agreement with the
results of the direct shear tests conducted on the MSW specimens in this study. These
results are presented in Figure 5.16. The interpreted cohesion was between zero and 23 kPa,
while the friction angle ranged between 24 and 41 degrees.
Figure 5.16 Large-scale DS test results on MSW from Canada (Landva and Clark, 1990).
146
Bray et al. (2009) integrated the results of 100 large-scale laboratory direct shear tests from
various investigations. As presented in Figure 5.17, the shear strength is defined by a
friction angle of 36o at a normal stress of 100 kPa, with the friction angle decreasing by 5o
for every log cycle increase in the normal stress (i.e., using Equations (2.13) and (2.14): c =
15 kPa, M 0 = 36o , and 'M = 5o).
Figure 5.17 Recommended static shear strength of MSW based primarily on direct shear
tests and field observations of static slope stability (Bray et al. 2009).
Furthermore, the results of the in situ direct shear test conducted on the MSW sample by
Caicedo et al. (2002a) confirmed the results obtained from the laboratory tests in this
research. They executed in situ large scale direct shear tests with a cross sectional area of
0.63 m2 on MSW samples to evaluate the causes of a landslide at the Dona Juana landfill in
Columbia. MSW specimens having a diameter of 900mm were tested in situ using direct
shear apparatus specifically designed for that purpose. The results of the in situ direct shear
tests are presented in Figure 5.18. The shear strength of the MSW was characterised by c =
78 kPa and φ=23 degrees.
147
Figure 5.18 Results of in situ direct shear tests on MSW (Caicedo et al. 2002a).
One of the aims of this study was an investigation of the hydraulic conductivity of MSW
stabilised with fly ash-quicklime because previous studies provide limited data. A number
of constant head permeability tests were conducted on treated and untreated MSW samples
(as explained in Section 4.8.3.2). Due to particle size limitations, as explained in Section
4.4, the materials were sieved and particles smaller than 19 mm were used for the tests. The
permeability tests indicated that the coefficient of permeability for the fly ash is between
10-6 and 10-7 m/s. Figure 5.19 presents the results of the coefficient of permeability of
MSW specimens with and without fly ash-quicklime treatment, as well as the effect of
curing time. The results show that increasing the rate of fly ash-quicklime in the MSW
specimens reduced the flow rate and consequently the coefficient of permeability of the
MSW specimens decreased after this treatment. As shown in Figure 5.19, by mixing 6.7%
and 13.3% fly ash-quicklime with MSW samples, the coefficient of permeability of the
specimens reduced rapidly from 1.6u10-5m/s to 7u10-6m/s and 2.4u10-6m/s, respectively,
but increasing the content of fly ash-quicklime from 13.3% to 26.7% produced no
significant changes in the coefficient of permeability. This reduction in the permeability
coefficient can be attributed to the following reasons:
148
x The conversion of soluble calcium hydroxide to cementitious compounds decreases
the bleed channels and void spaces and thereby reduces the permeability coefficient.
x The small finely divided particles of fly ash increased the density by acting like
micro-aggregates and facilitated to fill the voids in the MSW samples.
The k-value of MSW specimens stabilised with fly ash-quicklime was evaluated after 7, 28,
and 93 days of curing. As can be seen in Figure 5.19, no significant changes were observed
in the coefficient of permeability with time.
1.7E-05
After 7 days
1.5E-05
Coefficient of Permeability ( m/s)
After 28 days
1.3E-05 After 93 days
1.1E-05
9.0E-06
7.0E-06
5.0E-06
3.0E-06
1.0E-06
0.0 6.7 13.3 20.0 26.7
Fly ash-quicklime Content (%)
Figure 5.19 Coefficient of permeability of untreated and treated MSW specimens under 7 ,
28 and 93 days curing time
In this study the reduction in the coefficient of permeability of MSW treated with fly ash-
quicklime agreed well with the measured results reported by Phanikumar and Sharma
(2004) for samples of expansive soil mixed with fly ash. They presented a study of the
efficacy of fly ash as an additive in enhancing the engineering characteristics of expansive
soils. An experimental program was conducted to evaluate the effect of the content of fly
ash on the compaction, strength, and hydraulic conductivity of expansive soils. The
149
hydraulic conductivity of the blends decreased and the dry unit weight and strength
increased with an increase in the content of fly ash. The estimated values of hydraulic
conductivity k, for blends at various contents of fly ash are presented in Table 5.3. The tests
were performed on blends at their corresponding optimum moisture contents and dry unit
weights. The results indicated that the value of k decreased with an increase in the content
of fly ash. They noted that this decrease in hydraulic conductivity correlated well with the
increase in the maximum dry unit weight with an increase in content of fly ash.
Table 5.3 Effect of fly ash on compaction Behavior, and hydraulic conductivity (Modified
after Kumar and Sharma 2004)
Fly ash content (%)
Property 0 5 10 15 20
Optimum moisture content (%) 40 38 35 33 31
Max dry unit weight (kN/m3) 13.75 13.91 14.05 14.19 14.30
Hydraulic conductivity, k (m/s) 9.70ൈ10-9 Not tested 6.02×10-9 Not tested 3.95×10-9
Consolidation testing in a triaxial cell was used to determine several important soil
parameters, one of those is the coefficient of consolidation cv which governs the rate of
consolidation under a change in applied pressure. Wray (1986) described two graphical
procedures for determining the cv. The equation for the log-of-time method can be
expressed as:
0.196 d 2
cv (5-1)
t 50
where d is the initial thickness of the specimen before a change in the applied pressure. The
values of t50 reflect the time required for a 50% reduction in the thickness of the specimen
150
to occur (or 50% reduction in sample volume), and the cv values were obtained for each
consolidation pressure applied. Figure 5.20 shows the variation in the coefficient of
consolidation with different contents of fly ash-quicklime under different effective
confining pressures. These results confirmed increasing the content of fly ash-quicklime in
the specimen reduced the coefficient of consolidation. For example, under 200 kPa of
average effective confining pressure, the cv values were 1.6×10-5 m2/s and 1.2×10-5m2/s for
the untreated specimens and the specimens mixed with 26% fly ash-quicklime,
respectively. In addition, with an increase in the confining pressure up to the
preconsolidation pressure, the cv value reduced and changes was insignificant for higher
7.0E-05
Coefficient of Consolidation, cv (m 2/s)
4.0E-05
3.0E-05
2.0E-05
1.0E-05
0.0E+00
0 6.7 13.4 20 26.7
Fly ash + quick lime Content (%)
Figure 5.20 Coefficient of consolidation of MSW specimens for different fly ash-quicklime
contents under various effective confining pressures
k cv .mv .J w (5-2)
where J w is the unit weight of water, cv is the coefficient of consolidation and mv is the
coefficient of volume compressibility, defined as the volume change per unit volume per
unit increase in the effective stress. The value, given by the above equation, represents the
vertical component of the hydraulic conductivity tensor.
Table 5.4 summarises the results of the permeability tests through consolidation of
specimens in triaxial apparatus, while Figure 5.21 shows the coefficient of permeability of
MSW specimens stabilised with different contents of fly ash-quicklime. These results
show that by increasing the content of fly ash-quicklime in the MSW specimen, the
coefficient of permeability was reduced. The coefficient of permeability for an untreated
specimen was 6.2×10-8 m/s and was reduced to 3.2×10-8 m/s for specimens mixed with
26% fly ash-quicklime (under average confining pressure of 250 kPa). This parameter is
4.7×10-8 m/s, 4.1×10-8 m/s and 3.5×10-8 m/s for MSW specimens mixed with 6.7%, 13.3%
and 20% fly ash-quicklime respectively. Furthermore, Figure 5.22 shows the effect of
effective confining pressure on the permeability of untreated and treated MSW specimens
where increasing the effective confining pressure up to the pre-consolidation pressure
caused no significant change in the coefficient of permeability. However at higher
confining pressures the coefficient of permeability was reduced. As explained earlier, this
reduction can also be attributed to the conversion of soluble calcium hydroxide to
cementitious compounds, a reduction in the bleed channels and void spaces, and an
increase in the density.
152
Table 5.4 Summary of permeability test results through consolidation of specimen in
triaxial cell
Fly ash- Effective Average effective
CV Permeability Average
quicklime Stage confining confining
(m2/s) (m/s) void ratio
content (%) pressure (kPa) pressure (kPa)
1 20-50 35 6.3E-05 1.4E-07 0.62
2 50-100 75 4.9E-05 1.4E-07 0.60
0-0 3 100-150 125 1.8E-05 1.1E-07 0.56
4 150-200 175 1.6E-05 8.4E-08 0.52
5 200-300 250 1.6E-05 6.2E-08 0.47
1 20-50 35 5.4E-05 1.1E-07 0.59
2 50-100 75 4.1E-05 1.1E-07 0.58
5-1.7 3 100-150 125 1.6E-05 8.5E-08 0.55
4 150-200 175 1.4E-05 6.7E-08 0.51
5 200-300 250 1.4E-05 4.7E-08 0.46
1 20-50 35 5.2E-05 1.0E-07 0.58
2 50-100 75 3.9E-05 1.0E-07 0.56
10-3.3 3 100-150 125 1.4E-05 7.1E-08 0.53
4 150-200 175 1.3E-05 5.5E-08 0.49
5 200-300 250 1.2E-05 4.1E-08 0.45
1 20-50 35 4.7E-05 9.0E-08 0.56
2 50-100 75 3.6E-05 9.0E-08 0.54
15-5 3 100-150 125 1.3E-05 5.8E-08 0.52
4 150-200 175 1.2E-05 4.9E-08 0.49
5 200-300 250 1.2E-05 3.5E-08 0.45
1 20-50 35 4.5E-05 8.5E-08 0.54
2 50-100 75 3.5E-05 8.4E-08 0.53
20-6.7 3 100-150 125 1.3E-05 4.9E-08 0.51
4 150-200 175 1.2E-05 4.4E-08 0.48
5 200-300 250 1.2E-05 3.2E-08 0.44
153
1.6E-07
Average confining pressure
Coefficient of permeability, k (m/s) 1.4E-07 35 kPa 75 kPa
1.2E-07 125 kPa 175 kPa
250 kPa
1.0E-07
8.0E-08
6.0E-08
4.0E-08
2.0E-08
0.0E+00
0 6.7 13.4 20 26.7
Fly ash + quick lime Content (%)
Figure 5.21 Coefficient of permeability of MSW specimens for different fly ash-quicklime
contents under various effective confining pressures
1.6E-07
untreated specimen
5% fly ash+ 1.7% quicklime
Coefficient of permeability, k (m/s)
1.4E-07
10% fly ash + 3.3% quicklime
15% fly ash + 5% quicklime
1.2E-07 20% fly ash + 6.7% quicklime
1.0E-07
8.0E-08
6.0E-08
4.0E-08
2.0E-08
0 50 100 150 200 250 300
Average Confining Pressure (kPa)
Figure 5.22 Coefficient of permeability of MSW specimens for different fly ash-quicklime
contents under various effective confining pressures
154
5.4.3.3 Permeability Change Index
The value of the permeability coefficient depends on the size, shape and void distributions
of particles. Taylor (1948) proposed an empirical linear relationship between log k and e.
The relationship can be expressed as:
e0 e
log k log k 0 (5-3)
ck
where ck is the permeability change index, and k 0 and e0 are the in-situ values of
permeability and the void ratio, respectively. This type of relationship has been commonly
used to describe the variation of permeability with the void ratio.
Figures 5.23 to 5.28 show the relationship between the void ratio-permeability (e-log k) for
different contents of fly ash-quicklime. Based on linear regression analysis, the relationship
between the void ratio-permeability and ck values has been shown for different contents of
fly ash-quicklime. The results show that increasing the content of fly ash-quicklime
reduced the ck value. For example, by mixing 26% of fly ash-quicklime, the ck value
reduced from 0.38 (for untreated specimen) to 0.21, while the ck values were 0.34, 0.28 and
0.25 for specimens with 20%, 13.6% and 6.7% of fly ash-quicklime, respectively.
0.650
Untreated sample
0.600
Void Ratio, e
0.550
0.400
1.0E-08 1.0E-07 1.0E-06
Coefficient of permeability, k (m/s)
155
0.650
5% fly ash - 1.7% quicklime
0.600
Void Ratio, e
0.550
log k = 2.97e - 8.61
0.500 Ck =0.34
k 0 = 1.1E-07
e0 = 0.593
0.450
0.400
1.0E-08 1.0E-07 1.0E-06
Coefficient of Permeability, k (m/s)
Figure 5.24 Void ratio-permeability relationship of treated MSW specimen with 6.7% fly
ash-quicklime
0.640
10% fly ash - 3.3% quicklime
0.600
0.560
Void Ratio, e
0.520
log k = 3.53e - 9.02
0.480 Ck =0.28
k 0 = 1.0E-07
0.440 e0 = 0.575
0.400
1.0E-08 1.0E-07 1.0E-06
Coefficient of Permeability, k (m/s)
Figure 5.25 Void ratio-permeability relationship of treated MSW specimen with 13.3% fly
ash-quicklime
156
0.600 15% fly ash - 5% quicklime
0.560
Void Ratio, e
0.520
0.400
1.0E-08 1.0E-07 1.0E-06
Coefficient of permeability, k (m/s)
Figure 5.26 Void ratio-permeability relationship of treated MSW specimen with 20% fly
ash-quicklime
0.600
20% fly ash - 6.7% quick lime
0.560
Void Ratio, e
0.520
0.400
1.0E-08 1.0E-07 1.0E-06
Coefficient of Permeability, k (m/s)
Figure 5.27 Void ratio-permeability relationship of treated MSW specimen with 26.7% fly
ash-quicklime
157
0.640
Untreated specimen
0.610 5% fly ash + 1.7% quicklime
10% fly ash + 3.3% quicklime
0.580 15% fly ash + 5% quicklime
20% fly ash + 6.7 quicklime
0.550
Void Ratio, e
0.520
0.490
0.460
0.430
0.400
1.0E-08 1.0E-07
Triaxial compression tests were used to measure the shear strength of MSW specimens
under controlled drainage conditions. Specimen preparation and testing procedures were
described in Chapter 4. As explained earlier, the specimens were tested at three different
effective confining pressures (i.e. 100, 200, and 300 kPa), the confinements were applied
until the specimens consolidated completely, and then the loading was applied. Table 5.5
shows the results of the peak and residual principal stress differences obtained from the
triaxial compression tests. The stress-strain-volumetric response of MSW specimens in the
triaxial compression tests are depicted in Figures 5.29 to 5.36 for untreated and treated
specimens with various percentages of fly ash-quicklime. Based on these figures,
increasing the content of fly ash-quicklime caused a significant change in the peak strength,
residual response, stiffness, and brittleness of the MSW specimens. For example, Figure
5.36 illustrates the effect of various contents of fly ash-quicklime on stress-strain behaviour
of specimens at an effective confining pressure of 300 kPa. It shows that the peak principal
158
stress of MSW specimens increased from 600 kPa (without treatment) to 1150 kPa (mixed
with 26.7% fly ash-quicklime). Comparing the curves indicates that by increasing the
content of fly ash-quicklime in MSW specimens the shear strength and stiffness increased
significantly, while the residual strength experienced a 30% increase by adding a 13.3%
admixture of fly ash-quicklime and remained relatively constant after extra fly ash-
quicklime was added. It was observed that by increasing the amount of additive, the
volumetric response was initially less compressive and then more expansive, while the post
peak behaviour became more brittle.
