You are on page 1of 460

PURDUE UNIVERSITY

GRADUATE SCHOOL
Thesis Acceptance
This is to certify that the thesis prepared
By Joonho Hwang
Entitled
Effects of Cement Treatment on the 1-D Consolidation Behavior of a Highly Organic
Soil
Complies with University regulations and meets the standards of the Graduate School for originality
and quality

For the degree of

Doctor of Philosophy

Final examining committee members

Maria K. Santagata

Co

Antonio Bobet

Co-Chair

, Chair

Cliff T. Johnston

Darcy Bullock

Approved by Major Professor(s): Maria K. Santagata


Antonio Bobet

Approved by Head of Graduate Program:

Darcy Bullock

Date of Graduate Program Head's Approval: 12/15/2006

EFFECTS OF CEMENT TREATMENT ON THE 1-D CONSOLIDATION


BEHAVIOR OF A HIGHLY ORGANIC SOIL

A Dissertation
Submitted to the Faculty
of
Purdue University
by
Joonho Hwang

In Partial Fulfillment of the


Requirements for the Degree
of
Doctor of Philosophy

December 2006
Purdue University
West Lafayette, Indiana

UMI Number: 3304583

Copyright 2008 by
Hwang, Joonho
All rights reserved.

UMI Microform 3304583


Copyright 2008 by ProQuest Information and Learning Company.
All rights reserved. This microform edition is protected against
unauthorized copying under Title 17, United States Code.

ProQuest Information and Learning Company


300 North Zeeb Road
P.O. Box 1346
Ann Arbor, MI 48106-1346

ii

ACKNOWLEDGMENTS

First of all, I would like to thank God for giving me strength and guiding me
through the most important period of my life.
I would like to express my most sincere gratitude to my co-advisors, Professor
Antonio Bobet and Professor Maria K. Santagata for their guidance, invaluable
advising, encouragement, and patience they gave me throughout the course of
the research. This research could not have been accomplished without their
support and guidance.
I would also like to thank my advisory committee, especially Professor Darcy
Bullock in the School of Civil Engineering, Professor Cliff Johnston in the
Department of Agronomy, and Professor Arvid Johnson in the Department of
Earth and Atmospheric Sciences for their valuable suggestions and insight.
I offer special thanks to Professor Darrell G. Schulze in the Department of
Agronomy for his assistance and guidance with the XRD analysis. I thank Ms.
Shelley Finnigan and Dr. G.S. Premachandra in the Department of Agronomy for
their assistance with the fractionation and FTIR spectroscopy analysis. I
especially thank Ms. Janet Lovell, our lab manager, for all her support and help
with the experimental work. I would like to extend special thanks and gratitude to
my office mates and friends: especially Yeoun-Ike Kang for his great
companionship, Yiannis Zevgolis for always being there for me, Dong Wook Kim,
Chi Hyun Park, and Ho Young Seo for their time and support.
I express my sincere gratitude to my parents and parent-in-laws for their prayer,
encouragement and support.
Most of all, I would like to thank my wife for her love, support, and sacrifice,
without whom this would never have been possible.

iii

TABLE OF CONTENTS

Page
LIST OF TABLES .................................................................................................vi
LIST OF FIGURES ...............................................................................................ix
LIST OF SYMBOLS.......................................................................................... xviii
ABSTRACT ........................................................................................................xxi
CHAPTER 1. INTRODUCTION ............................................................................ 1
1.1. Problem statement ..................................................................................... 1
1.2. Scope of work and research approach ....................................................... 4
1.3. Organization of the thesis ........................................................................... 7
CHAPTER 2. BACKGROUND............................................................................ 10
2.1. Types of organic soils ............................................................................... 12
2.2. Characteristics of soil organic matter........................................................ 15
2.3. Soil - organic matter interaction ................................................................ 21
2.4. Effects of organic matter on engineering properties of soils ..................... 24
2.5. 1-D consolidation behavior of organic soils .............................................. 27
2.6. Deep mixing methods ............................................................................... 31
2.7. Organic soil and cement reactions ........................................................... 35
2.8. 1-D consolidation behavior of chemically treated organic soils................. 43
CHAPTER 3. LINDBERG ROAD SOIL............................................................... 85
3.1. Introduction............................................................................................... 85
3.2. Characterization and classification of LR soil ........................................... 87
3.2.1. Water content ..................................................................................... 87
3.2.2. Organic content .................................................................................. 88
3.2.3. Specific gravity ................................................................................... 88
3.2.4. Particle size distribution ...................................................................... 89
3.2.5. Atterberg limits.................................................................................... 90
3.2.6. Fiber content....................................................................................... 92
3.2.7. Soil acidity .......................................................................................... 92
3.2.8. Soil taxonomy classification................................................................ 93
3.3. Mineral composition of LR soil.................................................................. 95
3.3.1. Fractionation of LR soil ....................................................................... 96
3.3.2. Mineral composition of LR soil.......................................................... 102
3.4. Characterization of the organic matter in LR soil .................................... 105
3.4.1. Extraction and fractionation of the humic substances in LR soil ....... 106
3.4.2. Elemental composition of the organic matter in LR soil .................... 109
3.4.3. FT-IR analysis of the humic substances of LR soil ........................... 111

iv
Page
CHAPTER 4. EXPERIMENTAL METHODS ..................................................... 141
4.1. Introduction............................................................................................. 141
4.2. Experimental Program ............................................................................ 141
4.3. Laboratory sample preparation procedure.............................................. 142
4.3.1. Mixing ............................................................................................... 142
4.3.2. Compaction ...................................................................................... 144
4.3.3. Curing ............................................................................................... 144
4.3.4. Extrusion........................................................................................... 145
4.4. CRS consolidation test ........................................................................... 145
4.4.1. General overview.............................................................................. 145
4.4.2. Procedure for CRS consolidation test............................................... 150
4.4.3. CRS apparatus calibration................................................................ 152
4.4.4. Data reduction .................................................................................. 154
4.4.5. CRS consolidation theory ................................................................. 155
CHAPTER 5. 1-D CONSOLIDATION BEHAVIOR OF LINDBERG ROAD
SOIL .......................................................................................... 179
5.1. Introduction............................................................................................. 179
5.2. 1-D consolidation behavior of reconstituted LR soil................................ 180
5.2.1. Compression behavior and generation of excess pore pressure...... 181
5.2.2. Constrained Modulus........................................................................ 188
5.2.3. Hydraulic conductivity (k).................................................................. 190
5.2.4. Coefficient of Consolidation (Cv)....................................................... 191
5.2.5. Repeatability..................................................................................... 193
5.3. Strain rate selection and rate effects ...................................................... 194
5.3.1. Strain rate effects on compression behavior..................................... 196
5.3.2. Strain rate effect on excess pore pressure generation ..................... 197
5.3.3. Strain rate effect on hydraulic conductivity and coefficient of
consolidation............................................................................................... 199
5.4. 1-D consolidation behavior of intact LR soil............................................ 201
5.4.1. Compression behavior of intact LR soil ............................................ 203
5.4.2. Hydraulic conductivity and coefficient of consolidation of intact
LR soil.............................................................................................. 210
5.5. Creep behavior of reconstituted LR soil.................................................. 213
5.5.1. Determination of the End-Of-Primary points ..................................... 215
5.5.2. Comparison of the results from IL and CRS consolidation tests....... 216
5.5.3. Evaluation of C and C/Cc of reconstituted LR soil ......................... 218
5.5.4. Evaluation of C and C/Cc of intact LR soil ..................................... 221
CHAPTER 6. 1-D CONSOLIDATION BEHAVIOR OF PC TREATED LR
SOIL .......................................................................................... 269
6.1. Introduction............................................................................................. 269
6.2. 1-D consolidation behavior of PC treated LR soil ................................... 270
6.2.1. Calculation of the initial void ratio ..................................................... 272
6.2.2. 1-D Consolidation behavior of 8.0% PC treated LR soil ................... 275
6.2.3. 1-D consolidation behavior of 18.7% PC treated LR soil .................. 280

v
Page
6.2.4. 1-D Consolidation behavior of 51.4% PC treated LR soil ................. 283
6.2.5. 1-D Consolidation behavior of 100% PC treated LR soil .................. 290
6.2.6. Change of pH of LR soil with PC treatment ...................................... 295
6.2.7. Destructuring behavior of PC treated LR soil.................................... 296
6.2.8. Discussion: 1-D consolidation behavior of PC treated LR soil .......... 298
6.3. Effects of curing...................................................................................... 305
6.3.1. Effects of curing time ........................................................................ 305
6.3.2. Effects of curing surcharge ............................................................... 311
6.3.3. Discussion: Effects of curing on the 1-D consolidation behavior
of PC treated LR soil........................................................................ 314
6.4. Framework: Effects of treatment on the 1-D consolidation behavior
of soils ................................................................................................... 315
6.5. Effects of treatment on the organic matter of LR soil.............................. 320
CHAPTER 7. CONCLUSIONS AND RECOMMENDATIONS........................... 388
7.1. Introduction............................................................................................. 388
7.2. Results and Conclusions ........................................................................ 388
7.2.1. Sampling and characterization of LR soil.......................................... 388
7.2.2. Review of the 1-D Consolidation Behavior of Intact and
Reconstituted LR soil....................................................................... 390
7.2.3. Conclusions on the 1-D Consolidation Behavior of LR soil............... 394
7.2.4. Review of the effects of treatment on the 1-D consolidation
behavior of LR soil ........................................................................... 397
7.2.5. Review of organic soil / cement interactions..................................... 403
7.2.6. Conclusions on the effects of cement treatment on the 1-D
consolidation behavior, the structure and the chemistry of a
highly organic soil ............................................................................ 405
7.3. Recommendations for Future Research ................................................. 410
LIST OF REFERENCES .................................................................................. 415
VITA ................................................................................................................. 433

vi

LIST OF TABLES

Table
Page
Table 2.1: Major characteristics of 12 soil orders in Soil Taxonomy (after
Brady and Weil, 1999) ..................................................................... 55
Table 2.2: Distribution of soil organic matter in Florida Histosols (after
Zelazny and Carlisle, 1974) ............................................................. 55
Table 2.3: Elemental composition of humic acids extracted from soils from
widely different climates (after Schnitzer and Khan, 1978) .............. 56
Table 2.4: Elemental composition of fulvic acids extracted from soils from
widely different climates (after Schnitzer and Khan, 1978) .............. 56
Table 2.5: Elemental composition of Humic Acids (HA) and Fulvic Acids
(FA) from various tropical and temperate regions soils (after
Tan, 2003) ....................................................................................... 57
Table 2.6: Elemental composition of humic substances of Florida Histosols
(after Zelazny and Carlisle, 1974).................................................... 58
Table 2.7: Distribution of oxygen-containing functional groups in humic
and fulvic acids extracted from soils of widely different climatic
zones (all numbers in meq/100g) (after Stevenson, 1994) .............. 59
Table 2.8: Concentration of organic functional groups in humic substances
from 11 Florida Histosols samples (average and standard
deviation values expressed in cmol/kg) (after Sparks, 2003)........... 59
Table 2.9: Relative values of various peat properties for amorphousgranular, fine-fibrous and coarse-granular peat (after
MacFarlane and Williams, 1974) ..................................................... 60
Table 2.10: Values of C/Cc for various types of geotechnical materials
(Terzaghi et al, 1996)....................................................................... 60
Table 2.11: Values of natural water content (wo), permeability, and C/Cc
for various peat deposits (after Mesri et al, 1997)............................ 61
Table 2.12: Composition of Ordinary Type I Portland cement
manufactured in the States (Lea, 1956)........................................... 61
Table 2.13: Index properties of Saguenay Fjord sediments, Port de
Qubec recent sediments, and dredged material from Qubec
harbor (Tremblay et al, 2001) .......................................................... 62
Table 3.1: Summary of index properties and unit weights of peats and
organic soils (after Edil and Wang, 2000) ...................................... 116
Table 3.2: Summary of the index properties of LR soil ..................................... 117

vii
Table
Page
Table 3.3: Particle size distribution of the inorganic component of LR soil
(LR-B9) .......................................................................................... 117
Table 3.4: Mineral composition of each of three size fractions of LR soil ......... 118
Table 3.5: Mass of the three organic matter fractions of LR soil....................... 118
Table 3.6: Elemental composition of each fraction of LR soil ........................... 118
Table 3.7: Elemental compositions of the humic substances from soils and
other sources (after Tan, 2003) ..................................................... 119
Table 3.8: Elemental composition of humic substances of Florida Histosols
(after Zelazny and Carlisle, 1974).................................................. 120
Table 3.9: Characteristic infrared adsorption frequencies of various
organic functional groups (after Stevenson, 1982) ........................ 120
Table 3.10: Organic functional groups identified with FT-IR analysis in the
three fractions of the organic matter of LR soil .............................. 121
Table 4.1: Ranges of the testing variables investigated ................................... 162
Table 4.2: Resolution and stability of the sensors used in CRS1 system ......... 162
Table 4.3: Resolution and stability of the sensors used in CRS2 system ......... 163
Table 4.4: Resolution and stability of the sensors used in CRS3 system ......... 163
Table 4.5: Coefficients for CRS system calibration........................................... 164
Table 5.1: Summary of reconstituted LR soil specimen data............................ 225
Table 5.2: Summary of testing process of CRS tests performed on
reconstituted LR soil ...................................................................... 225
Table 5.3: Summary of results of the CRS consolidation test performed on
reconstituted LR soil ...................................................................... 226
Table 5.4: Summary of intact LR soil specimen data........................................ 226
Table 5.5: Summary of testing process of CRS tests performed on intact
LR soil............................................................................................ 226
Table 5.6: Results of the CRS consolidation tests performed on intact LR
soil ................................................................................................. 227
Table 5.7: Coordinates of sedimentation compression line (SCL) (Burland,
1990) ............................................................................................. 227
Table 5.8: Summary of index properties of reconstituted LR soil specimen
data for IL consolidation tests ........................................................ 228
Table 5.9: Summary of data from 24h-IL consolidation test performed on
reconstituted April LR soil (IL025).................................................. 228
Table 5.10: Summary of data from EOP-IL consolidation test performed on
reconstituted July LR soil (IL037)................................................... 229
Table 5.11: Summary of data from EOP-IL consolidation test performed on
block LR soil (IL056) ...................................................................... 229
Table 5.12: Summary of data from EOP-IL consolidation test performed on
block LR soil (IL071) ...................................................................... 230
Table 6.1: Comparison of %PC vs. Cement factor ........................................... 329
Table 6.2: Summary of the PC treated LR soil specimens ............................... 329
Table 6.3: Test process of CRS consolidation tests on PC treated LR soil ...... 330
Table 6.4: Comparison of the back calculated LOI and the actual LOI............. 330