Table 5.5 Results of Peak and residual principal stress difference in triaxial compression
test
Specimen Effective Fly ash – Dry unit Curing Peak principal Residual
no confining quicklime weight time stress principal stress
pressure content (kN/m3) (days) difference difference
(kPa) (%) (kPa) (kPa)
TA1-1 100 0-0 11.3 28 224 200
TA1-2 200 0-0 11.3 28 413 367
TA1-3 300 0-0 11.3 28 600 532
TA2-1 100 5 – 1.7 11.5 28 314 257
TA2-2 200 5 – 1.7 11.5 28 555 476
TA2-3 300 5 – 1.7 11.5 28 794 695
TA3-1 100 10 - 3.3 11.6 28 388 303
TA3-2 200 10 - 3.3 11.6 28 675 545
TA3-3 300 10 - 3.3 11.6 28 958 784
TA4-1 100 15 - 5 11.7 28 440 305
TA4-2 200 15 - 5 11.7 28 761 547
TA4-3 300 15 - 5 11.7 28 1080 790
TA5-1 100 20 - 6.7 11.8 28 462 321
TA5-2 200 20 - 6.7 11.8 28 804 584
TA5-3 300 20 - 6.7 11.8 28 1142 844
159
700
Effective confining pressure
400
300
200
100
Untreated Specimen
0
0
Volumetric Strain (%)
-1
-2
-3
-4
0 5 10 15 20
Axial Strain (%)
Figure 5.29 Stress-strain-volumetric response of untreated MSW specimens
1000
Effective confining pressure
300 kPa
Principal Stress Difference (kPa)
400
200
-1
-2
0 5 10 15 20
Axial Strain (%)
Figure 5.30 Stress-strain-volumetric response of treated MSW specimens with 6.7% fly
ash-quicklime content
160
1200
Effective confining pressure
600
400
200
MSW treated with 10% fly ash + 3.3% quick lime
0
2
Volumetric Strain (%)
-1
-2
0 5 10 15
Axial Strain (%)
Figure 5.31 Stress-strain-volumetric response of treated MSW specimens with 13.3% fly
ash-quicklime content
1200
Effective confining pressure
300 kPa
Principal Stress Difference (kPa)
1000
200 kPa
800 100 kPa
600
400
200
MSW treated with 15% fly ash + 5% quick lime
0
2
Volumetric Strain (%)
-1
0 2 4 6 8 10
Axial Strain (%)
Figure 5.32 Stress-strain-volumetric response of treated MSW specimens with 20% fly ash-
quicklime content
161
1400
Effective confining pressure
600
400
200
MSW treated with 20% fly ash + 6.7% quick lime
0
2
Volumetric Strain (%)
-1
0 2 4 6 8
Axial Strain (%)
Figure 5.33 Stress-strain-volumetric response of treated MSW specimens with 26.7% fly
ash-quicklime content
500
Effective confining pressure 100 kPa
Principal Stress Difference (kPa)
400
300
200
20% fly ash + 6.7% quicklime
15% fly ash + 5% quicklime
100 10% fly ash +3.3% quicklime
5% fly ash + 1.7% quicklime
untreated specimen
0
0 5 10 15 20
Axial Strain (%)
Figure 5.34 Stress-strain responses of treated and untreated MSW specimens with different
percentages of fly ash-quicklime at effective confining pressure of 100 kPa
162
1000
600
400
20% fly ash + 6.7% quick lime
15% fly ash + 5% quick lime
200 10% fly ash + 3.3% quick lime
5% fly ash + 1.7% quick lime
untreated specimen
0
0 5 10 15 20
Axial Strain (%)
Figure 5.35 Stress-strain responses of treated and untreated MSW specimens with different
percentages of fly ash-quicklime at effective confining pressure of 200 kPa
1400
20% fly ash- 6.7% quicklime
15% fly ash- 5% quicklime
Principal Stress Difference (kPa)
800
600
400
0
10 15 20 25 0 5
Axial Strain (%)
Figure 5.36 Stress-strain responses of treated and untreated MSW specimens with different
percentages of fly ash-quicklime at effective confining pressure of 300 kPa
163
It should be noted that according to Figures 5.37 and 5.38, reported by Ghosh and Subbarao
(2007), the stress-strain response of stabilised fly ash was affected as a result of modifying
Class F fly ash with lime and gypsum. Those mixtures containing lime only and cured for 7
days showed stress–strain responses similar to mixes not stabilised with fly ash. The
specimens stabilised with gypsum and lime showed a sharp peak in the stress–strain curve,
and immediately after attaining the peak deviatoric stress there was a rapid reduction in
deviatoric stress and an increase in the strain for both 7-day and 28-day curing periods.
Figure 5.37 Stress–strain response of fly ash, 7 and 28 days curing (Ghosh and Subbarao
2007)
Figure 5.38 Stress–strain response of fly ash with 10% lime and 1% gypsum, 7 days curing
(Ghosh and Subbarao 2007)
The peak shear strength envelopes are presented in Figure 5.39. The data shown in Table
5.6, provide the results of the shear strength parameters, the modulus of elasticity at failure
164
stress (Ef) and at the 50% of failure stress (E50%) with and without treatments. It has been
observed that by increasing fly ash–quicklime contents from 0 to 26.7%, the peak friction
angle increased from 29q to 39q and the cohesion intercept increased from 11 kPa to 30
kPa. The residual strength envelopes are presented in Figure 5.40. The corresponding
parameters are also shown in Table 5.6. The results illustrate that the residual friction angle
is also increased from 27q for untreated MSW specimens to 34.5q for treated MSW
specimens with 26.7% fly ash-quicklime content, while the changes in the residual
cohesion values were not significant. It should be noted that the friction angle depends on
the effective stress under that the MSW specimens were tested.
Table 5.6 Peak and residual strength and elastic parameters for untreated and treated
MSW in triaxial compression test.
600
20% fly ash + 6.7% quicklime
15% fly ash + 5% quicklime
500 10% fly ash + 3.3% quicklime
5% fly ash + 1.7% quicklime
(kPa)
Untreated specimen
400
(σ'1-σ'3)/2
300
200
100
0
0 200 400 600 800 1000
(σ'1+σ'3)/2 (kPa)
300
200
100
0
0 200 400 600 800
(σ'1+σ'3)/2 (kPa)
Figure 5.40 Residual-strength envelopes of untreated and treated MSW specimens
Table 5.6 provides the results of the modulus of elasticity at failure stress (Ef) and at the
50% of failure stress (E50%) with and without treatments. Figures 5.41 and 5.42 provide an
additional analysis of the modulus of elasticity at failure stress (Ef) and at 50% of the
failure stress (E50%) for untreated and treated MSW specimens under different effective
confining pressures, respectively. The values of Ef and E50% were calculated using
Equations (5.4) and (5.5):
(V 1c V 3c ) max
Ef (5.4)
Hf
(V 1c V 3c ) max
E 50% (5.5)
2 u H 50%
where εf is the axial strain when the principal effective stress difference is equal to
ሺߪԢͳ െ ߪԢ͵ ሻ݉ܽ ݔand ε50% is the axial strain when the principal effective stress difference is
ሺߪԢ ͳ െߪԢ ͵ ሻ݉ܽݔ
equal to ʹ
. It has been found that Ef and E50% increased with an increase in the
content of fly ash–quicklime. It can be inferred from Figure 5.41 and Table 5.6 that by
166
increasing the content of fly ash–quicklime from 0 to 26.7%, both Ef and E50% increased
from 3.1-5.5 MPa to 14-27.9 MPa and from 6.1-12.1 MPa to 19.6-36.1 MPa, respectively.
Meanwhile, as shown in Figure 5.43, the ratio of E50% to Ef decreased.
40
Young's Modulus at 50% Failure Stress, E 50%
20
15
10
5
(b)
0
0.0 6.7 13.3 20.0 26.7
Fly ash + quicklime Content (%)
Figure 5.41 Variation of Young’s modulus at 50% failure stress for untreated and treated
MSW specimens under different effective confining pressures.
30
Young's Modulus at Failure Point, E f (MPa)
15
10
5
(a)
0
0.0 6.7 13.3 20.0 26.7
Figure 5.42 Variation of Young’s modulus at failure stress for untreated and treated MSW
specimens under different effective confining pressures.
167
2.2
Effective confining pressure
100 kPa
1.9
E50% /Ef (MPa)
200 kPa
300 kPa
1.6
1.3
(c)
1.0
0.0 6.7 13.3 20.0 26.7
Fly ash + quicklime Content(%)
Figure 5.43 Variation of stiffness ratio of untreated and treated MSW specimens under
different effective confining pressures.
The dramatic drop in strength after achieving peak values can be determined by the
brittleness index IB, based on Equation (5.6), and as defined by Bishop (1967):
q f qu
IB (5.6)
qu
where qf and qu are the peak deviator stress and the residual deviator stress, respectively.
Figure 5.44 shows the variation of the brittleness index for untreated and treated MSW
specimens with various percentages of fly ash-quicklime. The results indicated that the
brittleness index of MSW specimens treated by up to 20% of fly ash-quicklime increased
from 0.13 to 0.35 under an effective confining pressure of 300 kPa, while an addition of
20% to 26.7% of fly ash-quicklime showed that the brittleness index was relatively
constant. Moreover, the brittleness index of specimens under higher effective confining
pressure had a relatively lower value.
168
0.50
Effective confining pressure
100 kPa
0.40
200 kPa
Brittleness Index
300 kPa
0.30
0.20
0.10
0.00
0.0 6.7 13.3 20.0 26.7
Fly ash + quicklime Content (%)
Figure 5.44 Effect of fly ash-quicklime contents on brittleness index of MSW specimens
under different effective confining pressures
Figure 5.45 indicates the variation of the effective confining pressure with void ratio
(during primary compression) for the untreated and treated MSW specimens with different
contents of fly ash-quicklime, with the results showing that adding fly ash-quicklime
reduced the compressibility characteristics of the MSW specimens. Filling the voids with
particles of fly ash-quicklime reduced the void ratio. It can be observed that the
preconsolidation pressure of treated materials increased with increasing fly ash-quicklime
contents and it attributed to the cementation induced due to the chemical reaction.
169
0.64
σ'pc
0.60
0.56
0.52
Void Ratio
0.48
0.44
Untreated specimen
treated with 5% fly ash - 1.7% quicklime
0.40 treated with 10% fly ash - 3.3% quicklime
treated with 15% fly ash - 5% quicklime
treated with 20% fly ash - 6.7% quicklime
0.36
1 50 100 150 200 300
20
Effective Confining Pressure (kPa)
Figure 5.45 Primary compression of untreated and treated MSW specimens with different
fly ash-quicklime contents
Figure 5.46 shows the effect of 28 day and 93 day curing times on the primary compression
of the MSW specimens treated with 20% fly ash-quicklime. Increasing the curing time
from 28 days to 93 days increased the preconsolidation pressure slightly but there was no
change in the compression index.
170
0.58
28 days
0.56
93 days
0.54
0.52
Void Ratio
0.5
trea ted with 15% fly
0.48 a sh + 5% quicklime
0.46
0.44
0.42
0.4
1
20 50 100 150 200 300
Figure 5.47 and Table 5.7 show the values of the compressibility index C c, for the MSW
specimens treated with various contents of fly ash-quicklime, and show that the value of Cc
decreased from 0.33 for untreated MSW specimens to 0.23 for specimens treated with
26.7% fly ash-quicklime. This value for MSW specimen treated with 6.7%, 13.3% and 20%
fly ash-quicklime was 0.29, 0.28, and 0.24, respectively. As the content of fly ash-
quicklime increased the treated specimens would resist compressive loading better and
show less compressibility characteristics. It should be noted that no change in Cc values
were observed for specimens cured for 93 days.
Table 5.7 Variation of primary compression index for MSW specimen treated with different
fly ash-quicklime content and curing time
Fly ash
Curing time Compression
quicklime
(days) index, Cc
Content (%)
0-0 28 0.33
5-1.7 28 0.29
10-3.3 28 0.28
15-5 28 0.24
15-5 93 0.24
20-6.7 28 0.23
171
0.34
0.32
0.28
0.26
0.24
0.22
0.20
0.0 6.7 13.3 20.0 26.7
Figure 5.47 Effect of fly ash-quicklime contents on compression index of MSW specimens
A similar trend was presented by Okoro et al. (2011) for soil samples stabilised with fly
ash. They reported the consolidation characteristics of two soils stabilised with class C fly
ash (CFA) and lime. The first group of specimens was prepared with 10% class C fly ash
and 6% lime and then cured for 28 days; and the second group consisted of specimens
prepared with raw soils and compacted at near optimum moisture content and maximum
dry unit weight. The results shown in Figures 5.48 and 5.49 reveal that stabilisation with
CFA and lime reduced the compressibility of soil due to new cementing products in the
stabilised specimens. As shown in Table 5.8, the compression index (Cc) and swelling
index (Cs) decreased due to CFA and lime stabilisation. The percentage changes in C c and
Cs varied with the type of stabilising agents.
172
Figure 5.48 Void ratio versus pressure of raw and 6% stabilised Alloway Clay (Okoro et
al. 2011)
Figure 5.49 Void ratio versus pressure curves of raw and 10% CFA-stabilised Made Land
(Okoro et al. 2011)
Table 5.8 A summary of swell and compression indices of raw and stabilised soils. (after
Okoro et al. 2011)
173
Furthermore, the results shown by Phanikumar and Sharma (2007) confirmed a reduction in
the compression index of specimens where the content of fly ash had been increased. They
reported laboratory test data on the effect of the content of fly ash on the change in volume
of two expansive clays and a non-expansive clay. The variation in the compression index
and the secondary consolidation coefficient of the clays with fly ash was studied. They
found that the compression index of the expansive and non-expansive clays decreased by
about 50% after using 20% fly ash in the samples, as shown in Figure 5.50. As Figures 5.51
and 5.52 show the addition of fly ash reduced the compressibility characteristics of the
expansive and non-expansive clays, but the fly ash had a greater effect on the
compressibility behaviour of expansive clays than on non-expansive clays.
Expansive clay
Cc
Non-Expansive clay
174
Figure 5.51 e-log p curves of expansive clay specimens (Phanikumar and Sharma 2007)
Figure 5.52 e-log p curves of nonexpansive clay (Phanikumar and Sharma 2007)
175
5.5.6 Secondary Compression
The pore pressure readings were used to verify completion of the consolidation process,
and after primary consolidation was completed and excess pore pressure was almost zero,
the tests continued to investigate the effect of creep with different contents of fly ash-
quicklime. These tests were conducted on samples that were under an effective confining
pressure of 300 kPa.
The time required for primary consolidation to end (Tp) and secondary consolidation to
begin are presented in Figure 5.53, while the effective confining pressure was increased
from 200 kPa to 300 kPa.
This also showed how the untreated and treated MSW specimens behaved in secondary
consolidation with various contents of fly ash-quicklime under an effective confining
pressure of 300 kPa. Here the change in the void ratio of the MSW specimens mixed with
fly ash-quicklime at the end of primary consolidation due to creep and particles sliding was
less, while the volumetric strains of treated and untreated MSW specimens during primary
and secondary compressions are shown in Figure 5.54.
0.51
Effective confining pressure
increa sed from 200 kPa to 300 kPa
0.48
Tp = 22 min
0.45
Void Ratio
0.42
Tp = 30 min
0.39
0.36
Untreated specimen Tp=22 min
treated with 5% fly ash - 1.7% quicklime Tp=26 min
0.33 treated with 10% fly ash - 3.3% quicklime Tp=28 min
treated with 15% fly ash - 5% quicklime Tp=29 min
treated with 20% fly ash - 6.7% quicklime Tp=30 min
0.3
0.01 0.1 1 10 100 1000
Time (min)
Figure 5.53 Primary and secondary compression of MSW specimens when effective
confining pressure increased from 200 kPa to 300 kPa
176
Time (min)
0.1 1 10 100 1000
0
Sta ge 5
Effective stress increa sed
from 200 kPa to 300 kPa
2
Volumetric Strain (Δєv %)
Untreated specimen
10 treated with 5% fly asg + 1.7% quicklime
treated with 10% fly ash + 3.4% quicklime
treated with 15% fly ash + 5% quicklime
treated with 20% fly asg + 6.7% quicklime
12
Figure 5.54 Volumetric strain of treated and untreated MSW specimens during primary
and secondary compression (when effective confining pressure increased from 200 kPa to
300 kPa)
Figure 5.55 and Table 5.9 present variations of the secondary compression index with
various admixtures of fly ash-quicklime. Table 5.9 clearly shows that the value of the
secondary compression index Cα for MSW specimens, decreased from 0.052 to 0.033,
while the content of fly ash-quicklime increased from 0 to 26.7%. This means that for any
given time period, the change in volume during secondary compression was greater for
untreated specimens than those treated with fly ash-quicklime. Hence, fly ash-quicklime
effectively reduces the change in volume during primary consolidation and secondary
consolidation. Consequently, the amount of settlement of structures built on MSW treated
with fly ash-quicklime decreased. This reduction in the primary compression index and the
secondary compression index of MSW treated with fly ash-quicklime can also be attributed
177
to the effect of cementation and the pozzolanic activity of fly ash-quicklime. Finally, Figure
5.56 shows that the curing time did not affect the secondary compression index of MSW
specimens when the effective confining pressure increased from 200 kPa to 300 kPa.
Table 5.9 Variation of secondary compression index for MSW specimen treated with
different fly ash-quicklime content and curing time
0-0 28 0.052 22
5-1.7 28 0.045 26
10-3.3 28 0.040 28
15-5 28 0.036 29
15-5 93 0.036 29
20-6.7 28 0.033 30
* Cα and Tp values corresponding to 300 kPa effective confining pressure
0.055
Secondary Compression Index, Cα
Effective confining
0.050 pressure 300 kPa
0.045
0.040
0.035
0.030
0.0 6.7 13.3 20.0 26.7
Fly ash + quicklime Content (%)
Figure 5.55 Effect of fly ash-quicklime contents on the secondary compression index of
MSW specimens
178
0.51
28 days
0.49 93 days
Tp = 29min
0.47
0.45
Void Ratio
0.43
Effective stress increased
0.41 from 200 kPa to 300 kPa
Tp = 29min
0.39
0.37
0.35
0.01 0.1 1 10 100 1000
Time (min)
Figure 5.56 Effect of curing time on primary and secondary compression of MSW
specimens when effective confining pressure increased from 200 kPa to 300 kPa
The results shown by Phanikumar and Sharma (2007) confirmed this reduction in the
secondary compression index of specimens as the content of fly ash increased. They
showed in Figure 5.57 and Table 5.10 that the coefficient of secondary consolidation of
expansive and non-expansive clay decreased by 40% after it was stabilised with 20% fly
ash. The results indicated that the volume change due to creep and sliding of particles after
the end of primary consolidation was less by clay mixed with fly ash. In addition, the time
required for primary consolidation to end and secondary consolidation to begin was
shortened when clays were mixed with fly ash. It can therefore be concluded that the
settlement of expansive clays improved with fly ash decreased, and the rate of settlement
increased, thus reducing the time expected for final settlement to be completed.
179
Figure 5.57 Effect of fly ash on secondary consolidation (Phanikumar and Sharma 2007)
5.6 Summary
It has been shown that the overall behaviour of municipal solid waste (MSW) was
significantly influenced by the addition of fly ash-quicklime. Mixing fly ash-quicklime into
the MSW specimens increased the unconfined compressive strength (UCS) of the
specimens. The optimum content of fly ash-quicklime needed to increase the compressive
strength of the MSW specimen is around 20% of its dry weight. A longer curing time
enhanced the unconfined compressive strength of the treated specimens. Based on the
180
results of the direct shear tests, it was shown that by increasing the content of fly ash-
quicklime in the MSW specimens, the peak shear stress of the treated specimens increased,
and the optimum content of fly ash-quicklime needed to increase the shear stress of the
MSW specimens was about 20% of its dry weight. In addition, by mixing up to 20% of fly
ash-quicklime into the MSW specimen, the friction angle and cohesion of the stabilised
samples increased. The shear strength parameters of specimens treated with fly ash-
quicklime were also enhanced slightly with curing time. The results of constant head
permeability tests showed that by increasing the content of fly ash-quicklime by 13% in the
MSW specimens, their coefficient of permeability was rapidly reduced but when more
additives were added there were no significant changes in the coefficient of permeability.
This reduction in permeability can be attributed to the conversion of soluble calcium
hydroxide to cementitious compounds and a subsequent reduction in the bleed channels and
void spaces. The results of permeability tests while the specimen was consolidating showed
that by increasing the effective confining pressure up to the preconsolidation pressure, there
was no significant change in the coefficient of permeability, but at higher pressures, the
coefficient of permeability was reduced. Furthermore, based on the results of the triaxial
tests, the treatment of MSW with fly ash-quicklime caused significant changes in the peak
strength, residual response, and the stiffness and brittleness of the MSW specimens. By
increasing the contents of fly ash-quicklime in the MSW specimens, the shear strength and
stiffness increased significantly, while the residual strength experienced a 30% increase
after a fly ash-quicklime admixture of up to 13.3% was added, and it remained relatively
constant even after any more fly ash-quicklime was added. The brittleness index of MSW
specimens treated with up to 20% of fly ash-quicklime increased, but it remained relatively
constant after more was added, whereas the brittleness index of specimens under higher
effective confining pressure were slightly lower. Based on the results of consolidation and
creep tests, the addition of fly ash-quicklime reduced the compressibility characteristics of
MSW specimens. The result showed that the value of primary and secondary compression
indices decreased after treatment with fly ash-quicklime, whereas the curing time did not
affect these indices.
181
CHAPTER 6
6 Numerical Analysis to Predict the Settlement of
Closed Landfills
6.1 Introduction
Landfill design and construction technology has progressed rapidly during the past two
decades in reaction to more strict controlling requirement and demands. However before
any construction on top of landfills, properties of decomposed waste materials should be
improved as required. Whether immediately or long time after construction, damage and
cracks in structures are a challenging issue for organisation that design and construct
foundations on improved landfill sites. Generally clients require a maximum post
construction settlement less than a certain value (e.g. 100 – 200 mm) over the structure’s
life-time, and limit the differential settlement to a certain change in grade. This chapter
briefs numerical analysis to predict the vertical and horizontal displacement of closed
landfills under traffic load with and without treatment by fly ash-quicklime using the finite
element program PLAXIS.
182
6.2 Finite Element Modeling
The finite element program PLAXIS version 9 has been used to evaluate the settlement of
stabilised landfill. The cross-section utilised for the numerical analysis is presented in
Figure 6.1. Five layers of solid waste were considered for the landfill to estimate the effect
of depth of stabilisation on landfill settlement. All the dimensions for model are given in
Figure 6.2. Accordingly, the solid waste properties need to be adjusted to consider various
improvement options. The landfill is modeled as a two dimensional plane strain model.
The 15 nodded triangle elements were employed in the modeling as shown in Figure 6.3
and the mesh generation algorithm of PLAXIS version 9.0 was used in this study. It
provides a fourth order interpolation for displacements and the numerical integration
involves twelve stress points. The powerful 15-node element provides an accurate
calculation of stresses and failure loads. The two vertical boundaries on both sides are free
to move in vertical direction, whereas the horizontal boundary at the base is considered to
be fixed in both vertical and horizontal directions as presented in Figure 6.2. The
foundation soil was considered to be stiff soil and its stability and deformation were not
considered directly in this analysis. Cross-sections of generated mesh have been shown in
Figures 6.4 and 6.5.
183
6m
11 m 3m
C 3m
3m
3m
24 m
12 m
30 m 36 m 30 m
B
184
Figure 6.5 Closer view of generated mesh
All soils exhibit some creep, and primary compression is thus always followed by a certain
amount of secondary compression. The secondary compression (for instance during a
period of 10 or 30 years) can be a certain percentage of the primary compression. This is
for instance the case when constructing embankments over closed landfill sites. Indeed, the
primary settlement of footings and embankments are usually followed by a substantial
creep settlement in later years. In such cases it is desirable to estimate the creep from FEM-
computations. The soft soil creep model was used for this analysis due to several
characteristics such as:
x Stress-dependent stiffness (logarithmic compression behaviour)
x Distinction between primary loading and unloading-reloading
x Secondary compression
x Memory of pre-consolidation stress
x Failure behaviour according to the Mohr-Culomb criterion
The Soft Soil Creep model utilized in PLAXIS requires the following material constants:
¾ Failure parameters as in the Mohr-Coulomb model:
x Cohesion (c),
x Friction angle (φ),
x Dilatancy angle (ψ ),
185
¾ Basic stiffness parameters:
x Modified compression index (λ㧖),
x Modified swelling index (κ㧖),
x Modified creep index (μ㧖).
Instead of defining the stiffness by the basic stiffness parameters, alternative stiffness
parameters can be utilised. These material constants are given by:
x Compression index (Cc),
x Swelling index (Cs),
x Creep index for secondary compression (Cα),
x Initial void ratio (einit).
Modified compression index (λ㧖), modified swelling index (κ㧖) and modified creep index
(μ㧖) can be obtained directly from an isotropic compression test or/and an oedometer test.
However, we are also able to make use of the relationship to internationally normalised
parameters to get those advanced stiffness parameters (λ㧖, κ㧖, and μ㧖) from commonly
used alternative stiffness parameters.
Cc
O (6.1)
2.3(1 e)
2 Cs
N
2.3 1 e (6.2)
CD
P
2.3(1 e) (6.3)
where e is the void ratio of soil.
The employed Soft Soil Creep constitutive model (adopted in PLAXIS software) considers
logarithmic creep law for secondary compression and is the 3D extension of 1D creep
model obtained from oedometer-type strain conditions. The model considers Modified Cam
Clay model as well as some aspects of viscous behaviour of soil under compression. The
total strain in this model is calculated as the summation of the elastic and visco-plastic
strains. The visco-plastic part is separated into a part occurring while pore water pressure is
186
dissipating (traditional primary consolidation) and another part occurring after primary
consolidation. The model ignores the creep deformation due to the shearing and assumes a
constant creep rate with time and stress level. It should be noted that many researchers (e.g.
Mesri and Castro, 1987; Yin 1999) believe that the creep rate decreases with time, and
consequently constant creep rate assumption results in an over estimation and unrealistic
prediction of the long term deformations. The detailed explanation of the Soft Soil Creep
model can be found in Vermeer and Neher (1999).
Buisman (1936) proposed a constitutive law for creep after examining that the soft soil
settlement cannot entirely be described by the classic consolidation theory. In the
framework of 1D-secondary compression, the work of other researchers such as Bjerrum
(1967), and Garlanger (1972) should be mentioned. More mathematical lines of research on
3D-creep were followed by Sekiguchi (1977), Adachi and Oka (1982) and Borja and
Kavazanjian (1985). Butterfield (1979) proposed the following creep equation:
W c tc
HH H cH P . ln( ) (6.4)
Wc
where H cH is the deformation during consolidation. The modified creep index P describes
the secondary compression per logarithmic time increment. In this equation H H is the
logarithmic strain. The superscript ‘H’ is applied because the logarithmic strain
measurement was originally used by Hencky. In case of large strains both Butterfield
(1979) and Den Haan (1994) showed that logarithmic strain supersedes the traditional
engineering strain.
The time W c is not a material parameter either as it not only depends on consolidation but
also on the geometry of the tested sample. Janbu (1969) developed a construction for
evaluating the parameter P and the time W c from experimental data. This and the
187
Figure 6.6 Consolidation and creep behaviour in standard Oedometer tests (Wehnert 2000)
Wehnert (2000) presented the following calculations to formulate the creep equation. The
strain rate from Equation (6-4) is as below:
. P
H (6.5)
W c tc
For isotropic stress conditions the same modified creep index ( P ) was observed. Thus,
secondary compression part can be included in the equation. Accordingly, the total
volumetric strain can be writes as:
pc pcpc W tc
Hv H ve H vcr H vce H vccr H vac
cr
N ln( ) (O N ) ln( ) P ln( c ) (6.6)
p0c pcp 0 Wc
where H v is the total volumetric strain due to an increase in mean effective stress from p 0c
to p c in a time period of tc+t'. The total volumetric strain is divided into two components:
an elastic part and a visco-plastic creep part, marked by e and cr, superscripts, respectively.
The visco-plastic part can be separated into: during consolidation and after consolidation.
This is given by the subscripts c and ac. Figure 6.7 explains these relations.
188
Figure 6.7 Logarithmic relationship between volumetric strain and mean stress including
creep (after Wehnert 2000)
It is clear that isotropic consolidation line (IC-line) is not obtained after the end of
consolidation, but it is occurred after some creep.