viii
Table
Page
Table 6.5: Summary of the results of consolidation tests performed on PC
treated LR soil specimens.............................................................. 331
Table 6.6: Summary of IL consolidation test performed on 8.00% PC
treated LR soil (IL048) ................................................................... 331
Table 6.7: Summary of IL consolidation test performed on 18.7% PC
treated LR soil (IL 042) .................................................................. 332
Table 6.8: Summary of IL consolidation test performed on 51.4% PC
treated LR soil (IL 050) .................................................................. 332
Table 6.9: Summary of IL consolidation test performed on 103.4% PC
treated LR soil (IL 054) .................................................................. 333
Table 6.10: LR soil specimens employed for evaluation of curing time
effects ............................................................................................ 333
Table 6.11: LR soil specimens employed for evaluation of the effects of
curing surcharge ............................................................................ 334
Table 6.12: Mass and organic C, H, O contents of each fraction of the
humic substances extracted from PC treated LR soil (LR-B9)....... 334
Table 6.13: Mass of the organic carbon in each fraction of LR soil .................. 335
Table 6.14: Mass of H and N in each fraction of LR soil................................... 335

ix

LIST OF FIGURES

Figure
Page
Figure 2.1: General degree of weathering, climatic and vegetative
conditions of 12 soil orders in Soil Taxonomy (after Brady and
Weil, 1999)....................................................................................... 63
Figure 2.2: Classification of soil organic matter (humus) .................................... 64
Figure 2.3: Initial water content (wo) and compression index (Cc) of peats
and soft clay and silt deposits (after Mesri et al, 1997) .................... 64
Figure 2.4: Comparison of compression curves and curves of hydraulic
conductivity versus void ratio of fibrous and amorphous
granular peat (after Berry and Poskitt, 1972) ................................... 65
Figure 2.5: Range in compression behavior of intact Middleton peats (after
Mesri et al, 1997) ............................................................................. 66
Figure 2.6: Hydraulic conductivity of Middleton peat (after Mesri et al,
1997) ............................................................................................... 67
Figure 2.7: Comparison of Ck for peat and soft clay and silt deposits (after
Mesri et al, 1997) ............................................................................. 68
Figure 2.8: C/Cc for Middleton peat (after Mesri et al, 1997) ............................. 69
Figure 2.9: Compression curves and curves of hydraulic conductivity
versus void ratio for fibrous peats from northern Quebec,
Canada (after Lefebvre et al, 1984) ................................................. 70
Figure 2.10: C/Cc for fibrous peats from northern Quebec, Canada (after
Lefebvre et al, 1984)........................................................................ 71
Figure 2.11: FHWA classification of Deep Mixing Methods (after Bruce
and Bruce, 2003) ............................................................................. 72
Figure 2.12: Schematic diagram of inorganic clay - cement reactions (after
Saitoh et al., 1985)........................................................................... 73
Figure 2.13: Increase in particle size due to pozzolanic reaction (after Rao
and Rajasekaran, 1996) .................................................................. 74
Figure 2.14: Effect of cement treatment on pore size distribution of
Singapore marine clay (after Chew et al, 2004)............................... 75
Figure 2.15: Compression curves of dredged materials untreated and
treated with 10% Portland cement (after Tremblay et al, 2001) ....... 76
Figure 2.16: Average normalized compression curves of: (a) dredged
material treated with 10% Portland cement, and (2) organic
soils treated with 5% Portland cement (after Tremblay et al,
2001) ............................................................................................... 77

x
Figure
Page
Figure 2.17: Schematic diagram showing the effect of organic matter
content on the position of the normalized compression curves
(after Tremblay et al, 2001).............................................................. 78
Figure 2.18: Compression curves of (a) untreated Raheenmore peat (RA)
and Ballydermot peat (BA), and (b) Ballydermot peat (BA)
treated with 200 kg/m3 of cement (after Hebib and Farrell,
2003) ............................................................................................... 79
Figure 2.19: Compression curves of untreated and cement treated (a)
Adria and (b) Correzzola peats (after Cortellazzo and Cola,
1999) ............................................................................................... 80
Figure 2.20: Effects of treatment on the compressibility of cement treated
(a) Adria and (b) Correzzola peats (after Cortellazzo and Cola,
1999) ............................................................................................... 81
Figure 2.21: Effects of treatment on the coefficient of consolidation of (a)
Adria and (b) Correzzola peats treated with 200kg/m3 of
cement (after Cortellazzo and Cola, 1999) ...................................... 82
Figure 2.22: Ranges in hydraulic conductivity measured on soils treated
with lime and cement (after hnberg, 2003).................................... 83
Figure 2.23: Relationship between the change in permeability and change
in water content of stabilized soils (after hnberg, 2003) ................ 84
Figure 3.1: The 30 cm x 30 cm x 30 cm block sampler made of Lexan
plates ............................................................................................. 122
Figure 3.2: Intact block sample sealed with plastic wrap and wax.................... 122
Figure 3.3: Linear relationship between the specific gravity and the loss on
ignition of LR soil ........................................................................... 123
Figure 3.4: Relationship between specific gravity and organic content
(after MacFarlane, 1969) ............................................................... 123
Figure 3.5: Particle size distribution of LR soil (LR-B9) .................................... 124
Figure 3.6: Influence of organic content on the Atterberg limits of LR soil........ 124
Figure 3.7: Influence of organic content on the Plasticity index of (a) LR
soil and (b) Juturnaiba organic clay from Brazil (after Mitchell,
1996) ............................................................................................. 125
Figure 3.8: Atterberg limits of LR soil on the plasticity chart ............................. 125
Figure 3.9: Comparison of the Atterberg limits of LR soil with other
amorphous and fibrous peats (after MacFarlane, 1969) ................ 126
Figure 3.10: Acidity (pH) vs. organic content (after MacFarlane, 1969)............ 127
Figure 3.11: Fractionation of particles using centrifuging method..................... 127
Figure 3.12: Particle size distribution of the inorganic component of LR soil
(LR-B9) .......................................................................................... 128
Figure 3.13: Fractionation procedure for XRD specimen.................................. 129
Figure 3.14: XRD patterns of the inorganic components of LR soil particles
in 5 2 m size fraction: from the bottom (a) Mg-glycerated,
(b) K-saturated at 21 C, (c) K saturated and heated at 300 C,
and (d) K-saturated and heat treated at 550 C ............................. 130

xi
Figure
Page
Figure 3.15: XRD patterns of the inorganic components of LR soil particles
in 2 0.2 m size fraction: from the bottom (a) Mg-glycerated,
(b) K-saturated at 21 C, (c) K saturated and heated at 300 C,
and (d) K-saturated and heat treated at 550 C ............................. 131
Figure 3.16: XRD patterns of the inorganic components of LR soil
particles smaller than 0.2 m: from the bottom (a) Mgglycerated, (b) K-saturated at 21 C, (c) K saturated and
heated at 300 C, and (d) K-saturated and heat treated at 550
C................................................................................................... 132
Figure 3.17: Organic matter extraction procedure ............................................ 133
Figure 3.18: Three vibration modes: bending (v2), symmetrical (v3) and
asymmetrical stretching (v1) vibration of water molecules (after
Johnston, 1996) ............................................................................. 134
Figure 3.19: FT-IR spectrum of the humin and inorganic component of LR
soil ................................................................................................. 135
Figure 3.20: FT-IR spectrum of humic acid of LR soil....................................... 136
Figure 3.21: Comparison of the FT-IR spectra of humic acid from a
Suwannee River humic acid (after Niemeyer et al., 1992), a
peat standard (H4) (after Niemeyer et al., 1992), and LR soil ........ 137
Figure 3.22: IR spectrum of the humic acids from four different sources
(after Tan, 2003), and LR soil ........................................................ 138
Figure 3.23: FT-IR spectrum of fulvic acid of LR soil ........................................ 139
Figure 3.24: IR spectrum of fulvic acids from four different sources (after
Tan, 2003) and LR soil .................................................................. 140
Figure 4.1: Mixer with two different mixing tools employed to prepare
samples for CRS consolidation tests. ............................................ 165
Figure 4.2: A plastic cylinder and modified compaction mold employed for
compaction. ................................................................................... 165
Figure 4.3: Modified mechanical proctor compactor employed for treated
LR soil samples: (1) original and (2) modified compaction
hammer.......................................................................................... 166
Figure 4.4: Cutting of plastic cylinder with Wells metal band saw for
preparation of soil samples for CRS consolidation tests ................ 166
Figure 4.5: Curing of soil specimen in water bath under a surcharge of 48
kPa ................................................................................................ 167
Figure 4.6: Schematic of the four components of a CRS consolidation test
apparatus; CRS Cell, Load Frame, Air-Water Interface
Pressure Regulator and Data Acquisition System (DAS)
(connecting wires behind PC, DAS, and load frame not shown).... 168
Figure 4.7: Specimen trimming apparatus: trimming frame, a thin spatula,
cutting blade and a wire saw.......................................................... 169
Figure 4.8: Schematic of CRS consolidation cell (CRS1) ................................. 170
Figure 4.9: Resolution of the Data Acquisition System and sensors
(CRS1)........................................................................................... 171

xii
Figure
Page
Figure 4.10: Stability of the sensors ................................................................. 172
Figure 4.11: Uplift calibration for CRS cell (The area subjected to cell or
backpressure is indicated with a thick line) .................................... 173
Figure 4.12: Calibration for uplift force.............................................................. 174
Figure 4.13: Calibration for machine deflection (compliance) ........................... 174
Figure 4.14: Flow chart for data reduction ........................................................ 175
Figure 4.15: Plot of the function F3 as a function of Tv ...................................... 176
Figure 4.16: Comparison of vertical effective stress computed using linear
and non-linear CRS theory (a) 0<uh/v<1.0 and (b) 0< uh/v
<0.5................................................................................................ 177
Figure 4.17: Comparison of coefficient of consolidation computed using
linear and non-linear CRS theory (a) 0<uh/v<1.0 and (b) 0<
uh/v <0.5 ....................................................................................... 178
Figure 5.1: Compression curves of reconstituted LR soil (CRS007) a) in
log scale and b) in arithmetic scale: A-B loading, B-C creep, CD unloading, D-E creep, and E-F reloading ................................... 231
Figure 5.2: Estimation of preconsolidation pressure (p) using the strain
energy method............................................................................... 232
Figure 5.3: Estimation of preconsolidation pressure (p) using
Casagrandes graphical construction............................................. 232
Figure 5.4: Generation of excess pore pressure during loading ....................... 233
Figure 5.5: Increase of actual strain rate with decrease in specimen height .... 233
Figure 5.6: Change of pore pressure ratio during loading ................................ 234
Figure 5.7: Time-settlement and dissipation of excess pore pressure
during creep after loading (CRS007). ............................................ 234
Figure 5.8: Time-settlement and dissipation of excess pore pressure
during creep after unloading (CRS007). ........................................ 235
Figure 5.9: Generation of excess pore pressure during reloading .................... 235
Figure 5.10: Comparison of actual strain rate during loading and reloading..... 236
Figure 5.11: Change in pore pressure ratio during reloading ........................... 236
Figure 5.12: Change in constrained modulus (D) during loading, unloading
and reloading ................................................................................. 237
Figure 5.13: Stress-strain curve in the over consolidated region ...................... 237
Figure 5.14: Change in hydraulic conductivity with void ratio ........................... 238
Figure 5.15: Change in hydraulic conductivity with vertical effective stress...... 238
Figure 5.16: Change in coefficient of consolidation with vertical effective
stress ............................................................................................. 239
Figure 5.17: Comparison of compression curves ............................................. 239
Figure 5.18: Comparison of excess pore pressure generation ......................... 240
Figure 5.19: Comparison of constrained modulus ............................................ 240
Figure 5.20: Comparison of hydraulic conductivity ........................................... 241
Figure 5.21: Comparison of coefficient of consolidation ................................... 241
Figure 5.22: Effect of strain rate on compression curve ................................... 242