Moreover, following Bjerrum (1967) study, it is supposed that the pre-consolidation stress
depends entirely on the amount of the creep strain being accumulated by time. Equation (6-
6) can be presented to capture this point:
pc p cp
Hv H ve H vcr N ln( ) (O N ) ln( ) (6.7)
p 0c p cp 0
where
'H vcr
p cp p cp 0 exp( ) (6.8)
O N
p cp W c tc
H vac
cr
H vcr H vccr (O N ) ln( ) P ln( ) (6.9)
p cpc Wc
189
If load is increased gradually and each load keep for period of t c t c W , and W =1 day. IC-
p cp
line with p'p = p' is then achieved. For OCR 1 , Equation (6.9) can be rewritten as:
pc
pc W c W tc
(O N ) ln( ) P ln( ) (6.10)
p cpc Wc
O N O N
W pc p cpc
( ) P
or Wc W .( ) P
(6.11)
Wc p cpc pc
Having derived an expression for W c it is now possible to formulate the differential creep
equation:
.
. . .
pc P
Hv H ve H cr
N . (6.12)
pc W c t c
v
. O N
. . .
p c P p cpc P
Hv H H e cr
N . ( ) (6.13)
p c W c p cp
v v
where p'p is defined in Equation (6.8). Combining Equations (6.13) and (6.11) gives:
. O N
. . .
pc P pc
Hv H H e cr
N . ( ) P
(6.14)
p c W p cp
v v
f p eq p eq
p
(6.15)
(Figure 6.8).
190
Figure 6.8 Yeild surface of the SS-model in p’-q plane (after Wehnert 2000)
O N
. . 1 P
p eq wp eq
H D Vc
1
( eq ) P
(6.16)
D W pp wV c
'H vcr
p eq
p p eq
p 0 exp( ) (6.17)
O N
wp eq
And α is defined as D . Hence the Soft Soil Creep model is an extension of the Soft
wp c
Soil model that takes creep into account.
Municipal solid waste parameters used in FEM analysis for different contents of stabiliser
are presented in Table 6.1. In addition, parameters used for compacted clay as a cover layer
of closed landfill model and road embankment are summarised in Table 6.2.
Some of the important waste properties required for the analyses are the unit weight, the
shear strength, the permeability, the primary compression index and the secondary
191
compression index. These parameters were obtained from the results of the extensive
laboratory program performed on the MSW samples as outlined in Chapter 5.
The unit weight of solid waste is an important factor in evaluating the settlement of
landfills. The unit weight values, estimated in this study (Chapter 4, Section 4.3) and used
in the analysis, are presented in Table 6.1. The unit weight values of MSW specimens
changed slightly for specimens mixed with different fly ash-quicklime contents.
The results obtained from consolidated drained (CD) triaxial tests confirmed that the shear
strength parameters are strongly influenced by the fly ash-quicklime content mixed with
MSW. The shear strength parameters of solid waste estimated in this study at different fly
ash-quicklime content (Chapter 5, Section 5.5.2) also presented in Table 6.1. It reveals that
c and M values of MSW specimens increased with increasing the fly ash-quicklime content
to MSW specimens.
In addition, the results from permeability tests in this study (Chapter 5, Section 5.4.3)
indicated that with an increase in fly ash-quicklime content in the MSW specimen, the
coefficient of permeability reduced. The corresponding numbers presented in Table 6.1.
Moreover, from consolidation test performed in this study (Chapter 5 Section 5.5.5), it has
been demonstrated that the value of Cc decreased by 30%, when the fly ash-quicklime
content increased from 0 to 26.7%. By increasing the fly ash-quicklime content, the treated
specimens indicated more resistance against the compressive loading and less
compressibility characteristics. Furthermore, the creep results in this study (Chapter 5,
Section 5.5.6) clearly shows that the value of the secondary compression index, Cα, for
MSW specimens decreased from 0.052 to 0.033, while fly ash-quicklime content increased
from 0 to 26.7%. This means that, for any given time period, the volume change during
secondary compression is more for untreated specimens than treated with fly ash-
quicklime. Hence, fly ash-quicklime admixture is effective in reducing the volume change
during the primary consolidation as well as the secondary consolidation. The corresponding
values for the primary compression index and the secondary compression index from this
study are also reported in Table 6.1.
192
Table 6.1 Parameters for soft soil creep model in FEM analysis
Saturated
Fly ash- Void Secondary
unit Cohesion Friction Permeability Recompression Compression
quicklime ratio compression
weight c angle m/day index index
content e0 index
γsat kN/m2 Φo Cr Cc
% 3 Cα
kN/m
Unit Poison’s
Ef
ID Material weight c ratio
φo ψo
model γt kPa kPa
kN/m
3 ν
Mohr-
Cover 17.5 10 25 0 2.5E4 0.35
Coulomb
Road Mohr-
18 5 30 5 6.5E4 0.3
embankment Coulomb
The settlement of proposed landfill model in PLAXIS was computed using ‘consolidation
analysis’ as a calculation type and ‘staged construction’ as a loading input. Based on Table
6.3 the effects of various fly ash-quicklime contents and different depths of improvement
were examined. The model presents 30 years old closed landfill and selected MSW
materials are decomposed waste. In the first stage of modelling, the filling process of
landfill was completed in 4 stages (each stage consisted of 4 metres over 30 days). After
filling the site with decomposed waste material, in landfill model, the waste was then
covered with a 0.5m layer of compacted clay. After this the model evaluated the settlement
of the landfill after 30 years of self weight of MSW. In the next stage a 1m thick road
embankment was constructed on the closed landfill model over a 6 month period, and then
a traffic loading of 20 kPa with a reduction factor of 0.5 was applied on to the road
embankment. In the final step, the vertical and horizontal displacement of the landfill
model 10 and 20 years after applying the traffic load for different amounts of fly ash-
quicklime and various depths of improvement were calculated.
Table 6.3 Fly ash-quicklime contents and depths of treatment for closed landfill model
Addative
5%FA+1.7%QL 10% FA+ 3.3% QL 15%FA+5%QL 20%FA+6.7%QL
content(%)
Treatment
0,3,6,9,12,24 0,3,6,9,12,24 0,3,6,9,12,24 0,3,6,9,12,24
depths (m)
FA=fly ash, QL=quicklime
194
6.7 Results and Discussion
The results of the numerical predictions of the model were estimated in PLAXIS. Table 6.4
presents the vertical displacements of the landfill model 10 years after applying the traffic
load at the midpoint below the embankment (point C in Figure 6.2). These numbers
reported based on different depths of treatment and various fly ash-quicklime contents. The
results in Figure 6.9 predicted that the vertical settlement of untreated landfill model is
about 370 mm 10 years after applying traffic load. In addition, the result illustrated that
treating MSW with fly ash-quicklime reduced the vertical displacement of the model
significantly. This reduction is more with higher depths of improvement. Furthermore,
Figures 6.10 to 6.14 presented vertical displacement of landfill model treated with 26.7%
fly ash-quicklime with different depths of improvement, 10 years after applying load. It
revealed that for 3m improved landfill with 26.7% fly ash-quicklime the vertical settlement
reduced 20% (from 370 mm to 296 mm) and this reduction for 6m, 9m, 12m and 24m
improved landfill was 32%, 40%, 46% and 58%, respectively.
Table 6.4 Vertical displacement of the landfill model 10 years after applying traffic load
195
Figure 6.9 Vertical displacement for untreated landfill 10 years after applying traffic load
Figure 6.10 Vertical displacement for 3-m treated with 26.6% fly ash-quicklime 10 years
after applying traffic load
196
Figure 6.11 Vertical displacement for 6-m treated with 26.6% fly ash-quicklime 10 years
after applying traffic load
Figure 6.12 Vertical displacement for 9-m treated with 26.6% fly ash-quicklime 10 years
after applying traffic load
197
Figure 6.13 Vertical displacement for 12-m treated with 26.6% fly ash-quicklime 10 years
after applying traffic load
Figure 6.14 Vertical displacement for 24-m treated with 26.6% fly ash-quicklime 10 years
after applying traffic load
198
6.7.2 Vertical Settlement 20 Years after Applying the Traffic Load:
The outcomes of the numerical estimation of the model are predicted in PLAXIS. Table 6.5
shows the vertical displacement of the landfill model 20 years after applying the traffic load
at the midpoint below the embankment. These numbers are based on different depths of
treatment and various amounts of fly ash-quicklime. The results showed that the settlement
of the untreated landfill model is about 536 mm 20 years after applying traffic load.
Moreover, these result indicated that treatment of MSW with fly ash-quicklime reduced
vertical displacement of the model significantly, and this reduction was more for higher
depths of improvement. It showed that 20 years after applying traffic load on 3m improved
landfill with 26.7% fly ash-quicklime the vertical settlement reduced by 19% (from 536
mm to 444 mm) and this reduction for 6m, 9m, 12m, and 24m improved landfill is 29%,
37%, 43%, and 55%, respectively.
Table 6.5 Vertical displacements of the landfill model, 20 years after applying the traffic
load
Furthermore, Figures 6.15 to 6.19 indicate the effects that the amounts of fly ash-quicklime
have on the vertical displacement versus time for 3m to 24m improvement depths at the
midpoint below the embankment. These figures clearly indicated that increasing the amount
of fly ash-quicklime reduced vertical settlement to a large extent. Figures 6.20 to 6.23 show
the effect of depths of improvement on the vertical displacement versus time for different
199
fly ash-quicklime contents. It can be inferred from these figures that increasing the depth of
improvement significantly decreases vertical settlement in the landfill model.
Time (Days)
1 10 100 1,000 10,000
0
10 years
-5
-10 20 years
-15
Settlement (mm)
-20
-25
-30
-35 3-m Improvement
-40
20% fly ash- 6.7% quicklime
-45 15% fly ash- 5% quicklime
-50 10% fly ash- 3.3% quicklime
5% fly ash- 1.7% quicklime
-55
Untreated landfill
-60
Figure 6.15 Vertical settlement versus time for 3-m improved landfill with various fly ash-
quicklime contents under traffic load
Time (Days)
1 10 100 1,000 10,000
0
10 years
-5
-10 20 years
-15
Settlement (mm)
-20
-25
-30
-35 6-m Improvement
-40 20% fly ash- 6.7% quicklime
-45 15% fly ash- 5% quicklime
-50 10% fly ash- 3.3% quicklime
5% fly ash- 1.7% quicklime
-55
Untreated landfill
-60
Figure 6.16 Vertical settlement versus time for 6-m improved landfill with various fly ash-
quicklime contents under traffic load
200
Time (Days)
1 10 100 1,000 10,000
0
10 years
-5
-10 20 years
-15
Settlement (mm)
-20
-25
-30
-35 9-m Improvement
-40
20% fly ash- 6.7% quicklime
-45 15% fly ash- 5% quicklime
-50 10% fly ash- 3.3% quicklime
-55 5% fly ash- 1.7% quicklime
Untreated landfill
-60
Figure 6.17 Vertical settlement versus time for 9-m improved landfill with various fly ash-
quicklime contents under traffic load
Time (Days)
1 10 100 1,000 10,000
0
-5
-10
-15
-20
Settlement (mm)
-25
-30
-35 12-m Improvement
-40
20% fly ash- 6.7% quicklime 10 years
-45 15% fly ash- 5% quicklime
-50 10% fly ash- 3.3% quicklime 20 years
5% fly ash- 1.7% quicklime
-55
Untreated landfill
-60
Figure 6.18 Vertical settlement versus time for 12-m improved landfill with various fly ash-
quicklime contents under traffic load
201
Time (Days)
1 10 100 1,000 10,000
0
-5
-10
-15
-20
Settlement (mm)
-25
-30
-35 24-m Improvement
-40
20% fly ash- 6.7% quicklime 10 years
-45 15% fly ash- 5% quicklime
-50 10% fly ash- 3.3% quicklime
5% fly ash- 1.7% quicklime 20 years
-55
Untreated landfill
-60
Figure 6.19 Vertical settlement versus time for 24-m improved landfill with various fly ash-
quicklime contents under traffic load
Time (Days)
1 10 100 1,000 10,000
0
10 years
-5
-10 20 years
-15
-20
Settlement (mm)
-25
-30 Treated with 5% fly ash-1.7% quicklime
-35
24m Improvement
-40 12m Improvement
-45 9m Improvement
-50 6m Improvement
3m Improvement
-55
Untreated landfill
-60
Figure 6.20 Vertical settlement versus time for landfill treated with 6.7% fly ash-quicklime
content under traffic load
202
Time (Days)
1 10 100 1,000 10,000
0
10 years
-5
-10 20 years
-15
-20
Settlement (mm)
-25
-30 Treated with 10% fly ash-3.3% quicklime
-35
24m Improvement
-40 12m Improvement
-45 9m Improvement
-50 6m Improvement
3m Improvement
-55 Untreated landfill
-60
Figure 6.21 Vertical settlement versus time for landfill treated with 13.3% fly ash-
quicklime content under traffic load
Time (Days)
1 10 100 1,000 10,000
0
-5
-10
-15
-20
Settlement (mm)
-25
-30 Treated with 15% fly ash-5% quicklime
-35
24m Improvement
-40 12m Improvement 10 years
-45 9m Improvement
-50 6m Improvement
20 years
3m Improvement
-55
Untreated landfill
-60
Figure 6.22 Vertical settlement versus time for landfill treated with 20% fly ash-quicklime
content under traffic load
203
Time (Days)
1 10 100 1,000 10,000
0
-5
-10
-15
-20
Settlement (mm)
-25
-30 Treated with 20% fly ash-6.7% quicklime
-35
24m Improvement
-40 12m Improvement 10 years
-45 9m Improvement
-50 6m Improvement
20 years
3m Improvement
-55
Untreated landfill
-60
Figure 6.23 Vertical settlement versus time for landfill treated with 26.7% fly ash-
quicklime content under traffic load
The results in Figure 6.24 predicted the horizontal displacement of untreated landfill 10
years after applying a traffic load. In addition, this result showed that the treatment of
MSW with fly ash-quicklime reduced the maximum horizontal displacement of the model,
and this reduction was more for higher depths of improvement. Figures 6.25 to 6.27 show
the horizontal displacement of landfill model treated with 26.7% fly ash-quicklime for
different depths of improvement, 10 years after applying a traffic load, while Figure 6.28
shows the horizontal displacement versus depth for landfill treated with 26.7% fly ash-
quicklime with various depths of improvement in a section below the toe of embankment
(Section B-B in Figure 6.2). It can be seen that at any specific depth below the toe of
embankment, with increasing depth of improvement, the horizontal displacement was
reduced and maximum horizontal displacement occurred at higher depths. Moreover,
Figure 6.29 shows the horizontal displacement versus depth for 9m improved landfill
treated with various amounts of fly ash-quicklime below the toe of embankment. This
proves that increasing the amount of fly ash-quicklime reduced the horizontal displacement
204
at any specific depth, and also that maximum horizontal displacement occurred at higher
depths.
Figure 6.24 Horizontal displacements for untreated landfill 10 years after applying load
Figure 6.25 Horizontal displacement for 6-m treated with 26.6% fly ash-quicklime 10 years
after applying load
205
Figure 6.26 Horizontal displacement for 12-m treated with 26.6% fly ash-quicklime 10
years after applying load
Figure 6.27 Horizontal displacement for 24-m treated with 26.6% fly ash-quicklime 10
years after applying load
206
Horizontal displacement (mm)
0 10 20 30 40
0
-5
Depth (m)
-10
-15
20% fly ash-6.7quicklime
untreated
3-m Improvement
-20 6-m Improvement
9-m Improvement
12-m Improvement
24-m Improvement
-25
Figure 6.28 Horizontal displacement versus depth for the landfill treated with 26.7% fly ash-
quicklime content in various improvement depths at a section below the toe of embankment
0 10 20 30 40
0
-5
Depth (m)
-10
-15
9-m Improvement
untreated
-20 5% fly ash- 1.7% quicklime
10% fly ash- 3.3% quicklime
15% fly ash- 5% quicklime
20% fly ash-6.7% quicklime
-25
Figure 6.29 Horizontal displacement versus depth for 9-m improved landfill treated with various
fly ash-quicklime contents at a section below the toe of embankment
207
6.8 Summary
A finite element program, PLAXIS version 9, was used to evaluate the settlement of the
landfill model under traffic load, with and without treatment by fly ash-quicklime. The soft
soil creep model was used for this analysis. Five layers of solid waste were used for the
landfill model to evaluate how the depth of stabilisation affected the vertical and horizontal
displacement of the model. Treated and untreated MSW parameters used for the model
were obtained from the results of the extensive laboratory program performed on treated
and untreated MSW samples, as outlined in Chapter 5. The vertical and horizontal
displacement of the landfill model 10 and 20 years after applying the traffic load, and with
different amounts of fly ash-quicklime and various depths of improvement, were estimated.
The results of the model analyses with PLAXIS showed that treating MSW with fly ash-
quicklime reduced the vertical displacement under traffic load at the midpoint below the
embankment. Horizontal displacement of the landfill model at a specific depth in a section
below the toe of the embankment under traffic load was also significantly reduced. These
reductions in vertical and horizontal displacement were greater with higher depths of
improvement.
208
CHAPTER 7
7 Summary and Conclusions
7.1 Summary
Any form of construction on top of closed landfills is a challenging task due to the complex
behaviour of settlement, and the weak shear strength of waste. The potential for using these
sites for redevelopment remains high in spite of the potential risk for human health and the
environment. This study made a broad examination of the effectiveness of treating the
geotechnical properties of municipal solid waste (MSW) taken from a closed landfill with
fly ash-quicklime, the results of which provide details on the compaction, shear strength,
stiffness, permeability and compressibility of municipal solid waste. These findings also
include the improved engineering properties of MSW stabilised with different percentages
of fly ash-quicklime.
The samples of decomposed municipal solid waste used in this study were collected from
former Bankstown landfill, a landfill site in the southwest of Sydney that was closed aroud
30 years ago. The main in-situ activities included excavating and bulk sampling of waste
materials. A Dynamic Cone Penetrometer (DCP) test was conducted on the landfill site but
the existence of large size particles such as pieces of brick and construction rubble mixed
209
with putrescible materials meant that the results were inconsistent and therefore have not
been reported in detail. After completing the excavation, the materials were visually
characterised and representative samples of the material were collected. Twelve, 70 litre
drums were filled with waste materials extracted from different depths. The materials were
then divided into particles that were smaller and larger than 19 mm. Particles larger than 19
mm in MSW were generally not considered to be decomposed waste materials, but particles
smaller than 19 mm were characterised according to typical geotechnical practices such as
the moisture content, organic content and sieve analysis. The field moisture content of the
samples ranged from 19.8% to 23.3%, and the measured organic content were between
20% and 28%. Samples were taken from the collected MSW depending on the size of the
particles; those with a maximum of 19 mm were classified as silty sandy gravel (GM), and
those with a maximum of 9 mm were classified as silty sand (SM), as per the Unified Soil
Classification System. Material smaller than 9 mm had a solid specific gravity (Gs) of 1.8.
The grain size distribution showed that 22% of the particles were in the range of gravel,
44% in the range of sand, and 34% were fine grained materials, with a coefficient of
uniformity (Cu) of 58. Atterberg limits of the fine portion of material were: 38% Liquid
Limit (LL), and 28% Plastic Limit (PL), which yielded a Plasticity Index (PI) of 10%.
Eraring fly ash and quicklime were used as additives to waste material for stabilisation.
Eraring fly ash is a natural pozzolan, and reacts chemically with calcium hydroxide in the
presence of moisture. The reaction of fly ash with lime results in an immediate
improvement of the soil’s mechanical properties during stabilisation because it reduces the
moisture content of the soil, which has a fast stabilising effect, increases the soil pH in
preparation for secondary pozzolanic reactions, and it also produces heat which accelerates
the chemical reactions.
The maximum dry density and optimum water content of untreated MSW samples with
particles smaller than 19 mm and 9 mm were measured using a 2.7 kg hammer and 25
blows per layer. The treated MSW samples were then prepared by mixing MSW materials
and fly ash-quicklime in various ranges, after which the optimum water content and
maximum dry density of the treated MSW samples were determined. When fly ash-
quicklime was mixed into the MSW samples, the maximum dry density increased and the
210
optimum moisture content decreased. By mixing 26.7% fly ash-quicklime with MSW
samples that were smaller than 9 mm, the maximum dry density increased from 1243 kg/m3
to 1292 kg/m3 and the optimum moisture content decreased from 15.7% to 13.3%, and
when 26.7% of fly ash-quicklime was mixed with MSW samples that were smaller than
19mm, the maximum dry density increased from 1442 kg/m3 to 1502 kg/m3 and the
optimum moisture content decreased from 14.4% to 12.2%. Because the MSW samples in
landfills were not compacted as tightly as normal layers of soil, a revised compaction test
was carried out by applying 15 blows per layer. All the samples prepared for this study
were also compacted to achieve the revised maximum dry density. Because the moisture
content of the MSW samples was on the wet side of optimum, a dry mixing method (no
water added to MSW samples during mixing) was adopted.
The results from the experimental program of this research can be summarised as follows:
x The unconfined compressive strength of MSW specimens increased by 150% from 102
kPa for an untreated specimen to 256 kPa for a specimen treated with 20% fly ash-
quicklime and cured for 7 days. The optimum content of fly ash-quicklime required to
increase the compressive strength of the MSW specimen was around 20% of its dry
weight, although there was a slight reduction in the compressive strength when the
content of fly ash-quicklime was more than 20%. Furthermore, the elasticity modulus
at the first loading (Esecant), the failure axial strain (εf), and the brittleness changed as a
consequence of including fly ash-quicklime, indeed its addition led to an increment of
Esecant and a decrement of the failure axial strain (εf) in the specimens. An increase in
the curing time from 7 days to 93 days enhanced the unconfined compressive strength
of the treated specimens and the MSW specimens had a 91% increase in their
unconfined compressive strength when 20% of fly ash-quicklime was mixed into them
(i.e. an increase from 256 kPa to 490 kPa).
x Based on the results of direct shear tests, increasing the content of fly ash-quicklime in
MSW increased the value of the peak shear stress of treated specimens. The optimum
211
content of fly ash-quicklime required to increase the shear stress of MSW specimen
was estimated to be 20% of the its dry weight, although there was a slight reduction in
the shear stress when the content was more than 20%. The cohesion and friction angles
of untreated MSW specimens were 17 kPa and 31º, respectively. By mixing 20% fly
ash-quicklime to an MSW specimen, the friction angle and cohesion increased from
31º to 39º and from 17 kPa to 33 kPa, respectively. The shear strength parameters of
specimens treated with fly ash-quicklime increased slightly with a longer curing time
compared to the untreated MSW specimens. By increasing the curing time of
specimens treated with 20% fly ash-quicklime from 28 days to 93 days, the friction
angle increased from 39º to 40º and the cohesion value increased from 33 kPa to 40
kPa. An increase in the normal stress intensified the peak shear stress and the
maximum vertical displacement for the specimen tested under a higher normal stress
had a lower value. The peak vertical displacement had a higher value for treated MSW
specimen with the higher amount of fly ash-quicklime.
x Based on the results of constant head permeability tests with an increasing rate of fly
ash-quicklime, the flow rate in the MSW specimens was reduced and the coefficient of
permeability decreased after treatment. By mixing 6.7% and 13.3% fly ash-quicklime
with the MSW samples the coefficient of permeability reduced rapidly from 1.6u10-5
m/s to 7u10-6 m/s and 2.4u10-6 m/s, respectively, but increasing the content of fly ash-
quicklime from 13.3% to 26.7% made no significant changes in the coefficient of
permeability. This reduction in the coefficient of permeability can be attributed to a
reduction in the bleed channels and void spaces due to the conversion of soluble
calcium hydroxide to cementitious compounds. An increase in density also caused a
reduction in the permeability of the small, finely divided particles of fly ash which act
like micro-aggregates and facilitates to fill the voids in the MSW samples. No
significant changes were observed in the coefficient of permeability with curing time.
x The coefficient of consolidation was reduced by increasing the fly ash-quicklime in the
specimen. Under 200 kPa of average effective confining pressure, the cv values were
1.6×10-5 m2/s and 1.2×10-5 m2/s for the untreated specimens and the specimens mixed
with 26% fly ash-quicklime, respectively. With an increase in the confining pressure
up to the preconsolidation pressure, the cv value decreased and for higher confining
pressures the change was insignificant.
x Based on a linear regression analysis, the relationship between the void ratio and
permeability was found for different contents of fly ash-quicklime, such that increasing
the fly ash-quicklime reduced the ck value. By mixing 26% fly ash-quicklime, the ck
value reduced from 0.38 (for untreated specimen) to 0.21, while the ck values were
0.34, 0.28 and 0.25 for specimens with 20%, 13.6% and 6.7% fly ash-quicklime
contents, respectively.
x From the triaxial tests results it was found that increasing the fly ash-quicklime content
caused a significant change in the peak strength, residual response, the stiffness and
brittleness of the MSW specimens. The peak principal stress difference of MSW
specimens under an effective confining pressure of 300 kPa increased from 600 kPa
(without treatment) to 1150 kPa (mixed with 26.7% fly ash-quicklime). By increasing
the content of fly ash-quicklime in MSW specimens the shear strength and the stiffness
significantly increased, while the residual strength increased by 30% after adding a fly
ash-quicklime admixture up to 13.3%, and it remained relatively constant even after
extra fly ash-quicklime was added. Also, increasing the amount of additive resulted in
213
a volumetric response that was initially less compressive and then more expansive,
while the post peak behaviour became more brittle. The peak friction angle increased
from 29q to 39q and the cohesion intercept increased from 11 kPa to 30 kPa after the
content of fly ash–quicklime increased from 0 to 26.7%. The residual friction angle
also increased from 27q for untreated MSW specimens to 34.5q for treated MSW
specimens with a 26.7% content of fly ash-quicklime. Changes in the residual cohesion
values were not significant.
x By increasing the fly ash–quicklime from 0 to 26.7%, both the Ef and E50% increased
from a range of 3.1-5.5 MPa to 14-27.9 MPa and from 6.1-12.1 MPa to 19.6-36.1 MPa,
respectively, while these results indicated that the ratio of E50% to Ef of MSW samples
decreased when the content of fly ash–quicklime increased.
x The addition of fly ash-quicklime reduced the compressibility of the MSW specimens,
and filling the voids with particles of fly ash-quicklime reduced the void ratio of the
sample. Increasing the curing time from 28 days to 93 days increased the
preconsolidation pressure slightly, whereas there was no change in the compression
index.
x The primary compression index Cc, decreased from 0.33 for untreated MSW specimens
to 0.23 for specimens treated with 26.7% fly ash-quicklime. The values of Cc for
treated MSW specimens with 6.7%, 13.3% and 20% fly ash-quicklime content were
0.29, 0.28 and 0.24, respectively. As the content of fly ash-quicklime increased the
treated specimens would resist the compressive loading better and consequently
214
showed less compressibility characteristics. No change in the Cc values was observed
in specimens cured for 93 days.
x The change in the void ratio due to creep and the sliding of particles after the end of
primary consolidation was less in MSW specimens mixed with fly ash-quicklime. The
secondary compression index Cα of MSW specimens decreased from 0.052 to 0.033,
when the content of fly ash-quicklime increased from 0 to 26.7%. This means that for
any given period of time the change in volume during secondary compression was
greater for untreated specimens than specimens treated with fly ash-quicklime. Hence,
fly ash-quicklime effectively reduced the change in volume during primary
consolidation and secondary consolidation. Consequently, the amount of settlement of
structures built on fly ash-quicklime treated MSW decreased. This reduction in the
primary and secondary compression indices of MSW treated with fly ash-quicklime
can also be attributed to the cementation and pozzolanic activity in the fly ash-
quicklime. The curing time did not affect the secondary compression index of MSW
specimens when the effective confining pressure increased from 200 kPa to 300 kPa.
A finite element program, PLAXIS version 9, has been used to evaluate the settlement of
landfill model under traffic load with and without treatment by fly ash-quicklime. The soft
soil creep model was used for this analysis. Five layers of solid waste were considered for
the landfill to evaluate the effect of depth of stabilisation on settlement of landfill model.
Treated and untreated MSW parameters are obtained from the results of the extensive
laboratory program performed on treated and untreated MSW samples in this study. The
settlement of the landfill model 10 and 20 years after applying the traffic load for different
fly ash-quicklime contents and various depths of improvement was estimated. The findings
of numerical modelling are summarised as follows:
x From the results of model analyses with PLAXIS, treatment of MSW with fly ash-
quicklime reduced the vertical displacement of the landfill model under traffic load
significantly. This reduction is more with higher depths of improvement. For 3m
improved landfill with 26.7% fly ash-quicklime, the vertical settlement 10 years after
applying traffic load below the embankment was reduced 20% (from 370 mm to 296
215
mm) and this reduction for 6m, 9m, 12m and 24m improved landfill was 32%, 40%,
46% and 58%, respectively. Treatment of MSW with fly ash-quicklime also reduced
the maximum horizontal displacement of the model below the toe of embankment.
This reduction is more for the model with higher depths of improvement.
Based on the findings of this research, and in an effort to gain a better understanding of the
mechanics of waste materials, a number of research activities are recommended for future
study:
216
investigated. The original technique was known internationally as the deep mixing
method (DMM), which is an in-situ soil treatment technology where in-situ properties
of MSW in closed landfills, blended with cementitious materials can be examined as a
field study.
x Finite element analyses should be performed to examine the effects of incompatible
strain and progressive failure in the stability of waste slopes. These analyses may
provide a more robust criterion to define the shear strength of MSW. Evaluation of
failed waste slopes may provide useful insight on closed landfill behaviour and further
numerical study is recommended in this area.
x The MSW in actual site are heterogeneous due to existence of different types of waste
material but for the sake of simplicity in the FEM waste materials assumed
homogeneous. For further FEM analysis the effect of heterogeneity of wastes in model
can be considered.
x Verification for finite elements results can be performed to check the reliability of the
model. One way is to simulate triaxial-test with an axisymmetric model with number of
elements.