xiii
Figure
Page
Figure 5.23: Effect of strain rate on generation of excess pore pressure for
a) April and b) July specimens....................................................... 242
Figure 5.24: Excess pore pressure normalized by actual strain rate ................ 243
Figure 5.25: Effect of strain rate on pore pressure ratio ................................... 243
Figure 5.26: Effect of strain rate on hydraulic conductivity ............................... 244
Figure 5.27: Effect of strain rate on coefficient of consolidation of April soil ..... 244
Figure 5.28: Effect of strain rate on coefficient of consolidation of July soil...... 245
Figure 5.29: Compression curves of intact LR soil ........................................... 245
Figure 5.30: Comparison of the compression curve of intact LR soil with a
peat (Middleton peat, Mesri et al, 1997) and an inorganic clay
(RBBC, Force, 1998) ..................................................................... 246
Figure 5.31: Comparison of the compression index of intact LR soil with
other soils (after Mesri and Rokhsar, 1974) ................................... 247
Figure 5.32: Comparison of the compression index ......................................... 248
Figure 5.33: The intrinsic compression line in (a) e-logv and (b) Void
index (Iv)- logv (after Burland, 1990) ........................................... 248
Figure 5.34: Intrinsic and sedimentation compression lines normalized
with void index (Burland, 1990)...................................................... 249
Figure 5.35: Sedimentation compression curves of intact LR soil .................... 250
Figure 5.36: Idealization of the compression behavior of reconstituted and
structured soils (after Liu and Carter, 2002)................................... 250
Figure 5.37: Influence of the destructuring index b during compression
(afterLiu and Carter, 2000)............................................................. 251
Figure 5.38: Destructuring behavior of (a) stiff Vallericca clay and (b) four
stiff clays and (c) an artificially bonded soft clay and a soft
calcerenite soil during compression (after Liu and Carter,
2000) ............................................................................................. 252
Figure 5.39: Destructuring of intact LR soil....................................................... 253
Figure 5.40: Constrained modulus of intact LR soil .......................................... 253
Figure 5.41: Constrained modulus against normalized vertical effective
stress ............................................................................................. 254
Figure 5.42: Generation of excess pore pressure in intact LR soil
specimens...................................................................................... 254
Figure 5.43: Hydraulic conductivity of intact LR soil ......................................... 255
Figure 5.44: Comparison of the hydraulic conductivity of LR soil to that of
RBBC an Middleton peat ............................................................... 255
Figure 5.45: Coefficient of consolidation of intact LR soil ................................. 256
Figure 5.46: Coefficient of consolidation against normalized vertical
effective stress............................................................................... 256
Figure 5.47: Comparison of the coefficients of consolidation of intact LR
soil and an inorganic clay............................................................... 257
Figure 5.48: Determination of the EOP with Casagrandes procedure
(IL025) ........................................................................................... 257

xiv
Figure
Page
Figure 5.49: Determination of the EOP based on the excess pore
pressures measurement (IL037).................................................... 258
Figure 5.50: Compression curves of reconstituted LR soil obtained from
CRS and IL consolidation tests...................................................... 258
Figure 5.51: Hydraulic conductivity of reconstituted LR soil obtained from
CRS and IL consolidation tests...................................................... 259
Figure 5.52: Coefficient of consolidation of reconstituted LR soil obtained
from CRS and IL consolidation tests.............................................. 259
Figure 5.53: Long term creep test for determination of secondary
compression index (C) of reconstituted LR soil (IL037)................ 260
Figure 5.54: Comparison of the secondary compression index of
reconstituted LR soil at different vertical effective stress levels ..... 261
Figure 5.55: EOP consolidation curve from IL025 ............................................ 262
Figure 5.56: Estimation of Cc associated with secondary consolidation for
determination of C/Cc: (a) compression curve and (b) time
settlement curve at v = 69.9 kPa (IL025)..................................... 263
Figure 5.57: Calculation of C/Cc for reconstituted April LR soil (IL025) ........... 264
Figure 5.58: Calculation of C/Cc for reconstituted July LR soil (IL037)............ 264
Figure 5.59: EOP compression curves of two block LR soil specimens ........... 265
Figure 5.60: Constrained modulus of two block LR soil specimens.................. 265
Figure 5.61: Hydraulic conductivity of two block LR soil specimens ................. 266
Figure 5.62: Coefficient of consolidation of two block LR soil specimens......... 266
Figure 5.63: Long term creep test for determination of secondary
compression index (C) of Block LR soil (IL056) ........................... 267
Figure 5.64: Long term creep test for determination of secondary
compression index (C) of Block LR soil (IL071) ........................... 268
Figure 5.65: C/Cc of Block LR soil................................................................... 268
Figure 6.1: Four phases of PC treated LR soil (not drawn to scale) ................. 336
Figure 6.2: The initial void ratio of PC treated LR soil....................................... 336
Figure 6.3: Compression curves of LR soil treated with 8.00%PC ................... 337
Figure 6.4: Normalized compression curves of LR soil treated with 8.00%
PC.................................................................................................. 337
Figure 6.5: Constrained modulus of LR soil treated with 8.00% PC ................. 338
Figure 6.6: Cumulative excess pore pressure generated in CRS test on LR
soil treated with 8.00% PC............................................................. 338
Figure 6.7: Hydraulic conductivity of LR soil treated with 8.00% PC ................ 339
Figure 6.8: Coefficient of consolidation of LR soil treated with 8.00% PC ........ 339
Figure 6.9: Time settlement and excess pore pressures dissipation curves
from long term creep test performed on LR soil treated with
8.00% PC....................................................................................... 340
Figure 6.10: Compression curves of LR soil treated with 18.7% PC ................ 340
Figure 6.11: Normalized compression curves of LR soil treated with 18.7%
PC.................................................................................................. 341
Figure 6.12: Constrained modulus of LR soil treated with 18.7% PC ............... 341

xv
Figure
Page
Figure 6.13: Excess pore pressure generated in CRS test on LR soil
treated with 18.7% PC ................................................................... 342
Figure 6.14: Hydraulic conductivity of LR soil treated with 18.7% PC .............. 342
Figure 6.15: Coefficient of consolidation of LR soil treated with 18.7% PC ...... 343
Figure 6.16: Time settlement and excess pore pressures dissipation
curves from a creep test performed on LR soil treated with
18.7% PC....................................................................................... 343
Figure 6.17: Compression curves of LR soil treated with 51.4% PC ................ 344
Figure 6.18: Normalized compression curves of LR soil treated with 51.4%
PC.................................................................................................. 344
Figure 6.19: Development of preconsolidation pressure during creep in LR
soil treated with 51.4% PC............................................................. 345
Figure 6.20: Time elapsed since treatment with PC at each stage of CRS
consolidation test ........................................................................... 346
Figure 6.21: Constrained modulus of LR soil treated with 51.4% PC ............... 346
Figure 6.22: Excess pore pressures generated in CRS test on LR soil
treated with 51.4% PC ................................................................... 347
Figure 6.23: Hydraulic conductivity of LR soil treated with 51.4% PC .............. 347
Figure 6.24: Coefficient of consolidation of LR soil treated with 51.4% PC ...... 348
Figure 6.25: Time settlement and excess pore pressure dissipation curves
from a creep test performed on LR soil treated with 51.4% PC ..... 348
Figure 6.26: Consolidation curves of LR soil treated with 49.1% PC................ 349
Figure 6.27: Increase in the preconsolidation pressure after 7 day creep at
1000 kPa........................................................................................ 349
Figure 6.28: Increase in the preconsolidation pressure after 7 day creep at
257 kPa.......................................................................................... 350
Figure 6.29: Increase in the preconsolidation pressure after 7 day creep at
850kPa........................................................................................... 351
Figure 6.30: Increase in the preconsolidation pressure after 7 day creep at
2450 kPa........................................................................................ 352
Figure 6.31: Compression curves of LR soil treated with 100% PC ................. 352
Figure 6.32: Development of preconsolidation pressure during creep in LR
soil treated with 103.4 % PC.......................................................... 353
Figure 6.33: Development of preconsolidation pressure during creep in LR
soil treated with 100.8 % PC.......................................................... 354
Figure 6.34: Normalized compression curves of LR soil treated with 100%
PC.................................................................................................. 354
Figure 6.35: Constrained modulus of LR soil treated with 100% PC ................ 355
Figure 6.36: Cumulative excess pore pressure generated in LR soil
treated with 100% PC .................................................................... 355
Figure 6.37: Hydraulic conductivity of LR soil treated with 100% PC ............... 356
Figure 6.38: Coefficient of consolidation of LR soil treated with 100% PC ....... 356

xvi
Figure
Page
Figure 6.39: Time settlement and excess pore pressures dissipation
curves from a creep test performed on LR soil treated with
1003.4% PC................................................................................... 357
Figure 6.40: Change in pH of LR soil with treatment ........................................ 357
Figure 6.41: Quantification of post-yield destructuring of LR soil treated
with 8.00% PC ............................................................................... 358
Figure 6.42: Quantification of post-yield destructuring of LR soil treated
with 18.7% PC ............................................................................... 359
Figure 6.43: Quantification of post-yield destructuring of LR soil treated
with 51.4% PC: .............................................................................. 360
Figure 6.44: Quantification of post-yield destructuring of LR soil treated
with 103.4% PC ............................................................................. 361
Figure 6.45: Quantification of post-yield destructuring of LR soil treated
with 100.8% PC ............................................................................. 362
Figure 6.46: Compression curves and destructuring indices of lime treated
Louiseville Clay (after Liu and Carter, 2000).................................. 363
Figure 6.47: Effect of treatment on compression behavior of LR soil ............... 364
Figure 6.48: Compression curves of lime stabilized Louiseville clay (after
Locat et al, 1996) ........................................................................... 365
Figure 6.49: Compression curves of cement treated soft Bangkok clay: (a)
5%, (b) 10%, and (c) 15% cement contents (after Lorenzo and
Bergado, 2004) .............................................................................. 366
Figure 6.50: Increase in the preconsolidation pressure with treatment............. 367
Figure 6.51: Development of structure with treatment ...................................... 368
Figure 6.52: Increase in the constrained modulus with treatment..................... 368
Figure 6.53: Constrained modulus vs. normalized vertical effective stress ...... 369
Figure 6.54: Increase in hydraulic conductivity with treatment ......................... 369
Figure 6.55: Increase in coefficient of consolidation with treatment.................. 370
Figure 6.56: Coefficient of consolidation vs. normalized vertical effective
stress ............................................................................................. 370
Figure 6.57: Effect of treatment on the C/Cc of LR soil ................................... 371
Figure 6.58: Effect of curing time on the compression curve of LR soil
treated with 50% PC ...................................................................... 371
Figure 6.59: Effect of curing time on the development of structure in LR
soil treated with 50% PC................................................................ 372
Figure 6.60: Effect of curing time on the constrained modulus of LR soil
treated with 50% PC ...................................................................... 372
Figure 6.61: Effect of curing time on hydraulic conductivity of LR soil
treated with 50% PC ...................................................................... 373
Figure 6.62: Effect of curing time on the coefficient of consolidation of LR
soil treated with 50% PC................................................................ 373
Figure 6.63: Effect of curing time on creep behavior of LR soil treated with
50% PC.......................................................................................... 374

xvii
Figure
Page
Figure 6.64: Effect of curing time the on compression curve of LR soil
treated with 100.8% PC ................................................................. 374
Figure 6.65: Effect of curing time on the development of structure in LR
soil treated with 100.8% PC........................................................... 375
Figure 6.66: Effect of curing time on the constrained modulus of LR soil
treated with 100.8%PC .................................................................. 375
Figure 6.67: Effect of curing time on the hydraulic conductivity of LR soil
treated with 100.8% PC ................................................................. 376
Figure 6.68: Effect of curing time of the coefficient of consolidation of LR
soil treated with 100.8% PC........................................................... 376
Figure 6.69: Effect of curing time on creep behavior of LR soil treated with
100.8% PC..................................................................................... 377
Figure 6.70: Effect of curing surcharge on compression behavior of LR soil
treated with 50% PC ...................................................................... 377
Figure 6.71: Increase in preconsolidation pressure with increasing curing
surcharge....................................................................................... 378
Figure 6.72: Effect of curing surcharge on constrained modulus of LR soil
treated with 50% PC ...................................................................... 378
Figure 6.73: Destructuring of 52.6%PC treated LR soil cured under 96 kPa.... 379
Figure 6.74: Destructuring of 48.4%PC treated LR soil cured under 192
kPa ................................................................................................ 379
Figure 6.75: Effect of curing surcharge on the hydraulic conductivity............... 380
Figure 6.76: Effect of curing surcharge on the coefficient of consolidation....... 380
Figure 6.77: Schematic diagram of the effect of cement treatment on the
position of the virgin compression curve ........................................ 381
Figure 6.78: Schematic diagram of the effect of cement treatment on the
preconsolidation pressure.............................................................. 381
Figure 6.79: Schematic diagram of the effect of cement treatment on the
constrained modulus...................................................................... 382
Figure 6.80: Hydraulic conductivity of lime stabilized Louiseville clay
(Locat et al, 1996).......................................................................... 383
Figure 6.81: Schematic diagram of the effect of cement treatment on the
coefficient of consolidation............................................................. 384
Figure 6.82: Changes in FT-IR spectra of non-extracted fraction of the PC
treated LR soil................................................................................ 385
Figure 6.83: Changes in FT-IR spectra of humic acid extracted from the
PC treated LR soil.......................................................................... 386
Figure 6.84: Changes in FT-IR spectra of fulvic acid extracted from the PC
treated LR soil................................................................................ 387

xviii

LIST OF SYMBOLS

ADIO

Analog Digital Input Output

BBC

Boston Blue Clay

CRS

Constant rate of strain

DAS

Data acquisition system

DCDT

Direct current displacement transducer

DMM

Deep mixing method

FA

Fulvic acid

FT-IR

Fourier transform Infrared

HA

Humic acid

ICL

Intrinsic compression line

IL

Incremental loading

LIR

Load increment ratio

LVDT

Linear voltage displacement transducer

OC

Organic content (%)