217
References
Acosta, H.A. (2002). “Stabilization of soft subgrade soils using fly ash”, M.Sc. Thesis,
University of Wisconsin-Madison.
Adachi, T. and Oka, F. (1982). “Constitutive equation for normally consolidated clays
based on elasto-viscoplasticity”, Soils and Foundations, vol. 22, pp. 57–70.
Aimin, X. and Sarkar, S.L. (1991). “Microstructural study of gypsum activated fly ash
hydration in cement paste”, cement and concrete research, Vol 21, No 6, pp. 1137–
1147.
Ampera, B. and Aydogmus, T. (2005). “Recent experiences with cement and lime-
stabilization of local typical poor cohesive soil”, Geotechnik-Kolloquium Freiberg,
March 11, 2005, Heft 2005-2, pp. 121-144.
Anon (1985). “Lime Stabilization Construction Manual”, Eighth Edition. National Lime
Association, Arlington, Va.
Anon (1990). “Lime Stabilization Manual”, British Aggregate Construction Materials
Industry, London.
AS 1289.2.1.1. (2005). “Determination of the moisture content of a soil - Oven drying
method”.
AS 1289.3.6.1. (2009). “Determination of the particle size distribution of a soil - Standard
method of analysis by sieving”.
AS 1289.4.1.1. (1997). “Determination of the organic matter content of a soil”.
AS 1289.5.1.1. (2003). “Determination of the dry density/moisture content relation of a
soil using standard compactive effort”.
AS 1289.6.2.2. (1998). “Determination of the shear strength of a soil - Direct shear test
using a shear box”.
AS 1289.6.4.2. (1998). “Determination of compressive strength of a soil - Compressive
strength of a saturated specimen tested in triaxial compression with measurement
of pore water pressure”.
AS 1289.6.6.1. (1998). “Determination of the one-dimensional consolidation properties of
a soil”.
AS 1289.6.7.1. (2001). “Determination of permeability of a soil - Constant head method
for a remoulded specimen”.
ASTM C 618-12 (2012). “Standard Specification for Coal Fly Ash and Raw or Calcined
Natural Pozzolan for Use in Concrete”.
ASTM D 5239 (2004). “Standard Practice for Characterizing Fly Ash for Use in Soil
Stabilization”.
218
Augello, A.J., Bray, J.D., Seed, R.B., Matasovic, N., and Kavazanjian, Jr., E. (1998).
“Performance of Solid-Waste Landfill During the Northridge Earthquake”,
Proceedings, NEHRP Conference and Workshop on Research on the Northridge,
California Earthquake of January 17, 1994, California Universities for Research in
Earthquake Engineering, Los Angeles, CA, pp 71-80.
Augello, A.J., Matasovic, N., Bray, J. D., Kavazanjian, Jr. E. and Seed, R. B. (1995),
“Evaluation of solid waste landfill performance during the Northridge
earthquake”, Earthquake design and performance of solid waste landfills, ASCE
Geotechnical Special Publication no. 54, M. K. Yegian, W. D. L. Finn, eds.
Proceedings, ASCE Annual Convention, San Diego, CA, pp. 17-50.
Beaven, R.P. (2000). “The hydrogeological and geotechnical properties of household
waste in relation to sustainable landfilling”, PhD dissertation, Queen Mary and
Westfield College, University of London.
Beaven, R.P. and Powrie, W. (1995). “Hydrological and geotechnical properties of refuse
using a large scale compression cell”, Proceedings of Fifth International landfill
Symposium, Sardinia, Cagliari, volume 2, pp 745-760.
Beaven, R.P., Powrie, W. and Zardava, K. (2011). “Hydraulic properties of MSW”, In,
Zekkos, Dimitrios (ed.) Geotechnical characterization, Field Measurements and
Laboratory Testing of Municipal Solid Waste. Virginia, US, American Society of
Civil Engineers, pp. 1-43.
Beeghly, J. H. (2003). “Recent experiences with Lime-Fly ash stabilization of pavement
subgrade soils, base and recycled asphalt”, In International Ash Utilization
Symposium, Univ. of Kentucky, Center for Applied Energy Research, Lexington,
KY, pp. 1-18.
Bell, F.G. (1996). “Lime stabilization of clay minerals and soils”, Engineering Geology,
Vol. 42, No. 4, pp. 223-237.
Bjarngard, A. B. and Edgers, L. (1990). “Settlement of municipal solid waste landfills”
Proceedings of 13th Annual Madison Waste Conference, University of Wisconsin,
Madison, pp. 192–205.
Bjerrum, L. (1967). “Engineering geology of norwegian normally consolidated marine
clays as related settlements of buildings”, Geotechnique Vol. 17, No. 2, pp. 81-
118.
Blacklock, J. R., Joshi, R. C. and Wright, P. J. (1982). “Pressure Injection Grouting in
Landfills Using Lime and Fly Ash”, Grouting in Geotechnical Engineering,
ASCE, pp. 708-721.
Bleiker, D.E., Farquhar, G. and McBean E. (1995). “Landfill settlement and the impact on
site capacity and refuse hydraulic conductivity”, Waste Management and Research
Vol. 13, pp. 533-554.
Borja, R. I. and E. Kavaznjian (1985). “A constitutive model for the stress-strain-time
behaviour of wet clays”, Geotechnique Vol. 35, No. 3, pp. 283–298.
219
Bouzoubaa, N., Zhang, M.H., Bilodeau, A. and Malhotra, V.M. (1997). “The effect of
grinding on the physical properties of fly ashes and a Portland cement clinker”,
cement and concrete research, Vol 27, No. 12, pp 1861–1874.
Bray, J.D., Zekkos, D., Kavazanjian, Jr., E., Athanasopoulos, G.A., and Riemer, M.F.
(2009). “Shear Strength of Municipal Solid Waste”, Journal of Geotechnical and
Geoenvironmental Engineering, ASCE, Vol. 135, No. 6, pp. 709-722.
Brooks, R., Udoeyo, F. and Takkalapelli, K. (2011). ”Geotechnical Properties of Problem
Soils Stabilized with Fly Ash and Limestone Dust in Philadelphia”, Journal of
Materials in Civil Engineering, Vol. 23, No. 5, pp. 711–716.
Buchanan, D. (1997). “The impact of waste processing on the hydraulic behaviour of
landfilled wastes”, Designing and Managing Sustainable Landfill, Scientific
Societies Lecture Theatre (26-27 February) IBC UK Conferences Ltd.
Buchanan, D., Clark, C.F., Ferguson, N.S. and Kenny, N.J. (2001). “Hydraulic
characteristics of wet-pulverised municipal waste”, Journal of Water and
Environment, March, pp. 14-20.
Buisman, A. S. K., (1936). “Results of Long Duration Settlement Tests”, Proceeding of
the 1st International Conference on Soil Mechanics and Foundation Engineering,
Vol. 1, pp. 103-106.
Burkart, B., Gross, G.C. and Kern, J.P. (1999). “The role of gypsum in production of
sulfateinduced deformation of lime-stabilized soils”, Environmental and
Engineering Geoscience, Vol. 5, No. 2, pp. 173-187.
Burwood Council. (2010). “Wangal Park Draft Plan of Management”, Burwood. NSW
Butterfield, R. (1979). “A natural compression law for soils (an advance on e-log p’)”,
technical note, Geotechnique, Vol. 29, No. 4, pp. 469–480.
Caicedo, B., Giraldo, E., Yamin, L. and Soler, N. (2002b). "The landslide of Dona Juana
landfill in Bogota.A case study", Proceeding of the Fourth International Congress
on Environmental Geotechnics, Brazil, Vol. 1, pp. 171-175.
Caicedo, B., Yamin, L., Giraldo, E. and Coronado, O. (2002a). "Geomechanical
properties of municipal solid waste in Dona Juana sanitary landfill", Proceeding of
the Fourth International Congress on Environmental Geotechnics, Brazil, Vol. 1,
pp. 177-182.
Chen, T.H. and Chynoweth, D.P. (1995). “Hydraulic conductivity of compacted
municipal solid waste”, Bioresource Technology Vol. 51, No. 2-3, pp. 205-212.
Chen, Y. M., Ke, H., Fredlund, D. G., Zhan, L. T. and Xie, Y. (2010). "Secondary
Compression of Municipal Solid Wastes and a Compression Model for Predicting
Settlement of Municipal Solid Waste Landfills", Journal of Geotechnical and
Geoenvironmental Engineering, Vol. 136, No.5, pp.706-717.
Choquette, M., Berube, M.A. and Locat, J. (1987). “Mineralogical and microtextural
changes associated with lime stabilization of marine clays from Eastern Canada”,
Applied Clay Science, Vol. 2, pp. 215-232.
220
Cokca, E. (2001). “Use Of Class C Fly Ashes for the Stabilization of an Expansive Soil”,
Journal of Geotechnical and Geoenvironmental Engineering Vol. 127, No. 7, pp.
568-573.
Conner, J. R. (1990). “Chemical fixation and solidification of hazardous wastes”, Van
Nostrald-Reinhold, New York, pp. 280–427.
Consoli, N.C., Prietto, P.D.M., Carraro, J.A.H. and Heineck, K.S. (2001). “Behavior of
compacted soil-fly ash-crabide lime mixtures” Journal of Geotechnical and
Geoenvironmental Engineering, Vol. 127, No. 9, pp. 774-782.
Das, B. M. (1983). “Advanced Soil Mechanics”, Hemisphere Publishing Corporation.
Washington, DC, 511 pages.
Den Haan, E. J. (1994). ”Vertical Compression of Soils”, PhD thesis, Delft University
Diamond, S., White, J.L. and Dolch, W.L. (1964). “Transportation of clay minerals by
calcium hydroxide attack. Clays”, Clay Minerals, Vol. 12, pp. 359-379.
Dixon, N., Russell, D. and Jones, V. (2005). “Engineering properties of municipal solid
waste”, Geotextiles and Geomembranes, Vol. 23, No. 3, pp. 205-233.
Earth Tech Consultants (2001). “Settlement and density enhancement for leachate
recirculation landfills”, Proceeding of Waste Tech, San Diego.
Edil, T. B., Ranguette, V. J. and Wuellner, W. W. (1990). “Settlement of municipal
refuse” Geotechnics of waste fills, Theory and practice, STP1070, A. Landva and
G. D. Knowles, eds., ASTM, Philadelphia, pp 225–239.
Edil, T., Benson, C., Shafique, S., Tanyu, B., Kim, W., and Şenol A. (2002). “Field
evaluation of construction alternatives for roadway over soft subgrade”,
Geotechnical Engineering Report, 02-04. Madison, Wisconsin.
Fairfield city council report (2000), “Brenan Park, Smithfield, plan of management”.
Fang, H. Y. (1991). “Foundation engineering handbook”, Springer.
Fassett, J.B., Leonondro, G.A. and Reptto, P.C. (1994). "Geotechnical properties of
municipal solid wastes and prior use in landfill design", Proceedings, Waste Tech
94 Conference, National Solid Waste Management Association, Charleston, SC.
pp. 13-14.
Gabr, M.A. and Valero, S.N. (1995). “Geotechnical properties of municipal solid waste”,
Geotechnical Testing Journal, Vol. 18, pp. 241-254.
Garlanger, J. E. (1972). “The consolidation of soils exhibiting creep under constant
effective stress”, Geotechnique Vol 22, No. 1, pp. 71–78.
Geosyntec (2003). “Waste Characterization Report, Cherry Island Landfill Expansion
Project.”, Wilmington, Delaware, Project number ME0250, March.
Ghosh, A. and Subbarao, C. (2007). “Strength characteristics of class F fly ash modified
with lime and gypsum”, Journal of Geotechnical and Geoenvironmental
Engineering, Vol 133, No. 7, pp. 757–766.
221
Gibson, R. E. and Lo, K. Y. (1961). “A theory of soils exhibiting secondary
compression”, Acta Polytech. Scand, Vol. C, No.10, pp. 1–15.
Gomes, C., Ernesto, A., Lopes, M.L. and Moura, C. (2002), "Sanitary landfill of Santo
Tirso-municipal waste physical, chemical and mechanical properties", Proceeding
of the Fourth International Congress on Environmental Geotechnics, Brazil, Vol.
1, pp. 255-261.
Gourc, J. P., Olivier, F., Thomas, S., Chatelet, L., Denecheau, P. and Munoz, M. L.
(2001). “Monitoring of waste settlements on five landfills: Comparison of the
efficiency of different devices”, Proceedings Sardinia ‘01, 8th International Waste
Management and Landfill Symposium, Cagliari, Italy, pp. 515-524.
Gray, D.H. and Lin, Y.K. (1972). “Engineering properties of compacted fly ash”, Journal
of the Soil Mechanics and Foundations Division, Vol. 98, No. 4, pp. 361– 380.
Han-bing, L., Chang-yu L, Hai-bin W. (2009). “Experimental Research on the
Mechanical Parameter of Fly Ash Soil”, ICCTP: Critical Issues in Transportation
Systems Planning, Development, and Management , pp. 41-48
Henke, K. F. (1985), “Standsicherheit von Deponien aus Mull-und Abfallstoffen”, Mull-
und Abfallbeseitigung. Kennzahl 4546.
Hornsby Shire Council. (2004), “Dartford road, thorlnleigh report”.
Horpibulsuk, S., Phetchuay, C. and Chinkulkijniwat, A. (2011). “Soil stabilization by
calcium carbide residue and fly ash”, Journal of Materials in Civil Engineering,
Vol. 24, No. 2, pp. 184-193.
Hossain, M. S., and Gabr, M. A. (2005). “Prediction of municipal solid waste landfill
settlement with leachate recirculation”, In Proceedings of the Sessions of the Geo-
Frontiers 2005 Congress. Austin, Texas, USA, ASCE, Vol. 168, pp. 50.
Howland, J.D., Landva, A.O. (1992). “Stability analysis of a municipal solid waste
landfill”, American Society of Civil Engineers, ASCE. Geotechnical special
publication, Vol.31, pp. 1216-1231.
Huang, S. and Cheng, J. (1986). “Kinetic of reaction in the system fly ash- Ca(OH)2-
H2O”, Journal of Chinese Ceramic Society. Vol. 14, No. 1, pp. 191– 197.
Hudson, A.P., Beaven R.P. and Powrie W. (2009). “Assessment of vertical and horizontal
hydraulic conductivities of household waste in a large scale compression cell”,
Proceedings of the 12th Waste Management and Landfill Symposium Sardinia
2009, Ed by Cossu R, Diaz LF and Stegman R. CISA, S. Margherita di Pula, Italy
pp. 641-642.
Hudson, A.P., Beaven, R.P. and Powrie, W. (2001). “Interaction of water and gas in
saturated household waste in a large scale compression cell”, Proceedings of the
8th International Sardinia Landfill Conference, S.Margherita Dipula, Cagliari,
Italy. Vol. 3, pp. 585-594.