PC

Portland cement

RBBC

Resedimented Boston Blue Clay

SCL

Sedimentation compression line

XRD

X-ray diffraction

Destructuring index

Secondary compression index

Cc

Compression index

Cc*

Compression index of reconstituted soil

Ck

Change in hydraulic conductivity with change in void ratio

xix
Cr

Recompression index

Cv

Coefficient of consolidation

Constrained modulus

E50

Youngs modulus at 50% axial strain

EOP

End of primary

eo, ei

In-situ (Initial) void ratio

e*

Void ratio of reconstituted soil

e*xx

Void ratio of reconstituted soil at a vertical effective stress


of xx kPa

eL

Void ratio at liquid limit

eEOP

Void ratio at end of primary consolidation

Gs

Specific gravity

Height of specimen

Iv

Void index

Iv

Modified void index

Hydraulic conductivity

kvo

In-situ vertical hydraulic conductivity

LL

Liquid limit

Loven-dried

Liquid limit of a soil after oven-drying

LOI

Loss on ignition

mv

Volume compressibility

NC

Normally consolidated

OC

Over consolidated

OCR

Over consolidation ratio

Mean effective stress

pH

acidity

PI

Plasticity index

PL

Plastic limit

Mean effective stress

Pyi

Mean effective yield stress

xx
t

Time

teop

Time to end of primary consolidation

tf

Time for end of a load increment

tuhmax

Time for generation of maximum excess pore pressure

Tv

Time factor

uh

Pore pressure

Water content

wo, wn

In-situ (natural) water content

wi

Initial water content

Strain rate

ave

Average strain

uh

Excess pore pressure

uhmax

Maximum excess pore pressure

Vertical effective stress

vc

preconsolidation pressure developed from secondary


compression

vi

consolidation pressure at which secondary compression


takes place

Preconsolidation pressure

vave

Average vertical effective stress

Unit weight

Total unit weight

Unit weight of water

xxi

ABSTRACT

Hwang, Joonho. Ph.D., Purdue University, December, 2006. Effects of Cement


Treatment on the 1-D Consolidation Behavior of a Highly Organic Soil. Major
Professors: Antonio Bobet, Maria K. Santagata.
Highly organic soils pose significant problems to the geotechnical profession
often requiring innovative design solutions and special construction procedures.
The research performed addressed one construction methodology - deep soil
mixing - that has been widely used in soft clays, but that, especially in the US,
has found limited use in organic soils.
The study utilized constant rate of strain (CRS) and incremental loading (IL) tests
to study the one-dimensional compression behavior of a soil with 40-60% organic
matter in its natural intact and reconstituted state, as well as following treatment
with cement. Testing of the soil in its intact state relied on high quality block
samples and showed that the natural soil displays intermediate behavior between
that typical of inorganic soft clays and that observed in peats. The soil showed a
high tendency to creep, with C/Cc of 0.095, at the high end of values reported in
the literature. Comparison of the compression results for the natural and the
reconstituted soil served to evaluate the degree of structuring of the natural soil,
which was found to be consistent with that typical of natural sedimentary
inorganic clays.
To investigate the effects of cement treatment on the behavior of the soil, tests
were conducted varying the cement dosage (8% to 100% by dry mass of the soil,
consistent with soil mixing practices), the curing surcharge (48, 96, and 192 kPa)

xxii
and curing duration (14 and 28 days). Substantial changes in the 1-D
consolidation behavior derive from the hydration of the cement and the structure
developed within the soils. Increasing cement percentages were found to cause
a marked increase in the preconsolidation stress (dependent also on the curing
stress), a stiffer behavior in the initial recompression region, a slight increase in
the compression index, an increase in the hydraulic conductivity and the
coefficient of consolidation, and a marked reduction in the creep coefficient and
the C/Cc ratio.
The work included an in depth characterization of the treated and untreated soil
both in terms of the organic fraction and the inorganic substrate using x-ray
dirffraction RD, FT-IR spectroscopy, and elemental analyses (LECO tests).

CHAPTER 1. INTRODUCTION

1.1. Problem statement


Numerous projects have been designed and developed in the United States in
areas where poor soils, such as peats or highly organic soils, are encountered.
Peats and highly organic soils are generally associated with high compressibility,
high rates of creep, as well as sometimes unsatisfactory strength characteristics,
which increase the risk of inadmissible settlements and/or of foundation failure.
As a result, foundations, embankments, excavations, and other ground works
become very difficult and often require costly treatments.
Several options are generally available to deal with these ground conditions: (1)
strengthening of the foundation; (2) elimination of the problem soils; (3) treatment
of the problem soils; (4) relocation of the project. In many cases, given the
impracticality and the expense of other options, the only option is strengthening
the foundation (e.g. resorting to a deep foundation) or elimination of part of the
problem soils. Techniques that have been widely used in recent years to improve
peat deposits and ground with significant organic soils include: preloading (e.g.
Gruen and Lovell, 1984, Al-Shamrani and Dhowian, 1997, Russel et al. 1999,
Mesri et al, 1997); sand or wick drains (e.g. Koda et al., 1993, Koda and Wolski,
1994) (the two techniques listed above are often combined); and vibro-concrete
columns.
A new method called the Deep Mixing Method (DMM) was developed in Sweden
and Japan in the late 1960s as an alternative to the traditional soil improvement
methods to improve the performance of soft soils. The basic principle of the

2
deep mixing method is to mix hardening agents (generally lime or cement) with
soil in situ in controlled proportions to produce columns of hardened soil which
display higher strength and stiffness and lower hydraulic conductivity than the
original untreated soil. Depending on the nature and properties (in particular the
water content) of the soil requiring treatment, the binding agents can be
introduced in slurry (wet method) or dry form (dry method). The binding agents
are mixed with soil either at the distal end of the shaft or along the drill shaft by
pure rotation of the mixing augers or by a combination of rotation of the mixing
tool and injection of binder in slurry form at high pressure (Bruce et al, 2003).
According to Holm (2003), the application of deep mixing methods for soft ground
improvement has been gradually increasing in Europe and Japan since the early
1970s and 1980s, and the total volume of soils stabilized with this technique
significantly increased in Europe, especially in Finland and Sweden, and Japan
around the early 1990s. Deep mixing was introduced in the United States in the
late 1980s and since then has been widely used, especially over the past 10
years, for many applications, including the works at Jackson Lake Dam, WY
(Ryan and Jasperse, 1989), Logan Airport, Boston, MA (Nicholson and Chu,
1994), Fort Point Channel, Boston, MA (Bruce et al, 2003), and Boston Central
Artery (ORourke and McGinn, 2004). There have been a few cases in which the
deep mixing method has been successfully applied to improve the engineering
properties of organic soils in the United States: stabilization of a 3.5 to 5 m thick
organic silty clay deposit for construction of railroad embankment in a section of
the Hudson-Bergen Light Rail Transit System, New Jersey (Esrig et al, 2003),
and stabilization of a 2.5 7.5 m thick organic clay layer (with organic content of
4 to over 30%) for the widening of I-95, Alexandria, VA (Lambrechts et al, 2003).
Due to the successful implementation in the field and the proved effectiveness of
the deep mixing methods as a soil improvement method for soft soils, a
significant research effort has gone into understanding of the physical properties

3
and the mechanical behavior of chemically stabilized soft soils and the
mechanisms responsible for the improved behavior (e.g. hnberg, H., 1996, Rao
and Rajasekaran, 1996, Tremblay et al, 2001, Holm, 2003, Jeslic and Leppanen,
2003, Chew et al, 2004, Lorenzo and Bergado, 2004). However, as observed by
Kang (2007), most of the research performed has focused on the evaluation of
the strength characteristics of the chemically stabilized soft soils, often relying on
simple unconfined compression tests. In addition, most of the work documented
in the literature has focused on treatment of inorganic clays, while there
continues to be much more limited experience with organic soils and peats. In
fact, outside of Sweden, where considerable experience exists in treating organic
soils, the presence of organic matter is generally considered an impediment to
the use of deep mixing methods. This concern is founded in the known retarding
effects of some organic compounds on the process of hydration of cement, and
does not recognize the variability in nature and characteristics of organic matter
present in soils.
The research presented in this thesis sought to address the two limitations
outlined above through a comprehensive experimental program conducted on a
highly organic soil (LOI~40-60%) comprised mainly of highly processed humic
substances, in its natural intact state, following reconstitution and after treatment

with portland cement (at dosages of 8-100% by dry mass of the soil
[corresponding to 23-230 kg of cement per m3 of treated soil]). The study relied
on the collection of high quality block samples as well as additional disturbed
samples, and an experimental program consisting of constant rate of strain
(CRS) and incremental loading tests performed using state of the art computer
controlled equipment. This work was complemented by an in depth
characterization of the chemistry of the soil both before and after treatment using
fractionation procedures, Fourier Transform Infrared Spectroscopy and elemental
analyses. (See section 1.2 for details on the objectives of the study and the
research approach).

4
As discussed in the thesis, the work performed has produced a complete data
set illustrating the effects of cement treatment on important engineering
properties, critical to engineering design (compressibility, constrained modulus,
preconsolidation stress, hydraulic conductivity, coefficient of consolidation), as a
function of effective stress and void ratio. The work contributes to the
fundamental understanding of the 1D consolidation behavior of cement treated
soils illustrating aspects of the behavior of these soils previously not understood
(e.g. secondary compression behavior). Additionally, the results provide insight
into the interaction between cement, mineral soil and organic matter, and overall
demonstrate the viability of using deep mixing for improving the properties of a
highly organic soil. Finally, the work performed on the untreated soil offers
insight into the behavior and structure of transitional geomaterials, which are
neither inorganic clays nor peats.

1.2. Scope of work and research approach


The research presented in this thesis originated from interest by the Indiana
Department of Transportation (INDOT) to explore the effectiveness of cement
treatment (in particular through deep mixing) to improve the properties of soils
with considerable organic content. With this practical premise, the overall
scope of the research was to perform a fundamental study of cement treatment
in organic soils that was founded on the evaluation of engineering properties
relevant to design and construction on these materials.
Due to time and resource limitations testing focused on one highly humified soil
with 40-60% organic matter. In addition to opportunity considerations (the
sampling location was in immediate proximity to the Purdue campus and access
was possible due to ongoing construction operations), the selection of this
specific soil was dictated by the fact that its organic content fell in the range
considered to be problematic for the use of deep mixing methods. Additionally, its
organic content made it a transitional material between inorganic clays and

5
peats. While the behavior of natural inorganic clays is well documented in the
geotechnical literature, and much work has been performed on the engineering
properties of highly organic peats (in particular those with high fiber content),
much less is known about transitional geomaterials such as the one selected
for this research. This represented an additional opportunity for the study to
make a contribution to the geotechnical literature outside of the work on cement
treatment.
Full characterization of the effects of treatment with cement on the engineering
properties of the soil would have required conducting both 1-D consolidation and
consolidated drained and undrained shear tests both before and after treatment.
Again, in part, due to time limitations, the choice was made to limit the
investigation to the analysis of the effects of treatment on the 1-D consolidation
behavior. This decision was motivated also by the fact that, in practice,
especially due to the significance of long term creep deformations, settlement
problems are often the most critical consideration for the design on highly organic
soils. (In general strength parameters for organic soils exceed those typically
observed in inorganic clays). Additionally, a thorough understanding of the
consolidation behavior of these soils is a necessary requirement for stability
calculations as the shear strength profile in a deposit is a direct function of the
effective stress profile that develops over time with consolidation of the soil (this
is especially relevant to staged loading).
Within the broad scope outlined above, the specific objectives of this research
were to:
a) characterize the modifications to key engineering properties (compressibility,
preconsolidation stress, coefficient of consolidation, hydraulic conductivity,
secondary compression [creep] coefficient) due to treatment with various
dosages of Portland cement (the range considered in this study [8-100% by

6
dry mass of the soil, or 23-230 kg/m3 of treated soil] is consistent with field
applications [see Section 2.6]);
b) investigate the evolution of the nature of the structure of the soil as a result of
treatment;
c) gain insight into the interactions between binder, soil mineral substrate and
organic matter that take place as a result of the treatment;
d) identify important parameters to be considered during laboratory evaluation of
cement treated soils.
The objectives listed above were pursued through a comprehensive experimental
program that included:
a) Obtaining high quality samples of the soil using a block sampling technique.
Such an approach was considered critical to ensure the sample quality
necessary for characterizing the intact engineering behavior of a soft soil such
as the one examined in this research. (Additional disturbed samples were
also obtained for the index tests, the chemical tests, and the experiments with
cement treatment).
b) Developing a laboratory methodology for preparing cement treated samples
that replicated as closely as possible field conditions.

This methodology,

discussed in Chapter 4, involved mixing and compacting the soil through a


kneading action (intended to be representative of the action of deep mixing
augers), and curing the treated samples under a surcharge providing
continuous access to water (the role of the curing surcharge and the curing
time were specifically addressed in this study).
c) Performing a careful characterization of the physical, mineralogical and
chemical properties of both the mineral and organic components of the
natural soil.

The goal of this work, which made use also of techniques

outside the realm of traditional geotechnical engineering (see details in


Chapter 3), was: i) to gain an improved understanding of the nature and the
structure of the natural soil, that could aid the analysis of the mechanical

7
tests, and ii) to provide the basis for understanding the effects of adding a
binder such as cement.
d) Characterizing the 1 D consolidation behavior of the natural soil using both
constant rate of strain (CRS) and incremental loading (IL) tests performed
making use of state of the art computer controlled equipment. CRS tests
represent the most effective means to derive compressibility, hydraulic
conductivity and coefficient of consolidation as a (continuous) function of
vertical effective stress or void ratio, while IL tests are necessary for
investigating secondary compression. The results of this phase of the testing
program were intended to provide the means for later evaluating the effects of
cement treatment on the mechanical properties and the structure of the soil.
e) Evaluating the 1D consolidation behavior of the reconstituted natural soil
again using CRS and IL tests (same rationale as above). These results were
intended to represent the baseline against which to compare the results of the
natural intact soil to quantify the effects of structure. Much work has been
done in recent years to quantify and model the structure of natural clays, but it
is not clear whether the same frameworks apply to highly organic soils.
Additionally, the behavior of the reconstituted specimens would serve to
evaluate the effects of treatment of the soil through deep mixing. In deep soil
mixing it is in fact unavoidable that, due to the action of the augers, the
natural structure of the soil is removed.
f) Investigating the interactions occurring between mineral soil, cement and
organic matter using extraction procedures and characterization tests typically
used for soil analysis. This work was aimed at gaining insight into the nature
of the new geomaterial generated by the treatment of the natural soil and
into the mechanisms responsible for the observed mechanical behavior.

1.3. Organization of the thesis


This thesis consists of seven chapters, including this introduction.

8
Chapter 2 presents background information on the characteristics of organic
matter and the 1-D consolidation behavior of peats and highly organic soils. An
overview of the methods used in practice for improvement of the engineering
properties of highly organic soils obtained from the available literature is
presented. The reaction of cement hydration and the interaction of hydrated
cement and organic matter are discussed. The effects of treatment on the
compression behavior and hydraulic conductivity of organic soils, including soils
with low organic content and peats, obtained from the literature, are summarized.
Chapter 3 discusses the soil used in this research program. The chapter
presents the index properties as well as the results of the extensive work
performed to characterize both its mineral and organic components.
Chapter 4 describes the experimental methods employed in the research to
investigate the effects of treatment on the 1-D consolidation behavior of the soil.
The variables investigated in the experimental program are presented. The
sample preparation procedure developed to produce laboratory soil samples and
the experimental equipment used to perform the tests are discussed.
Chapter 5 presents the results obtained from constant rate of strain (CRS) and
incremental loading (IL) consolidation tests performed on the intact and the
reconstituted soils. The 1-D consolidation behavior, including both primary and
secondary consolidation, of the intact and the reconstituted soils and the strain
rate effects are discussed.
Chapter 6 presents the results obtained from the CRS and IL consolidation tests
performed on the PC treated soils to investigate the effects of treatment on the 1D consolidation behavior. The effects of curing, including curing time and
surcharge, are discussed. A general framework of the 1-D consolidation behavior
of chemically stabilized soils, developed based on the results obtained from this
experimental program and the available literature, is presented. Additionally, the
chapter discusses the factors affecting the 1-D consolidation behavior of
chemically stabilized soils, and the results of extraction tests performed on the

9
chemically stabilized soils which illustrate the changes in chemical makeup due
to treatment with Portland cement.
Chapter 7 summarizes the work performed and presents the conclusions drawn
from the analysis of the results. Recommendations for future research are also
presented.

10

CHAPTER 2. BACKGROUND

The importance of organic soils is recognized in many different disciplines


including geotechnical engineering, soil science, and environmental engineering.
In geotechnical engineering it is the engineering behavior of these soils which is
of greatest interest, and which has been the focus of considerable research. The
presence of organic matter confers soils distinct engineering properties (e.g. high
compressibility, high hydraulic conductivity, and high creep) which set them apart
from other soft soils and make them considered problem soils, requiring special
consideration both during design and construction. From a soil science and
environmental engineering perspective organic soils represent, instead, a
significant resource due to their role as a major source of nutrients for plants and
crops, and their influence on carbon dynamics and water quality. In these fields
the focus is primarily on the characterization of the physico-chemical and
structural properties of organic matter and the investigation of its interactions with
minerals and pore solution.
The term organic soil generally refers to a soil that contains a significant
amount of organic matter. Methods used to classify these soils very significantly
depending on the discipline. For example, in geotechnical engineering a soil is
classified as organic based either on the organic content or on the change in the
Atterberg limits after oven drying. In soil science, soils are generally classified
based on the presence or absence of major diagnostic characteristics, such as
horizon morphology, wetness, climate, mode of deposition, texture and so on
(Soil Taxonomy, 2003). Since soil genesis is the underlying factor in the

11
classification in soil science, soils with similar organic contents can be classifies
as one of the 12 orders of soil taxonomy.
As discussed in the introduction chapter, this research deals with the 1D
consolidation behavior of a highly organic soil in its intact and reconstituted state
as well as following treatment with cement. The work contributes to the
knowledge base on the mechanical behavior of organic soils, and provides
insight into the behavioral and structural changes associated with the treatment
of the soil using cement. The latter results are of significance, in particular, for
the application of deep mixing methods to highly organic deposits.
The present chapter is intended to provide background information on several
topics relevant to the work conducted. It is organized in eight sections that cover
two main topics: the nature and behavior of natural organic soils (sections 2.12.5), and cement treatment of organic soils (sections 2.6-2.8). Sections 2.1 2.3
focus on summarizing relevant information primarily from the soil science
literature on the approaches to classify organic soils and the main factors
influencing soil formation (section 21.), the chemical characteristics of organic
matter and the interaction of mineral soil and organic matter. The goal of these
sections is to provide the non soil scientist reader with the background required
for understanding some of the characterization tests performed both on the
natural soil (presented in Chapter 3) and the cement treated soil (see Chapter 6),
used in this research and the hypotheses presented later in the thesis (Chapter
6) on the interaction of organic soil and cement during treatment.
Sections 2.4 and 2.5 go on to summarize the effects of the presence of organic
matter on various index and engineering properties. Specifically, section 2.4
presents an overview of the role of organic matter on water holding capacity,
hydraulic conductivity, cation exchange capacity, and Atterberg limits, while
Section 2.5 discusses specifically the 1-D consolidation behavior of organic soils,

12
which represents the focus of the research presented in this thesis. In this
section, literature results from tests conducted on various highly organic soils are
presented and discussed to highlight the differences in behavior in comparison to
mineral soils, and the effects of organic matter and fiber content.
The remaining three sections address the topic of the treatment of organic soils
with cement. First, section 2.6 presents an overview of deep mixing, a soil
improvement method which is finding increasing application for treating organic
soils. In this section the aim is to summarize some key aspects of this method:
the general construction procedure, the various types of deep mixing methods,
the applications of this method, and the typical mechanical properties of deep
mixed soils. Sections 2.7 and 2.8 go on to discuss the interaction of organic soils
with cement, and summarize what is known about the 1-D compression behavior
of chemically stabilized organic soils.

2.1. Types of organic soils


Most classification systems for organic soils and peats are based on organic
content (Landva et al, 1982). The ASTM standard classifies a soil with organic
content less than 75% by dry mass of the soil as muck or organic soil, while a
soil with organic content higher than 75% is termed a peat. The International
Peat Society (IPS) uses an organic content of 20% to classify a soil as a peat
(>20% organic content) or as organic soil (less than 20% organic content). Other
classification systems identify yet different thresholds between minerals soils,
organic soils and peats. Other widely used system for classification of organic
soils and peats are the Von Post and the Radforth systems (the latter is
applicable only to soils with 80% or higher organic content). These systems use
various parameters including organic content, fiber content, type of botanical
matter, and degree of decomposition for classification purposes. The details of
these classification systems can be found in Radforth (1969) and Landva and
Pheeny (1980). The Unified Soil Classification System (USCS) does not classify

13
organic soils based on the organic content. Rather, it classifies a soil as organic if
the liquid limit of the soil after oven drying at 105 110 C is less than 75% of its
liquid limit value before oven drying. Recent research by Huang et al. (2006)
indicates that this reduction in liquid limit corresponds to an organic matter
content in the 15-20% range.
Fiber content is another important factor used for classification of organic soils
and peats, and has a significant impact on the structure and properties of highly
organic soils. Fiber content indicates the degree of decomposition of the organic
matter. There is an inverse relationship between fiber content and degree of
decomposition (Walmsley, 1973), i.e. a soil with higher fiber content is less
decomposed than a soil with less fiber content. Fiber content is normally
measured by a wet sieving method. ASTM standard D1997 provides the
procedure for determination of fiber contents. Depending on the fiber content and
the type of fiber, the Radforth classification system classifies peats (>80%
organic content) into three main groups: (1) amorphous granular peats, (2) fine
fibrous peats, and (3) coarse fibrous peats (MacFarlane and Williams, 1974). The
effects of organic matter and fiber content on the 1-D consolidation behavior of
organic soil will be discussed further in section 2.5.
In soil science, soils are classified into 12 orders of soil taxonomy based on the
presence or absence of major diagnostic characteristics, such as horizon
morphology, wetness, climate, mode of deposition, texture, absence or presence
of diagnostic horizons and so on (Soil Taxonomy, 2003). The 12 soil orders of
soil taxonomy and their characteristics are summarized in Table 2.1. Figure 2.1
shows the principal characteristics of the 12 soils orders of soil taxonomy. In
general, soils with 25% or more organic content are classified as Histosols.
According to soil taxonomy, soils that are not formed in permafrost, a condition
by which soils remain below 0 C for 2 or more years in succession, are
classified as Histosols if half or more of the upper 80 cm is organic. A soil is also

14
classified as a Histosol if the organic materials rest on rock or partially fill voids in
fragmental, cindery, or pumiceous materials (Soil Taxonomy, 1999). As shown in
Figure 2.1, most Histosols occur in wet-land environments. Histosols can form in
any moist climate in which plants can grow, from equatorial to arctic regions, but
are most prevalent in cold climates up to the limit of permafrost (Brady and Weil,
1999).
The amount of soil organic matter in a soil depends on the five soil-forming
factors: climate, vegetation, topography, parent material and time, in a
descending order of significance (Stevenson, 2003).
Climate, including temperature and moisture, is the most important factor since it
determines the type and the amount of plant material produced, and the intensity
of the microbial activity in the soil. In a humid climate, Spodosols and Alfisols are
formed in the forest, while in a semiarid climate where grassland is formed,
Mollisols are formed. In general, soils formed in grassland (Mollisols) have the
highest organic matter content since the harsh climatic conditions (such as cold
winters and hot and dry summers) preserve organic matter. In well-aerated
regions such as desert, semidesert and certain tropical areas, the microbial
activity is very high and thus the soils in these areas have the lowest organic
matter content. The amount of the organic matter in the soil formed under
restricted drainage condition, such as Histosols and Inceptisols, is not
significantly influenced by the climate. Anaerobic conditions completely prevent
microbial activity over a wide range of temperature.
Vegetation is the second most important factor influencing organic matter
content. Larger amounts of plant material in grassland area lead to a higher
organic matter content in the soil (Mollisols). In the soils formed in deciduous
forests, where the leaves are fallen off or shred seasonally, the soils with higher
calcium content (Alfisols) have organic matter concentrated in the top 10 15cm

15
of the soil, while the soils with lower calcium content (Spodosols) have organic
matter only in the topmost section.
Topography affects the organic matter content of a soil via its influence on
temperature, rainfall runoff or retention. Soils in depressions or down hill areas
have a higher chance for accumulation of rainfall runoff and thus anaerobic
conditions are facilitated compared to soils on slopes. The anaerobic condition
prevents microbial activity, so the soils in depressions have in general higher
organic content. The soils on slopes facing north have lower temperatures, and
have higher organic content than soils on slopes facing south.
The parent material affects the organic matter content through its influence on
the texture of the soil. Soils with higher clay content, especially smectite, have a
higher affinity for organic molecule adsorption, and thus tend to have higher
organic content than sandy or loamy soils.
The rate (time) of organic matter accumulation is rapid initially, decreases slowly,
and reaches equilibrium depending on the soil type in a time that ranges between
110 years for fine-textured parent material to 1500 years for sandy deposits. The
rate of the organic matter accumulation reaches equilibrium by (1) production of
humic acids that can resist microorganism attack or (2) protection of humus
decay through the interaction with mineral matter.