Hyder Consulting Pty Ltd., Oke, M., Allan. P., Goldsworthy, K. and Pickin, J. (2008).
“waste and recycling in Australia” A Report Prepared for the Department of
Environment, Water, Heritage and the Arts
222
Indraratna, B., Balasubramaniam, A.S. and Khan, M.K. (1995). “Effect of fly ash with
lime and cement on the behavior of soft clay”, Quarterly Journal of Engineering
Geology, Vol. 28, part 2, pp. 131-142.
Ismail, A.I.M. (2004). “Engineering and petrological characteristics of clayey silt soils to
be used as road base and their improvement by lime and cement”, Doctoral work
Thesis, Mathematisch-Naturwissenschaftliche Fakultaet der Technischen
Universitaet Clausthal.
Jain, P., Powell, J., Townsend, T.G. and Reinhart, D.R. (2006). “Estimating the hydraulic
conductivity of landfilled municipal solid waste using the borehole permeameter
test”, Journal of Environmental Engineering, ASCE Vol.132, pp. 645-652.
Janbu, N. (1985). “25th rankine lecture, Soil models in off shore engineering”,
Geotechnique Vol. 35, No. 3, pp. 239–281.
Janz, M. and Johansson, S.E. (2002). “The function of different binding agents in deep
stabilization”, Svensk djupstabilisering, Rapport 9.
Kavazanjian, E. Jr., Matasovic, N. (1995). “Seismic analysis of solid waste landfills”,
Proceedings of Geoenvironment 2000, ASCE Geotechnical Special Publication
No. 46, New York, NY, pp. 1066-1080.
Kavazanjian, E. Jr., Matasovic, N., Stokoe, K.H.II. and Bray, J.D. (1996). “In situ shear
wave velocity of solid waste from surface wave measurements”, Environmental
Geotechnics, Kamon, Balkema, Vol. 1, pp. 97-102.
Kavazanjian, E., Jr., Matasovic, N., Bonaparte, R., Schmertmann, G.R. (1995).
“Evaluation of MSW properties for seismic analysis”, Geoenvironment 2000,
ASCE Geotechnical Special Publication, No. 46, pp. 126-141.
Kawasaki, T., Niina, A., Saitoh, S., Suzuki, Y. and Honjo, Y. (1981). “Deep mixing
method using cement hardening agent”, Proceedings of 10th international
conference on soil mechanics and foundation engineering, pp. 721–724.
Kehew, EA. (1995). “Geology for engineers and environmental scientists”, 2nd ed. New
Jersey, Prentice Hall Englewood Cliffs, pp. 295–302.
Kitazume, M., Yamazaki. H. and Tsuchida, T. (2000). “Recent soil admixture
stabilization techniques for port and harbor constructions in Japan, deep mixing
method, premix method, light-weight method”, Proceedings of international
seminar on geotechnics in Kochi, Japan, pp. 23–40.
Klink, R. and Ham, R.K. (1982). “Effects of moisture movement on methane production
in solid waste landfill samples”, Resource Conservation, Vol. 8, pp. 29-41.
Korfiatis, G.P., Demetracopoulos, A.C., Bourodimos, E.L. and Nawy, E.G. (1984).
“Moisture transport in a solid waste column”, Journal of Environmental
Engineering, ASCE, Vol. 110, No.4, pp. 789-796.
Kumar, A., Walia, B. S. and Bajaj, A. (2007). “Influence of Fly Ash, Lime, and Polyester
Fibers on Compaction and Strength Properties of Expansive Soil”, Journal of
Materials in Civil Engineering, Vol. 19, No. 3, pp. 242-248.
223
Landva, A.O. and Clark, J. I. (1986). “Geotechnical testing of wastefill”, Proceedings,
39th Canadian Geotechnical Conference Ottawa, Ontario, pp. 371–385.
Landva, A.O. and Clark, J. I. (1990). “Geotechnics of waste fill, Theory and practice”
STP 1070, Landva and Knowles (ed.), ASTM, pp. 86-103.
Landva, A.O., Clark, J.I., Weisner, W.R. and Burwash, W.J. (1984) “Geotechnical
engineering and refuse landfills” Proc. 6th National Conference on Waste
management, Vancouver, pp. 1-37
Landva, A.O., Pelkey S.G. and Valsangkar, A.J. (1998). “Coefficient of permeability of
municipal refuse”, Environmental Geotechnics, Vols. 1-4, pp. 63-67.
Lav, A.H. and Lav, M.A. (2000). “Microstructural development of stabilized fly ash as
pavement base material”, Journal of Materials in Civil Engineering, Vol. 12, No.
2, pp. 157-163.
Leckie, J.O., Pacey, J.G. and Halvadakis, C. (1979). “Landfill management with moisture
control”, Journal of Environmental Engineering 105, pp. 337-347.
Lee, S. and Fishman, K. (1992). “Improved resilient modulus realized with waste product
mixtures”, Geotechnical Special Publication, Grouting, Soil improvement and
Geosynthetics, edited by Borden, R. and Holtz, R., published by ASCE, New
York. Vol. 2, No. 30, pp. 1356-1367.
Ling, H. I., Leshchinsky, D., Mohri, Y., and Kawabata, T. (1998). “Estimation of
municipal solid waste landfill settlement”, Journal of Geotechnical and
Geoenvironmental Engineering, Vol. 124, No. 1, pp. 21–28.
Mahler, C. F., and De Lamare Netto, A. (2003). “Shear resistance of mechanical
biological pre-treated domestic urban waste”, In Proceedings Sardinia, pp. 6-10.
Manassero, M., Van Impe, W.F. and Bouazza, A., (1996). “Waste Disposal and
containment”, Proceedings of 2nd International Congress on Environmental
Geotechnics, Osaka, Japan, Balkema, Rotterdam, Netherlands, Vol. 2, pp. 1425-
1474.
Manassero, M., Van Impe, W.F., Bouazza, A., (1996). “Waste disposal and containment”,
Proceeding of Second International Congress on Environmental Geotechnics,
Osaka, pp. 193-242.
Matasovic, N. and Kavazanjian, E.Jr. (1998). “Cyclic characterization of OII landfill solid
waste”, Journal of Geotechnical and Geoenvironmental Engineering, Vol. 124, No.
3, pp. 197-210.
Mazzucato, N., Simonini, P. and Colombo, S. (1999). “Analysis of block slide in a MSW
landfill”, Proceedings Sardinia 1999, Seventh International Waste Management
and Landfill symposium, Cagliari, Italy, 4-8 October 1999.
McDougall, J. (2011). “Settlement: The short and the long of it." Geotechnical
characterization, field measurement, and laboratory testing of municipal solid
waste”. Proceedings of the 2008 international symposium on waste mechanics,
Zekkos, D., ed., ASCE Geotechnical Special Publication No. 209, Geo- Institute,
Reston, Va, pp. 76-111.
224
McDougall, J. R. and Pyrah, I. C. (2001). “Settlement in landfilled waste: Extending the
geotechnical approach”, Proceedings Sardinia ‘01, 8th International Waste
Management and Landfill Symposium, Cagliari, Italy, pp. 481-490.
McDougall, J.R., Sarsby, R.W. and Hill, J.N. (1996). “A Numerical Investigation of
Landfill Hydraulics Using Variably Saturated Flow Theory”, Geotechnique, Vol.
46, No. 2, pp. 329-341.
Merry, S.M., Fritz, W.U., Budhu, M. and Jesionek, K. (2006). “The effect of gas on pore
pressures in wet landfills,” ASCE Journal of Geotechnical and Geoenvironmental
Engineering, Vol. 32, No. 5, pp. 553-561.
Mesri, G. and Castro, A. (1987). “Cα/Cc concept and K0 during secondary compression”,
Journal of Geotechnical Engineering, Vol. 113, No. 3, pp. 230-247.
Morris, D.V. and Woods, C.V. (1990). “Settlement and engineering design considerations
in landfill and final cover design”, Geotechnics of Waste Fills - Theory and
practice, ASTM STP 1070, eds. Landva, A.O. and Knowles, G.D., ASTM,
Philadelphia.
Muntohar, A.S. and Hantoro, G. (2000). “Influence of Rice Husk Ash and Lime on
engineering properties of a clayey sub-grade”, EJGE paper, pp. 2000-2094.
Nablantoglu, Z. and Tuncer, E. (2001). “Compressibility and hydraulic conductivity of a
chemically treated expansive clay”, Canadian Geotechnical Journal, Vol. 38, No.
1, pp. 154-160.
Nalbantoglu, Z. (2001). “Use of a self-cementing fly ash as a soil stabilization agent”,
Iranian Journal of Science and Technology, Vol. 25, No. 4, pp. 691-698.
Nalbantoglu, Z. and Gucbilmez, E. (2001). “Improvement of calcareous expansive soils in
semiarid environments”, Journal of Arid and Environments, Vol. 47, No. 4, pp.
453-463.
Nalbantoglu, Z. and Gucbilmez, E. (2002). “Utilization of an industrial waste in
calcareous expansive clay stabilization”, Geotechnical Testing Journal, Vol. 25,
No. 1, pp. 78-84.
NAVFAC, U.S. Department of the Navy. (1983). “Soil mechanics design manual 7.3”,
Naval Facilities Engineering Command, Alexandria, Va.
Nicholson, P. and Kashyap, V. (1993). “Fly ash stabilization of tropical Hawaiian soils”,
Geotechnical Special Publication, ASCE, New Work, N.Y., No. 36, , pp. 15-29.
Nicholson, P., Kashyap V. and Fuji C. (1994), “Lime and fly ash admixture improvement
of tropical Hawaiian soils”, Transportation Research Record, Washington, DC,
No. 1440, pp. 71-78.
Okoro, C., Vogtman, J., Yousif, M., Agnaou, M. and Khoury, N. (2011). “Consolidation
Characteristics of Soils Stabilized with Lime, Coal Combustion Products and
Plastic Waste”, Proceedings of the GeoFrontiers 2011, Dallas, Texas, USA, pp.
1202-1209.
225
Ormsby, W.C. and Kinter, E.B. (1973), “Effects of dolomitic and calcitic limes on
strength development in mixtures with two clay minerals”, Public Roads, 37, pp.
149-160.
Oweis, I. S. and Khera, R. (1998). “Geotechnology of waste management”, (2nd ed.),
PWS Publishing Company.
Park, H. I. and Lee, S. R. (1997). “Long-Term Settlement Behaviour of Landfills with
Refuse Decomposition”, Journal of Solid waste Technology and Management,
Vol. 24, No. 4, pp. 159–165.
Parson, R.L. and Milburn, J.P. (2003). “Engineering behavior of stabilized soils”,
Transportation research record, Journal of the Transportation Research Board,
Washington, DC- 0361-1981, pp. 20-29.
Pelkey, S.A., Valsangkar, A.J. and Landva, A. (2001). “Shear displacement dependent
strength of municipal solid waste and its major constituents”, Geotechnical Testing
Journal, Dec 2001, pp. 381-390
PhaniKumar, B. and Sharma, R. (2004). “Effect of Fly Ash on Engineering Properties of
Expansive Soils”, Journal of Geotechnical and Geoenvironmental Engineering,
Vol. 130, No. 7, pp. 764–767.
Phanikumar, B. and Sharma, R. (2007). ”Volume Change Behavior of Fly Ash-Stabilized
Clays” Journal of Materials in Civil Engineering. 19, Special Issue, Geochemical
Aspects of Stabilized Materials, pp. 67–74.
Pohland, F.G. (1975). “Sanitary Landfill Stabilization with Leachate Recycle and
Residual Treatment”, USEPA, Cincinnati, OH. EPA-600/2-75-043.
Powrie, W. and Beaven, R.P. (1999). “Hydraulic properties of household waste and
implications for landfills”, Proceedings of the Institution of Civil Engineers,
Geotechnical Engineering, Vol. 137, pp. 235-247.
Powrie, W., Beaven, R.P. and Hudson, A.H. (2008). “The influence of landfill gas on the
hydraulic conductivity of waste”, Proceedings of Geocongress 2008 Annual
Congress of the Geo-Institute of ASCE, Ed. ASCE, New Orleans, Louisiana,
USA, pp. 264-271.
Powrie, W., Beaven, R.P. and Hudson, A.P. (2005). “Factors affecting the hydraulic
conductivity of waste”, International Workshop "Hydro-physico-mechanics of
landfills", Ed. Gourc J.P., Grenoble.
Powrie, W., Richards, D. J. and Beaven, R. P. (1998). “Compression of waste and
implications for practice”, Geotechnical Engineering of Landfills, Thomas
Telford, pp. 3-18.
Prakash, K. and Sridharan, A. (1989), “Lime stabilization and curing effects on the index
and compaction characteristics of a montmorillonitic soil”, Journal of Southeast
Asian Geotechnical Society, Vol. 20, No. 1, pp. 39-47.
Qian, X., Koerner, R. M. and Gray, D. H. (2002). “Geotechnical aspects of landfill design
and construction”, New Jersey, Prentice Hall.
226
Qubain, B.S., Seksinsky, E.J. and Li, J.C. (2000). “Incorporating subgrade lime
stabilization intopavement design”, Geomaterials, Transportation Research
Record, (1721), pp. 3-8.
Rajasekaran, G., Murali, K. and Srinivasarghavan, R. (1997). “Fabric and mineralogical
studies on lime treated marine clays”, Ocean Engineering, Vol. 24, No. 3, pp. 227-
234.
Rajsekaran, G. and Narasimha Rao, S. (1998). “Particle size analysis of lime-treated
marine clays”, Geotechnical Testing Journal, Vol. 21, No. 2, pp. 109-119.
Reddy, K. R., Hettiarachchi, H., Parakalla, N. S., Gangathulasi, J. and Bogner, J. E.
(2009). “Geotechnical properties of fresh municipal solid waste at Orchard Hills
landfill, USA”, Waste Management Vol. 29, pp. 952–959
Richardson, G. and Reynolds, D. (1991). “Geosyntetic considerations in a landfill on
compressible clays”, Proceedings of Geosynthetics, Vol. 2, Industrial Fabrics
Association International, St. Paul, MN.
Sahu, B.K. (2001). “Improvement in California bearing ratio of various soils in Botswana
by fly ash”, Proceedings of the 2001 International Ash Utilization Symposium,
October 22, 2001, Lexington, Ky., 0-9674971-4-0, pp. 22-24.
Sanchez-Alciturri, J.M., Palma, J., Sagaseta, C. and Canizal, J. (1993). “Mechanical
properties of wastes in a sanitary landfill”, Waste Disposal by landfill, Green
1993, Sarsby (editor), 1995, Balkema, Rotterdam, pp. 357-363
Sekiguchi, H. (1977). “Rheological characteristics of clays. In Proceedings 9th
International Conference on Soil Mechanics and Foundation Engineering”,
Volume 1, Tokyo, pp. 289–292.