2.2. Characteristics of soil organic matter


The term soil organic matter is used as a synonym of humus, and is defined as
the total organic material in soils, including litter, light fraction, microbial
biomass, water-soluble organics, and stabilized organic matter (Stevenson,
1994). The undecayed plant and animal tissues and their partial decomposed
products are excluded from soil organic matter (Sparks, 2003). Litter is defined
as the macro-organic matter that lies on the soil surface. Light fraction includes

16
plant residues in varying stages of decomposition. The microbial biomass is the
organic matter present as live microbial tissue, including bacteria, fungi, algae,
and so on. Water soluble organics refers to the organics contained in the soil
solution (Stevenson, 1994).
The classification scheme for the components of soil organic matter is yet to be
developed and many researchers have suggested various classification schemes
(Stevenson, 1994). Figure 2.2 presents the most general classification of soil
organic matter. As shown in the figure, soil organic matter, or humus, consists of
two main components: non-humic substances and humic substances.
Non-humic substances consist of compounds belonging to the well-known
classes of organic chemistry, such as amino acids (including polypeptides),
carbohydrates (including monosaccharides, oligosaccharides, and
polysaccharides), and lipids (including fats, waxes, resins, and so on) (Schnitzer
and Kahn, 1972). Non-humic substances are easily attacked by soil microorganisms and exist in the soil for a relatively short period of time (Schnitzer and
Kahn, 1972, Sparks, 2003).
As a result, most of the organic carbon in soils is in the form of humic
substances. This term refers to a general category of naturally occurring,
biogenic, heterogeneous organic substances that can generally be characterized
as being yellow to black in color, of high molecular weight, and refractory
(Sparks, 2003). Humic substances can be divided into three fractions: humin,
humic acid, and fulvic acid, based on solubility characteristics. Humin refers to
the soil organic matter fraction that is insoluble in alkali and remains after
extraction of the humic and fulvic acids with dilute alkali. The humin fraction
consists of four main materials: (1) humic acid that is too strongly bound to
mineral matter to be separated, (2) highly humidified humic matter with a carbon

17
content higher than 60% and insoluble in alkali, (3) fungal melanins, and (4)
paraffinic substances (Stevenson, 1994).
Fulvic acid refers to the colored soil organic matter that is soluble in both alkali
and dilute acid. On the other hand, humic acid is the dark-colored organic matter
that is soluble in alkali but is insoluble and precipitates in dilute acid. In general,
fulvic acid contains significantly more acidic functional groups, such as carboxyl
acid (COOH), and thus is more reactive than humic acid. In addition, fulvic acid
has lower molecular weight, and contains higher oxygen but lower carbon than
humic acid (Stevenson, 1994).
The distribution of the three fractions of soil organic matter varies depending on
the soil type, climate, drainage and depth in the soil profile (Tan, 2003). However,
quantification of the three fractions of soil organic matter is very complex due to
lack of standardized extraction, fractionation, and purification methods
(Stevenson, 1994). Differences in the distribution of these three fractions (and in
particular of the humic acid to fulvic acid ratio [HA/FA]) are however reported to
occur for soils of different origin. In general, forest soils (Alfisols, Spodosols, and
Ultisols) have higher fulvic acid content, and peat and grassland soils (Mollisols)
have higher humic acid content (Stevenson, 1994). However, Tan (2003) stated
that a Spodosol from the border of Georgia and Florida has a humic acid
concentration 10 times higher than that of fulvic acid, with a FA/HA ratio of 0.13.
Vertisols have very high humin content (Stevenson, 1994). Kononova (1966)
showed that among several great soil groups obtained in the USSR, Mollisols
have the highest humic acid to fulvic acid ratio in the range of 0.7 2.0. The
distributions of organic matter in four different Histosols obtained from sugarcane
fields in Palm Beach County, Florida, are summarized in Table 2.2. The figure
shows that for the same type of soil the HA/FA ranges from 0.2 to 5.4, and that
even for a given soil the HA/FA ratio can vary significantly with depth. Zelazny
and Carlisle (1974) observed that the degree of decomposition increased with

18
depth, and soil samples from deeper horizons had higher humin and lower humic
acid and fulvic acid contents. Zelazny and Carlisle (1974) and Kononova (1966)
state that peats and Histosols contain higher humic acid content than fulvic acid.
According to Tan (2003), the humic acid content in peats and Histosols generally
increases with increasing degree of humification.
Although the distribution of organic matter varies depending on soil type and
depth, the elemental compositions of all three fractions of organic matter are very
similar regardless of soil type (Sparks, 2003). The elemental composition of
humic acid and fulvic acid extracted from soils from widely different climates is
shown in Tables 2.3 - 2.5. As shown in these tables, the elemental composition
of humic and fulvic acids fall in a very narrow range regardless of climate. Fulvic
acids in general have higher O and S contents but lower C, H, and N contents
than humic acids. The major elements of humic and fulvic acids are C and O.
Humic acids have about 51 59% C, 3 6% H, 1 6% N, 0.1 1.5% S, and 33
38% O. Fulvic acid has about 41 51% C, 4 7% H, 1 3% N, 0.1 3.5% S,
and 40 50% O. The elemental compositions of humic substances of the four
Histosols from Florida presented in Table 2.2.are shown in Table 2.6. The
elemental composition of humic substances of Florida Histosols fall in general in
the range formed by other organic soils (see data in Tables 2.3 2.5).
Understanding the macromolecular structure of humic substance is very
important since it can affect the chemistry of soil organic matter-mineral
complexes, and stability of organic matter-mineral aggregates. However, the
actual macromolecular structure of humic substances is not fully understood yet,
as it changes depending on the pH, electrolyte concentration, ionic strength of
solution, and the concentration of humic and fulvic acids (Sparks, 2003). In
addition, the complexity and heterogeneity of humic substances make it hard to
understand the basic structure of humic substances.

19
Chen and Schnitzer (1976) observed the shape, size and degree of aggregation
of humic and fulvic acids at various pH using a scanning electron microscopy
(SEM). Chen and Schnitzer (1976) observed that fulvic acids have relatively
more open structure under a highly acidic condition, pH between 2- 3, with
elongated fibers and bundles of fibers. As the pH increases, the structure of
fulvic acids changes from a fibrous to a sponge-like structure. At pH above 7, the
fulvic acid structure starts changing distinctively, to a structure with improved
orientation. At pH of about 8, the fulvic acid forms sheets. At pH of about 10,
fulvic acids shows fine, homogeneous grains.
Schnitzer and Khan (1978) observed that although the effect of pH on the
structure of humic acid is similar to that observed on fulvic acid, the major
transition of the humic acid structure changes at higher pH, generally between 6
and 10, due to the low solubility of humic acid in water. At a lower pH, humic
acid shows a fibrous structure, and gradually changes to a more sheet-like
structure as the pH increases. The size of the humic acid particle decreases as
the pH increases.
Unlike minerals, especially phyllosilicate minerals, humic substances have no
regularly repeating structural units, and thus no two molecules are the same
(Sparks, 2003). Based on the elemental composition of humic substances and
employing advanced analytical techniques such as 13C Nuclear Magnetic
Resonance (NMR) spectroscopy, gas chromatography and mass spectrometry,
models for the structure of humic substances have been proposed (e.g. Schulten
and Schnitzer, 1997, Stevenson, 1994). It is reported that humic substances are
made of a variety of functional groups, such as carboxyl (COOH), phenolic OH,
enolic OH, quinine, hydroxyquinone, lactone, ether, alcoholic OH, and etc.,
although the exact amounts of these functional groups are unknown (Stevenson,
1994).

20
According to a two-dimensional structure of humic acid proposed by Schulten
and Schnitzer (1993), humic acid is composed of aromatic rings bridged by O-,
-CH2-, -NH-, -N=, -S-, and other groups, and contains both free OH groups and
double linkage of quinines (Stevenson, 1994). Humic acid has high
microporosity with voids of different dimensions that can trap and bind other
organics, inorganic components, and water (Sparks, 2003). Humic acid contains
abundant carboxyl (COOH) and hydroxyl (OH) groups on both aromatic rings and
aliphatic side chains (Deng and Dixon, 2002).
According to a two-dimensional structural model of fulvic acid proposed by
Schnitzer and Khan (1972), fulvic acid consists of phenolic and
benxenecarboxylic acids held together through hydrogen bonds to form a
polymeric structure of considerable stability (Stevenson, 1994). Fulvic acid has
higher carboxyl group content than humic acid. As mentioned before, fulvic acid
contains higher O than humic acid, and most of its oxygen occurs in carboxyl
form.
The distribution of oxygen-containing functional groups in humic and fulvic acids
extracted from soils of widely different climatic zones is shown in Table 2.7. The
results show a very wide range of distribution of oxygen-containing functional
groups. In general, fulvic acids have higher total acidities, the sum of mineral and
carbonate acidities, than humic acid. The higher carboxyl (COOH) and acidic OH
(presumed to be phenolic OH) groups content contribute to the higher acidities of
fulvic acid, with carboxyl being the most important (Stevenson, 1994). Note that
acidic soils from cool temperature regions have higher carboxyl group content
and highest acidities. This may indicate the effect of climate, as mentioned in
section 2.2, in that cold weather prevents microbial activity and thus increases
the amount of humic substances in the soil. The concentration of organic
functional groups in humic substances from 11 Florida Histosols is summarized
in Table 2.8. Fulvic acid has the highest total acidity, in the range of 860 40,

21
which is related to the higher concentration of carboxyl (COOH), phenolic OH,
and carbonyl (COO) functional groups.
2.3. Soil - organic matter interaction
Most of the humic substances in soils occur in insoluble forms. There are three
mechanisms through which humic substances are bound: (1) as insoluble
macromolecular complexes, (2) as macromolecular complexes bound together
by di- or trivalent cations, such as Ca2+, Fe2+, and Al3+, or (3) in combination with
clay minerals, such as through bridging by polyvalent cations (clay metal
humus), hydrogen bonding, van der Waals forces, and so on. The first
mechanism is especially important in peat or highly organic soils, where clay and
metal ions contents are relatively lower than the organic content (Stevenson,
1994).
To better understand the interaction between clay and soil organic matter, it is
necessary to briefly review the structure and the surface chemistry of clay
minerals. The most important clay minerals in soil are kaolinite, mica, vermiculite,
and smectite (Stevenson, 1994).
Kaolinite consists of one silica tetrahedral sheet and one aluminum octahedral
sheet, and thus is called a 1:1 type phyllosilicate mineral. The chemical formula
for kaolinite is Si4Al4O10(OH)8. As the formula indicates, kaolinite is neutrally
balanced. Layers of kaolinite are held by strong hydrogen bonding (between a
hydrogen ion in the basal plane of octahedral sheet and an oxygen ion in the
basal plane of tetrahedral sheet of the adjacent kaolinite mineral), and thus there
is no interlayer space between kaolinite minerals (Sparks, 2003).
Mica is a 2:1 type phyllosilicate mineral and consists of two octahedral sheets
and one tetrahedral sheet. Depending on the type of the cation in the octahedral
sheet, mica has two subgroups: dioctahedral mica (such as muscovite) with
aluminum as the octahedral cation, and trioctahedral mica (biotite) with

22
magnesium as the octahedral cation. For each subgroup, one of the four silica
ions (Si4+) in the tetrahedral sheet is isomorphically substituted, without structural
alteration, with one aluminum ion (Al3+). This isomorphic substitution yields a
negative one (-1) layer charge per unit cell. The net negative layer charge of
mica is neutralized by a monovalent potassium ion (K+), which is fixed in the
interlayer position. The fixation of potassium ion causes no interlayer space
between mica minerals (Schulze, 2002).
Vermiculite is also a 2:1 type phyllosilicate mineral but with less isomorphic
substitution of the tetrahedral cation than in mica. Vermiculite has a net layer
charge of -0.6 to -0.9. The layer charge is balanced with exchangeable cations,
primarily Ca2+ and Mg2+, in the interlayer. In general, vermiculite can have two
planes of water molecules surrounding the hydrated exchangeable cations in the
interlayer space (Schulze, 2002).
Smectite is also a 2:1 type phyllosilicate mineral as mica and vermiculite, but has
a lower degree of isomorphic substitution of the tetrahedral cations than mica
and vermiculite. The unit cell formula of the tetrahedral sheet of smectite is
(Si3.67Al0.33). Isomorphic substitution gives smectite a net layer charge in the
range of -0.25 to -0.6. The low layer charge of smectite causes high swelling
potential, and smectite can have four layers of water molecules surrounding the
hydrated cations inside the interlayer space (Schulze, 2002).
The reactivity of the organic soil particles depend on the type of clay minerals,
organic matter content, and the pH of the soil. As mentioned before, the surface
of 2:1 type phyllosilicate minerals is permanently negatively charged due to
isomorphic substitution in the tetrahedral sheet, while 1:1 type phyllosilicate
minerals are neutrally charged. The charge sites on the 2:1 type phyllosilicate
minerals where the isomorphic substitution has occurred are called constant
charge sites. The surface of the tetrahedral sheet of 1:1 type phyllosilicate

23
mineral is referred to as the neutral siloxane surface. The siloxane surface
does not interact strongly with the H bond network of water molecule and is
termed a hydrophobic surface (Johnston and Tombcz, 2002). The surface
charge of edges sites of 1:1 and 2:1 type phyllosilicate minerals changes with
pH, and thus these sites are referred to as conditionally charged sites, variable
charge sites, or pH dependent sites (Johnston and Tombcz, 2002). The trivalent
aluminum ions (Al3+) are surrounded by six hydroxyl ions (OH-). Each hydroxyl
ion is bonded to two aluminum ions allocating -0.5 charges to aluminum ions.
The trivalent aluminum ion is neutrally balanced by -0.5 charges from six
hydroxyl ions. However, at the edge sites, the hydroxyl ions are bonded to only
one aluminum ion, and thus have net -0.5 charges. The charges at these sites
can change to +0.5 if additional hydrogen ions are bonded to the hydroxyl ions at
low pH.
According to a two-dimensional model of humic acid proposed by Schulten and
Schnitzer (1993), humic acid contains about 21 carboxyl groups, which are
responsible for the high cation exchange capacity of soil organic matter. Most of
the carboxyl groups are highly reactive and act as polar sites. The non polar
sites of humic acid, in general, interact with nonpolar organic compounds
(Johnston and Tombcz, 2002).
The reaction between soil organic matter and soil mineral can occur through two
different mechanisms: (1) by attachment to clay mineral surfaces, such as
through cation and anion exchange, bridging by polyvalent cations (clay metal
ion humic substances), Hydrogen bonding, and van der Waals forces, or (2) by
penetration into the interlayer spaces of expanding-type clay minerals, such as
smectite and vermiculite (Stevenson, 1994).
The organic matter and soil interaction can occur in both edge sites and surface
of clay minerals. Electrostatic bonding occurs through cation exchange or

24
protonation at the edge sites of clay minerals. Under basic conditions, the edge
sites of clay minerals are negatively charged and attract inorganic cations.
Electrostatic bonding through cation exchange occurs when positively charged
organic cations replace inorganic cations at the edge sites. The negatively
charged clay mineral surface and organic anions, such as carbonyl group (COO), repel each other. If polyvalent cations such as Ca2+, Fe3+, and Al3+ are
available in the solution, these can neutralize both the negatively charged clay
surface and organic anions, and adsorption of humic substances on to clay
minerals occurs. In general, the divalent Ca2+ ion does not form strong
complexes with organic molecules, while the trivalent Fe3+, and Al3+ form strong
complexes with organic compounds. Hydrogen bonding also occurs at the edge
sites of clay minerals. Hydrogen bonding is weaker than covalent bonds but is
stronger than Van der Waals forces.
The area of clay surface that forms complexes with organic substance depends
on organic content and type and amount of clay. In soils with higher organic
content, such as Histosols and Mollisols, almost all of the clay minerals are
covered with a thin layer of organic matter (Stevenson, 1994).

2.4. Effects of organic matter on engineering properties of soils


Humic substances reside on the external surface of soil minerals and strongly
influence the chemical behavior of solutes in the soil solution. The size, shape,
and aggregation state of humic substances are influenced by changes in water
content, pH and ionic strength (Johnston and Tombcz, 2002). The effect of
organic matter on the engineering properties of soil increases with increasing
organic content.
As mentioned in section 2.1, the Radforth classification system classifies peats
(>80% organic content) into three main groups: (1) amorphous granular peats,
(2) fine fibrous peats, and (3) coarse fibrous peats, based on the fiber content.

25
These three peat types are ranked based on water content, void ratio, unit weight,
shear strength, permeability, and compressibility in Table 2.9 (MacFarlane and
Williams, 1974). Amorphous-granular peats have the lowest shear strength and
tensile strength, while coarse-fibrous peats have the highest natural void ratio
and compressibility among the three types of peats. The effects of organic matter
and fiber content on the 1-D consolidation behavior of organic soil will be
discussed further in the following section.
The most important effect of organic matter is its high water holding capacity
(MacFarlane, 1969). The water holding capacity of organic soil varies depending
on the degree of decomposition, fiber content, and organic content. In general,
fibrous peats have higher water content than more decomposed and amorphous
peats. Pure peats, that have 0% ash content, can have water content between
750 and 1500 %. The water content of organic soil decreases sharply with
increasing ash (or inorganic material) content (MacFarlane, 1969). Many
researchers have observed that the compressibility of peats is closely related to
the water content (MacFarlane, 1969, Cook, 1956, Landva and La Rochelle,
1982, Mesri et al, 1997). The compressibility of peats increases with increasing
water content. Landva and La Rochelle (1982) observed a linear relationship
between compression index (Cc) and water content (w) as Cc = 0.115 w(%).
The hydraulic conductivity of organic soils and peats varies depending on organic
content, void ratio, and degree of decomposition (MacFarlane, 1969). Organic
matter binds soil particles together into structural units called aggregates. These
aggregates help an organic soil to maintain a loose, open, granular state. The
aggregation of soil particles by organic matter leads to an increase in the
hydraulic conductivity (Stevenson, 1994). According to Walmsley (1973), with
increasing degree of decomposition, there is a small decrease in porosity and a
considerable decrease in the effective diameter of the pores. This implies that
more decomposed amorphous peats have, in general, lower hydraulic

26
conductivity than less decomposed fibrous peats. MacFarlane (1969) mentioned
that the highly colloidal, amorphous peats tend to discourage permeability by
water, whereas the open-meshed fibrous peats are initially quite permeable. The
hydraulic conductivity of organic soils and peats covers a wide range depending
on the size and continuity of the pores, which are determined by the arrangement
of particles.
Although both clay and organic matter contribute to the cation exchange capacity
of soil, it is the organic matter that controls this important property. The cation
exchange capacities of kaolinite, mica, vermiculite (trioctahedral) and
montmorillonite (smectite) are in the range of 2 15, 10 40, 100 200, and 80
150 cmol/kg, respectively (Sparks, 2003). The cation exchange capacity of soil
organic matter ranges generally from 150 to 300 cmol/kg, and can be as high as
1400 cmol/kg (Stevenson, 1994). The contribution of organic matter to the cation
exchange capacity of organic soil changes depending on soil pH. As the pH of
soil increases, the contribution of soil organic matter increases (Sparks, 2003,
Stevenson, 1994). Approximately 25 90% of the total cation exchange capacity
of the mineral soil in surface layers is due to organic matter, while all the cation
exchange capacity of highly organic soils or peats is due to the organic matter. In
general, organic soils with higher degree of decomposition have higher cation
exchange capacity. The contribution of humic and fulvic acids mainly comes from
the ionization of carboxyl group (COOH), and to a lesser degree from phenolic
OH and NH groups (Stevenson, 1994).
The liquid and plastic limits of organic soils generally increase with increasing
organic content (Tremblay et al, 2003). Landva el al. (1982) showed that the
addition of organic carbon (organic matter) increased both liquid limit and plastic
limit significantly.

27
2.5. 1-D consolidation behavior of organic soils
Peats and organic soils exhibit the highest compressibility among geotechnical
materials (Mesri et al, 1997). The high compressibility of peats and organic soils
is caused by the high natural water contents and high void ratios. According to
Tan (2003), the high water holding capacity of humic substance and the
cementation effect of humic acid, which causes aggregation of inorganic minerals,
have an effect of creating and preserving a stable structure that provides a large
void space in the soil.
As illustrated in Figure 2.3, the compression index (Cc) and the natural water
content (wo) of peats are approximately one order of magnitude higher than
those of clay and silt deposits. According to Stevenson (1994), the higher
compressibility of organic soil is due to the high water holding capacity and the
high tendency of organic matter to form complexes with clay particles, as
summarized in section 2.3. The complexation of organic matter increases the
pore size and eventually increases the compressibility of soils (Stevenson, 1994).
Consolidation of organic soil and peat deposits is more significant due to large
secondary compression. Mesri and Godlewski (1977) observed that for any
natural soil, there is a unique relationship exists between secondary compression
index (C) and compression index (Cc). According to the C/Cc concept, the
magnitude and behavior of the secondary compression index (C) with time is
directly related to the magnitude and behavior of the compression index with
vertical effective stress (Mesri et al, 1997). The C/Cc concept can be applied in
both the over consolidated and the normally consolidated regions, so that Cc
represents the slope of the compression curve in both of the regions. The typical
values of C/Cc for various types of geotechnical materials are summarized in
Table 2.10. As shown in the table, the average value of C/Cc for peats is 0.06
0.01, the highest among natural soils. The C/Cc ratio increases with the
deformability of soil particles, and thus the high value of C/Cc for peats is a

28
reflection of the highly deformable organic particles. As shown in Table 2.11, the
values of C/Cc for various peat deposits vary from 0.035 0.10. The high values
of C/Cc and Cc for peats indicate that the secondary compression index (C) of
peats is more than one order of magnitude higher than that of clay and silt
deposits.
The 1-D consolidation behavior of organic soils and peats varies depending on
the type of soil organic matter, organic content, and fiber content. MacFarlane
and Radforth (1965) classified peats into two main types: amorphous granular
peat and fibrous peat, depending on fiber content. According to Berry and Poskitt
(1972), amorphous granular peats exhibit a plastic structural resistance to
compression, and show compression behavior similar to clay soils. This implies
that the secondary compression of amorphous granular peat is due to the
gradual adjustment of the soil grains into a more stable configuration following
the rupture of the soil structure that occurs during the primary stage (Berry and
Poskitt, 1972). On the other hand, for fibrous peat the primary consolidation
behavior is due to drainage of water from the macronetwork (i.e. the dissipation
of excess macropore water pressure), and the secondary compression is due to
the very slow drainage of water from the micropores into the macronetwork
(Berry and Poskitt, 1972).
Reference compression curves and curves of hydraulic conductivity versus void
ratio for a fibrous peat and an amorphous granular peat are compared in Figure
2.4. Fibrous peats have more open fabric or structure than amorphous peats,
and exist in general at higher void ratios. Fibrous peats exhibit steeper
compression curves than amorphous granular peats. In general, the compression
curve of fibrous peats lies above that of amorphous granular peats at the same
vertical effective stress (Figure 2.4 (a)). Due to the high water contents and void
ratios, the compressibility of organic soils and peats are higher than clay and silt
deposits. As shown in Figure 2.3, the natural water content and the

29
compressibility of peats are approximately one order of magnitude higher than
those of clay and silt deposits. Figure 2.4 (b) shows that the hydraulic
conductivity of fibrous peats is generally smaller than that of amorphous peats at
the same void ratio. This implies that the effective void space that serves as
channel for flow is smaller fibrous peat compared to amorphous granular peat
(Mesri et al, 1997). However, the rate of change of hydraulic conductivity with
logarithm of void ratio (Ck = e/logk) is almost the same for fibrous and
amorphous granular peats.
The 1-D consolidation behavior of a fibrous peat from Middleton, Wisconsin, was
experimentally investigated by Mesri et al (1997) in a extensive study which
provides excellent reference data for the behavior of this sol type (these data will
be used in Chapter 5 for comparisons between the behavior of the soil tested in
this research and that of highly fibrous peats). The total unit weight, natural water
content, in-situ void ratio, organic content, and specific gravity of Middleton peat
are = 9.1 10.2 kN/m3, w = 623 846%, e = 10.1 14.2, LOI = 90 95%, and
Gs = 1.53 1.65, respectively. The plant fibers are porous and not decomposed.
The range of the compression curves determined by testing specimens obtained
from block samples of Middleton peat are shown in Figure 2.5. The compression
curves of this fibrous peat are characterized by an S-shape and values of the
preconsolidation pressure in the 30 43 kPa range. Due to S-shape curve, the
compression index decreases from 8.5 at 2p to 2.0 at 1600 kPa. The swelling
index (Cs) is in the range of 0.3 0.9, and varies slightly depending on the value
of the vertical effective stress from which unloading takes place (Cs/Cc decreases
from 0.30 to 0.07 as the stress from which unloading takes place goes from 1600
kPa at 200 kPa) (Mesri et al, 1997).
Hydraulic conductivity data of Middleton peat measured from falling-head and
constant-head tests and from constant rate of strain (CRS) consolidation tests is
plotted in Figure 2.6. The figure shows that the initial vertical hydraulic

30
conductivity (kvo) of Middleton peat is in the 2x10-8 1x10-7 m/sec range. At an
in-situ void ratio of 12.0, the horizontal hydraulic conductivity (kho) is
approximately one order of magnitude higher than the vertical hydraulic
conductivity. The high anistropy in hydraulic conductivity is caused by the fabric
anisotropy of this peat. Photomicrographs show, in fact, that the plant materials
of Middleton peat are elongated and horizontally oriented.
Values of Ck (Ck =e/logk) normalized by the initial void for peats and soft clay
and silt deposits are compared in Figure 2.7. As shown in the figure, Ck for soft
clay and silt deposits is close to eo/2, and to eo/4 for peats. According to Mesri et
al. (1997), both the low value of Ck/eo and high values for Ck for peats as
compared to clays and silts suggest that only part of the pores in peat
macropores in between particles are serving as flow channels. The high initial
hydraulic conductivity of peat is due to large macropores between fairly large
peat particles, and the decrease in hydraulic conductivity with compression is
due to the reduction of the size of the macropores.
The secondary compression index (C) and the compression index (Cc) of
Middleton peat, obtained in the v/ p range of 0.36 52.2, are shown in Figure
2.8. The results show a linear relationship between the primary and secondary
compression coefficients. The C/Cc ratio of the Middleton peat, which is the
slope of the linear best fitting curve of C and Cc data, is 0.052. This value lies in
the lower range of average values of C/Cc for peats (0.06 0.01).
Figure 2.9 shows the compression curves and the curves of hydraulic
conductivity versus void ratio for two other fibrous peats, denoted as NBR-2 and
NBR-3, both from James Bay, in northern Quebec, Canada (Lefebvre et al, 1984).
The soil denoted as NBR-2 samples was characterized by natural water content,
LOI, and fiber content of about 1460%, 100%, and 67%, respectively. The NBR-3
samples had, instead, a natural water content of 860%, and LOI of about 70%.

31
The average values of the preconsolidation pressure were 4.8 and 6.4 kPa for
the NBR-2 and NBR-3 samples, respectively. Comparison of the two curves
provides insight into the effects of changes in the organic matter content alone.
Compared to the NBR-3 peat, the NBR-2 soil, which is characterized by the
higher LOI, exhibits a higher initial void ratio (22.0 - 28.0, compared to 11.8 for
the NBR-3 soil), greater compressibility (Cc = 6.4 - 15.0 compared to 4.5 - 9.1 for
the lower LOI soil), a more distinct S-shaped compression curve, higher in situ
hydraulic conductivity, lower hydraulic conductivity than the NBR-3 peat at the
same void ratio, and a higher value of the parameter Ck (Ck = e/logk = 3.2 for
NBR-2 versus 1.5 for NBR-3).
Lefebvre et al. (1984) observed that the compression index increased with the
initial void ratio, and that the variation of the compression index with vertical
effective stress was smaller for samples with lower initial void ratio. As shown in
Figure 2.10, for these peats the average value of C/Cc is 0.06, which falls in the
range reported for peats, 0.06 0.01 (Mesri et al, 1997).