Senol, A., Bin-shafique, M.S., Edil T.B. and Benson, C.H. (2002). “Use of class C fly ash
for stabilization of soft subgrade”, Fifth International Congress on Advances in
Civil Engineering, 25-27 September 2002, Istanbul Technical University, Istanbul,
Turkey.
Shao, L., Liu, S., Du, Y., Jing, F. and Fang, L. (2008). “Experimental Study on the
Stabilization of Organic Clay with Fly Ash and Cement Mixed Method”, ASCE
Geotechnical Special Publication, Vol. 179, pp.20-27.
Sharma, H. D. (2000). “Solid waste landfills: Settlements and postclosure perspectives”,
Proceeding, ASCE National Conference on Environmental and Pipeline
Engineering, R. Y. Suranpalli, ed., ASCE, Reston, Va., pp. 447–455.
Sharma, H. D. and Reddy, K. R. (2004). “Geoenvironmental engineering site remediation,
waste containment, and emerging waste management technologies, Wiley,
Hoboken, N.J.
Sharma, H.D. and De Ariban, M. (2007). “Municipal Solid Waste Landfill Settlement,
Postclosure Perspectives”, Journal of Geotechnical and Geoenvironmental
Engineering, Vol.133, No.6, pp. 619-629.
227
Shi, C. (1999). “Studies on several factors affecting hydration and properties of lime-
pozzolan cements”, Journal of Materials in Civil Engineering, Vol. 29, No. 4, pp.
467–472.
Shi, C. and Day, R.L. (1995). “Acceleration of the reactivity of fly ash by chemical
activation”, Cement and Concrete Research, Vol. 25, No. 1, pp. 15-21.
Shi, C. and Day, R.L. (2000). “Pozzolanic reaction in the presence of chemical
activators”, Part 1. Reaction kinetics, Cement and Concrete Research, Vol. 30, No.
1, pp. 51-58.
Shimizu, K. (1997). “Geotechnics of waste landfill, Environmental Geotechnics, Kamon,
(editor), 1996 Balkema, Vol. 1, pp. 1475-1491.
Shirazi, H., (1999). “Field and laboratory evaluation of the use of lime and fly ash to
replace soil cement as a base course”, Seventh international conference on low-
volume Roads, Transportation Research Record, (1652), Vol. 1, pp. 270-275.
Siegel, R. A., Robertson, R. J. and Anderson, D. G. (1990). “Slope stability investigations
at a landfill in southern California”, American Society for Testing and Materials,
ASTM. Special Technical Publication. STP 1070, pp. 259-284.
Singh, D., Ghabchi, R., Laguros, J. G., and Zaman, M. (2010). “Laboratory Performance
Evaluation of Stabilized Sulfate Containing Soil with Lime and Class C Fly Ash”,
GeoFlorida 2010, pp. 757-766.
Sloane, R.L. (1965). “Early reaction determination in two hydroxide-kaolinite systems by
electron microscopy and diffraction”, Proceedings, 13 th National Conference on
Clays and Clay Minerals, New York, Pergamon press, pp. 331-339.
Solanki, P., and Zaman, M. (2011). “Characterization of Lime and Fly Ash-Stabilized Soil
by Indirect Tensile Testing”, Geo-Frontiers 2011, pp. 4438-4448.
Solanki, P., Khoury, N. and Zaman, M. (2009). ”Engineering Properties and Moisture
Susceptibility of Silty Clay Stabilized with Lime, Class C Fly Ash, and Cement
Kiln Dust”, Journal of Materials in Civil Engineering, Vol. 21, No. 12, pp. 749-
757.
Sowers, G. F. (1973). “Settlement of waste disposal fills” Proceeding, 8th International
Conference on Soil Mechanics and Foundation Engineering, Moscow, pp. 207-210
Staub, M., Galietti, B., Oxarango, L., Khire, M.V. and Gourc, J.P. (2009). “Porosity and
hydraulic conductivity of MSW using laboratory scale tests”, HPM3 – Third
international conference on Hydro-Physico-Mechanics of landfills Braunschweig,
Germany, 10th -13th March. 8 pp.
Tastan, E., Edil, T., Benson, C. and Aydilek, A. (2011). “Stabilization of Organic Soils
with Fly Ash”, Journal of Geotechnical and Geoenvironmental Engineering, Vol.
137, No. 9, pp. 819-833.
Taylor, D. W. (1948). “Fundamentals of Soil Mechanics”, New York: John Wiley and
Sons, Inc., New York.
Tchobanoglous, G. (1977) Solid Wastes, McGraw-Hill, New York.
228
Terzaghi, K. (1943). “Theoretical Soil Mechanics”, John Wiley and Sons. New York, NY.
510 pages.
Thomas, Z.G. and White D.J. (2003), “Engineering properties of self-cementing fly ash
subgrade mixtures”, Proceedings of the International Ash Utilization Symposium,
Lexington, pp. 346-378.
Toth, P. S., Chan, H. T., and Cragg, C. B. (1988). ‘‘Coal ash as a structural fill with
special reference to Ontario experience’’, Canadian Geotechnical Journal, Ottawa,
Vol. 25, pp. 694–704.
Towhata, I., Kawano, Y., Yonai, Y. and Koelsh, F. (2004). “Laboratory tests on dynamic
properties of Municipal Wastes”, Proceeding 11th Conference in Soil Dynamics
and Earthquake Engineering and 3rd International Conference on Earthquake
Geotechnical Engineering, Vol. 1, pp. 688-693.
Turner, J.P. (1997). “Evaluation of western caol fly ashes for stabilization of low-volume
roads. Testing Soil Mixed with Waste or Recycled Materials”, ASTM, STP 1275,
American Society of Testing and Materials, pp. 157-171.
Vermeer, P. A., and Neher, H. P. (1999). “A soft soil model that accounts for creep”,
Proceedings of the International Symposium Beyond 2000 in Computational
Geotechnics, pp. 249-261.
Vilar, O.M. and Carvalho, M.F. (2002). “Shear strength properties of municipal solid
waste” Proceeding of the Fourth International Congress on Environmental
Geotechnics, Brazil,Vol. 1, pp. 59-64.
Virendra, S. and Narendra, K. (1997). “Use of fly ash in base and subbase of roads”,
Proceedings of the International Conference on Soil Waste Technology and
Management, Vol. 1.
Wall, D.K. and Zeiss, C. (1995). “Municipal landfill biodegradation and settlement”,
Journal of Environmental Engineering, ASCE, Vol. 121, No 3, pp 214-223.
Waste Management Association of Australia. (2007), Landfill Survey.
Weber, A. (2001), “Verformungs und Festigkeitseigenschaften von verbessertem
Loesslehm”, Diplomarbeit, Bauhaus-Universitaet Weimar, Professur Grundbau,
unpublished.
Wehnert, H. R. M. (2000). “An evaluation of soft soil models based on trial
embankments”, In Computer Methods and Advances in Geomechanics:
Proceedings of the 10th International Conference on Computer Methods and
Advances in Geomechanics, Tucson, Arizona, USA, 7-12 January 2001, p. 373.
Wesche, K. (1991). “Fly ash in concrete: properties and performance”, CRC Press LLC,
Vol. 7
Wiemer, K. (1982). “Qualitative und quantitative Kriterien zur Bestimmung der Dichte
von Abfallen in geordnetern Deponien”, Disertation, TU Berlin.
229
Wild S., Abdi M.R. and Lengward G. (1993). “Sulfate expansion of lime-stabilized
kaolinite. 2. Reaction-Products and expansion”, Clay Minerals, Vol. 28, No. 4, pp.
569-583.
Withiam, J.L., Tarvin, P.A., Bushell, T.D., Snow, R.E. and German, H.W. (1995).
“Prediction and performance of Municipal Landfill slope”, Proceedings
International Conference The Geoenvironment 2000, ASCE GSP, NY, No. 46,
pp. 1005-1019.
Wray, W. K. (1986). “Measuring Engineering Properties of Soil”, Prentice-Hall, Inc.
Englewood Cliffs, NJ. 276 p.
Xie, Z. and Xi, Y. (2001). “Hardening mechanisms of an alkaline-activated class F fly
ash”, Cement and Concrete Research, Vol. 31, No. 9, pp. 1245-1249.
Yen, B.C. and Scanlon, B. (1975). “Sanitary landfill settlement rates”, Journal of
Geotechnical Engineering Division ASCE, Vol 101, No. 5, pp. 475-487.
Yesiller, N., Hanson, J.L., Rener, A.T. and Usmen, M.A. (2001). “Ultrasonic Testing for
evaluation of stabilized mixtures”, Geomaterials, Transportation Research Record,
1757, pp. 32-39.
Yusuf, F.A.M.S., Little D.N. and Sarkar S.L. (2001). “Evaluation of structural
contribution of lime stabilization of subgrade soils in Mississippi”, Geomaterials,
Transportation Research Record, 1757, pp. 22-31.
Zeccos, D. P. (2005). “Evaluation of Static and Dynamic Properties of Municipal Solid
Waste.” A dissertation submitted for the degree of Doctor of Philosophy In
Engineering - Civil and Environmental in the graduate division of the university of
California, Berkeley
Zekkos, D., Kabalan, M. and Flanagan, M. (2012). “Lessons Learned from Case Histories
of Dynamic Compaction at Municipal Solid Waste Sites”, Journal of Geotechnical
and Geoenvironmental Engineering.
Zhang, J.R. and Cao, X. (2002). “Stabilization of expansive soil by lime and fly ash”,
Journal of Wuhan University of Technology, Materials Science Edition, Vol. 17,
No. 4, pp. 73-77.
230
Appendix
231
232
Figure A.1 Vertical displacement from PLAXIS for untreated landfill 10 years after applying load
233
Figure A.2 Vertical displacement from PLAXIS for untreated landfill 20 years after applying load
234
Figure A.3 Vertical displacement from PLAXIS for 3-m treated with 6.7% fly ash-quicklime 10 years after applying load
235
Figure A.4 Vertical displacement from PLAXIS for 3-m treated with 6.7% fly ash-quicklime 20 years after applying load
236
Figure A.5 Vertical displacement from PLAXIS for 3-m treated with 13.3% fly ash-quicklime 10 years after applying load
237
Figure A.6 Vertical displacement from PLAXIS for 3-m treated with 13.3% fly ash-quicklime 20 years after applying load
238
Figure A.7 Vertical displacement from PLAXIS for 3-m treated with 20% fly ash-quicklime 10 years after applying load
239
Figure A.8 Vertical displacement from PLAXIS for 3-m treated with 20% fly ash-quicklime 20 years after applying load
240
Figure A.9 Vertical displacement from PLAXIS for 3-m treated with 26.6% fly ash-quicklime 10 years after applying load
241
Figure A.10 Vertical displacement from PLAXIS for 3-m treated with 26.6% fly ash-quicklime 20 years after applying load
242
Figure A.11 Vertical displacement from PLAXIS for 6-m treated with 6.6% fly ash-quicklime 10 years after applying load
243
Figure A.12 Vertical displacement from PLAXIS for 6-m treated with 6.6% fly ash-quicklime 20 years after applying load
244
Figure A.13 Vertical displacement from PLAXIS for 6-m treated with 13.3% fly ash-quicklime 10 years after applying load
245
Figure A.14 Vertical displacement from PLAXIS for 6-m treated with 13.3% fly ash-quicklime 20 years after applying load
246
Figure A.15 Vertical displacement from PLAXIS for 6-m treated with 20% fly ash-quicklime 10 years after applying load
247
Figure A.16 Vertical displacement from PLAXIS for 6-m treated with 20% fly ash-quicklime 20 years after applying load
248
Figure A.17 Vertical displacement from PLAXIS for 6-m treated with 26.6% fly ash-quicklime 10 years after applying load
249
Figure A.18 Vertical displacement from PLAXIS for 6-m treated with 26.6% fly ash-quicklime 20 years after applying load
250
Figure A.19 Vertical displacement from PLAXIS for 9-m treated with 6.6% fly ash-quicklime 10 years after applying load
251
Figure A.20 Vertical displacement from PLAXIS for 9-m treated with 6.6% fly ash-quicklime 20 years after applying load
252
Figure A.21 Vertical displacement from PLAXIS for 9-m treated with 13.3% fly ash-quicklime 10 years after applying load
253
Figure A.22 Vertical displacement from PLAXIS for 9-m treated with 13.3% fly ash-quicklime 20 years after applying load
254
Figure A.23 Vertical displacement from PLAXIS for 9-m treated with 20% fly ash-quicklime 10 years after applying load
255
Figure A.24 Vertical displacement from PLAXIS for 9-m treated with 20% fly ash-quicklime 20 years after applying load
256
Figure A.25 Vertical displacement from PLAXIS for 9-m treated with 26.6% fly ash-quicklime 10 years after applying load
257
Figure A.26 Vertical displacement from PLAXIS for 9-m treated with 26.6% fly ash-quicklime 20 years after applying load
258
Figure A.27 Vertical displacement from PLAXIS for 12-m treated with 6.6% fly ash-quicklime 10 years after applying load
259
Figure A.28 Vertical displacement from PLAXIS for 12-m treated with 6.6% fly ash-quicklime 20 years after applying load
260
Figure A.29 Vertical displacement from PLAXIS for 12-m treated with 13.3% fly ash-quicklime 10 years after applying load
261
Figure A.30 Vertical displacement from PLAXIS for 12-m treated with 13.3% fly ash-quicklime 20 years after applying load
262
Figure A.31 Vertical displacement from PLAXIS for 12-m treated with 20% fly ash-quicklime 10 years after applying load
263
Figure A.32 Vertical displacement from PLAXIS for 12-m treated with 20% fly ash-quicklime 20 years after applying load
264
Figure A.33 Vertical displacement from PLAXIS for 12-m treated with 26.6% fly ash-quicklime 10 years after applying load
265
Figure A.34 Vertical displacement from PLAXIS for 12-m treated with 26.6% fly ash-quicklime 20 years after applying load
266
Figure A.35 Vertical displacement from PLAXIS for 24-m treated with 6.6% fly ash-quicklime 10 years after applying
load
267
Figure A.36 Vertical displacement from PLAXIS for 24-m treated with 6.6% fly ash-quicklime 20 years after applying load
268
Figure A.37 Vertical displacement from PLAXIS for 24-m treated with 13.3% fly ash-quicklime 10 years after applying load
269
Figure A.38 Vertical displacement from PLAXIS for 24-m treated with 13.3% fly ash-quicklime 20 years after applying load
270
Figure A.39 Vertical displacement from PLAXIS for 24-m treated with 20% fly ash-quicklime 10 years after applying load
271
Figure A.40 Vertical displacement from PLAXIS for 24-m treated with 20% fly ash-quicklime 20 years after applying load
272
Figure A.41 Vertical displacement from PLAXIS for 24-m treated with 26.6% fly ash-quicklime 10 years after applying load
273
Figure A.42 Vertical displacement from PLAXIS for 24-m treated with 26.6% fly ash-quicklime 20 years after applying load