2.6. Deep mixing methods


Several options are generally available in practice when organic soils and peat
deposits are encountered: (1) strengthening of the foundation; (2) elimination of
the problem soils; (3) treatment of the problem soils; (4) relocation of the project.
In many cases, the only options are strengthening the foundation (e.g. resorting
to deep foundations) or eliminating part of the problem soils, since the other
options are impractical or too expensive. In recent years, the most widely used
technique to improve the ground in presence of organic soils or peat has been
preloading (Gruen and Lovell, 1984, Al-Shamrani and Dhowian, 1997, Russel et
al. 1999, Mesri et al, 1997), often in combination with sand drains or wick drains
to reduce the time for primary consolidation (Koda et al., 1993, Koda and Wolski,
1994).

32
An alternative to the methods traditionally used to deal with highly organic soils,
which is gaining increased acceptance is the deep mixing method. The Deep
Mixing Method (DMM) was developed in Sweden and Japan in the late 1960s
and was introduced in the United States in the late 1980s. Since then it has
been widely used, especially over the past 10 years, for many applications,
including the works at Jackson Lake Dam, WY (Ryan and Jasperse, 1989),
Logan Airport, Boston, MA (Nicholson and Chu, 1994), Fort Point Channel,
Boston, MA (Bruce et al, 2003), and Boston Central Artery (ORourke and
McGinn, 2004).
The basic concept of the deep mixing method is to improve the ground by mixing
the soils in situ with binding agents to produce columns of hardened soil which
display higher strength and stiffness and lower hydraulic conductivity. The most
widely employed binders include lime, fly ash, or ordinary type I Portland cement.
When deep mixing is applied in practice, the first step is to determine the area
and the depth to be stabilized. The mixing tool, usually hollow drill shafts, is
penetrated to the target depth by a rotary process. Depending mainly on the insitu water content, the binding agent can be introduced into the soil in slurry or
grout form (wet method) or in dry form (dry method). Generally, the dry method
is applied when the in-situ water content of the soil is higher than 60%. In both
methods, the binder can be mixed with the soil either by pure rotation of the
mixing tool at relatively low stress (rotary) or by the combination of rotation of the
mixing tool and injection of the binding slurry or grout into the soil at high stress
(rotary + jet). The binder can be introduced at the distal end of the shaft (end) or
along the length of the shaft (shaft).
Several types of deep mixing methods are currently available, with the main
differences being the types of equipment (e.g. type and number of the shafts)
and the construction procedure. The Federal Highway Administration (FHWA,
2000) proposed a classification system for deep mixing methods based on three

33
fundamental operational characteristics: (1) binder type (wet/dry methods), (2)
mixing principle (rotary/rotary and jet), and (3) mixing location (shaft/end).
Although a total of eight classification groups are theoretically available using the
three operational characteristics as proposed by FHWA, only four groups are
used in practice. The four classification groups include: Wet-Rotary-Shaft (W-RS), Wet-Rotary-End (W-R-E), Wet-Rotary+Jet-End (W-J-E), and Dry-Rotary-End
(D-R-E) (FHWA, 2000). Figure 2.11 shows the classification of the deep mixing
methods currently available in the United States (Bruce and Bruce, 2003).
As mentioned before, deep mixing methods have been widely used for various
applications, including hydraulic cut-off walls, excavation support walls, mass
ground treatment, liquefaction mitigation, in situ reinforcement, piles, and gravity
walls, and environmental remediation (FHWA, 2000). Depending on the type of
application, deep mixing can be employed to produce a wide range of treated soil
structures, such as a single column, rows of overlapping columns, grids or
lattices, or blocks (Bruce and Bruce, 2003). The diameter of each column ranges
typically from 0.6 to 1.5 m, and the length ranges from 20 to 40 m.
The strength/stiffness requirements of the stabilized soils may be different
depending on the application. In general, higher strength and stiffness are
required for in-situ reinforcement (typically unconfined compression strength
higher than 1 MPa) than for block ground treatment (unconfined compression
strength about 0.15 0.5 MPa). The strength and the stiffness of the soils
treated with the deep mixing methods are affected by many factors. According to
Terashi (1997), the increase in the strength of the treated soil depends on the
followings:
(1) characteristics of hardening agent, such as the type and the amount of
hardening agent, and the ratio of mixing water and additives (for wet
method);

34
(2) characteristics and conditions of the soil, such as physical, chemical,
and mineralogical properties of the soil, organic content, water content,
and pH of pore water;
(3) mixing conditions, such as degree of mixing, timing of mixing/re-mixing,
and the quality of the hardening agent;
(4) curing conditions, including temperature, curing time, and surcharge.
The most important factor that controls the improvement of the stabilized soil is
the type and the amount of binder. Although significant increase in the
engineering properties of clayey soils can be achieved by treatment with lime,
many researchers observed that cement is the most effective binder for
treatment of organic soils and peats (hnberg, 1996, Tremblay et al, 2002,
Humphrey, 2001, Hebib and Farrell, 2003). The amounts of cement used in
deep mixing are often expressed in terms of the cement factor. This factor
represents the amount of cement (kg) used in a unit volume (1 m3) of the treated
soil. It typically ranges from 100 to 500 kg/m3 for the wet method and from 100 to
300 kg/m3 for the dry method. In the wet method, the cement slurry is prepared
with a water to cement ratio ranging from 0.4 to 2.0, and an amount of slurry
equivalent to 20 to 40 % of the volume of the soil to be treated is injected for
treatment. Since larger amounts of cement are used in wet method, the strength
and the stiffness of soils treated with the wet method are generally higher than
those of soils treated using dry method. The unconfined compression strength of
clays treated with wet method ranges from 0.5 to 5 MPa, with the maximum
strength up to 10 MPa. The typical range of the unconfined compression
strength of clay treated with dry method is between 0.2 and 2 MPa. The stiffness
of the cement treated clays measured through the Youngs modulus (E50) ranges
from about 300 to 1,000 times the unconfined compression strength for the wet
method, while this ratio reduces to about 50 to 200 for the dry method.

35
2.7. Organic soil and cement reactions
The improvement in the engineering properties and the mechanical behavior of
cement treated organic soils are a result of the chemical reactions occurring
between cement and soil. As discussed in the previous sections, organic soils
are highly reactive and very complex in nature. Therefore, it is very important to
understand the nature of the chemical reactions that occur in the organic soil
cement mixture, as they are the main mechanism associated with the
improvement of the engineering properties and the modification of the soil
structure.
Before considering the presence of organic matter it appears useful to review
what is thought to happen (e.g. Aitoh et al. 1985) when cement is used to treat
an inorganic clay.
The chemical reactions of inorganic clays and cement can be divided into three
stages: cement hydration and double layer modification, pozzolanic reaction, and
hardening of cement particles. The schematic of the chemical reaction of cement
and clay minerals is shown in Figure 2.12. The next paragraphs describe the
details of the three reaction stages.
Stage 1: Cement hydration and double layer modification.
Ordinary type I Portland cement is composed of oxides of calcium, silicon,
aluminum, and iron with small amounts of sulfur (or gypsum) and magnesium. In
this section, consistently with cement chemistry notation, the oxides are
expressed as follows:
SiO2 = S,

Al2O3 = A,

CaO = C,

Fe2O3 = F,

and the major cement compounds are described as:


(CaO)3SiO2 ( Tricalcium Silicate) = C3S
(CaO)2SiO2 ( Dicalcium Silicate) = C2S
(CaO)3Al2O3 ( Tricalcium Aluminate) = C3A

H2O = H.

36
(CaO)4(Al2O3)(Fe2O3) ( Tetracalcium Aluminoferrate) = C4AF
As shown in Table 2.12, C3S and C2S are the two main constituents of ordinary
type I Portland cement. The contents of C3S and C2S are about 33 50% and 22
26%, respectively.
When cement particles hydrate, C3S and C2S create a cementitious calcium
silicate gel (C3S2H3) on the surface of cement grains, and release Ca2+ and OHions into solution. The reactions are as follows:
C3S : (CaO)3SiO2 + 6(H2O) = (CaO)3(SiO2)2(H2O)3 + 3Ca2+ + 6(OH)C2S : (CaO)2SiO2 + 4(H2O) = (CaO)3(SiO2)2(H2O)3 + Ca2+ + 2(OH)The reaction of the third constituent of Portland cement, C3A, occurs very rapidly
and causes immediate hardening of the cement (within a few minutes).
According to Jolicoeur and Simard (1998), the relative reactivity of the major
constituents of Portland cement with water is given as C3A > C3S > C2S C4AF.
Thus, the early behavior of hydrating cements is governed mainly by the reaction
of C3A, but setting and early strength development mostly depends on the
hydration of C3S. The reaction of C3A is as follows:
C3A: (CaO)3Al2O3 + 6(H2O) = (CaO)3(Al2O3)(H2O)6
In general, gypsum (CaSO42H2O) is added to cement to inhibit the reaction of
C3A. However, gypsum does not affect the hydration of C3S and C2S (Lea,
1956).
The Ca2+ and OH- ions released into solution increase the ion concentration of
the solution. The released Ca2+ may replace monovalent cation, such as Na+,
which are held on the exchangeable sites of clay minerals. These two
phenomena, in general, reduce the double layer thickness of clay particles.
The double layer thickness of clay particles depends on the type and
concentration of ions near the surface of the clay minerals. Based on Gouy-

37
Chapman or diffuse double layer theory, the thickness of the double layer of a
clay particle (1/k) can be calculated using the following equation:

1 o RT
=
k 2 F 2 I 103

1 2

(2.1)

where 1/k= double layer thickness (m)

= dielectric constant of water (78.5 at 298K)


o = permittivity of free space (8.854 x 10-12 C2J-1m-1)
R = molar gas constant (8.314 Jmol-1K-1)
T = temperature in degrees Kelvin (K)
F = Faraday constant (9.648 x 104 Cmol-1)
I = ionic strength of the solution in moles per decimeter
The ionic strength (I) of the solution can be calculated as:
I=

1 n
Z 2j m j
2 j =1

(2.2)

where Zj = valance of an ion, j


and mj = molar concentration of the ion, j
For example, the release of 0.3 mole of Ca2+ and 0.6 mole of (OH)- ions into 1
liter of deionized water will increase the ionic strength of the solution to :

I=

1
2
2
( +2 )Ca2+ 0.3 + ( 1)OH 0.6 = 0.9 M

The reduction of the double layer thickness of clay minerals causes the soil
particles to form a flocculated structure. However, any increase in clay strength
due to double layer modification is negated by remolding of the clay during the
deep mixing process (Jameson, 1996). In addition, the introduction of additional
water to the clay during mixing stage results in a more flocculated structure in
cement treated clay particles than in the untreated clay particles.

38
Stage 2: Pozzolanic reaction
The hydroxyl ions (OH-) released into the soil solution during hydration of cement
particle increase the pH of the soil solution. If the pH of the soil solution increases
above 12, silica (SiO2) and alumina (Al2O3) of the clay particles are dissolved into
the soil solution. The dissolved silica and alumina react with calcium and
hydroxyl ions to form calcium silicates (CSH) or calcium aluminates (CAH) on the
surface of cement particles. The reactions are as follows:
CSH : Ca2+ + 2(OH)- + SiO2 = CaOSiO2H2O
CAH : Ca2+ + 2(OH)- + Al2O3 = CaOAl2O3H2O
CSH and CAH gels cause interparticle bonding and flocculation of the clay
particles. These cementation reactions are termed pozzolanic reaction. As
shown in the reactions above, the pozzolanic reaction takes hydroxyl ions, and
thus decreases the pH of the solution. This effect will be further discussed in the
following section.
The dissolution of silica and alumina from clay minerals and formation of CSH
and CAH gels between flocculated clay particles lead to an increase in particle
size and create a more open clay structure (Chew et al, 2004). Figure 2.13
shows the change in particle size distribution of a lime treated marine clay from
the coast of Madras, India (Rao and Rajasekaran, 1996). The figure clearly
shows the increase in particle size due to aggregation of particles by pozzolanic
reaction. Figure 2.14 shows the effect of cement treatment on the pore size
distribution of Singapore marine clay (Chew et al, 2004). As shown in the figure,
the cement treated clay matrix has approximately 1 2 orders of magnitude
higher pore size than the untreated clay. This indicates that cement treated soil
has a more open structure than untreated soil.

39
Stage 3: Cement hardening reaction
The calcium silicate gel (C3S2H3) created during hydration of cement particles
forms hydrate crystals as C3S2H3 gel adsorbs pore water, and hardens with time.
The calcium silicate hydrate crystals are needle-shaped. These hydrate crystals
grow and bind adjacent cement or clay particles to form a more rigid structure.
The calcium silicate gel (C3S2H3) and the calcium aluminate hydrates (C3AH6)
are stable only a pH higher than 12.6. If the pH of the soil solution does not
increase above 12.6 during initial hydration, or even if the pH increases above
12.6 but then decreases below this value due to pozzolanic reactions, C3S2H3 will
hydrolyze and transform to CSH gel. The reaction is as follows:
At pH < 12.6: C3S2H3 = CSH and lime
In this case, hardening of C3S2H3 gel and thus increase of the strength of the
soil-cement mixture with time will not occur.
All the three reaction stages contribute to the increase of the strength and the
modification of the structure of soil-cement mixture. The most important stage is
the hardening stage, followed by pozzolanic reaction and double layer
modification.
The reactions described will clearly be modified in presence of organic
compounds. The organic chemicals adsorbed on the surface of hydrated cement
particles can, in fact, interfere with the nucleation and the growth of cementitious
reaction products. While limited work has been done to investigate the interaction
between organic matter and cement during hydration from the perspective of
treating organic soils, significant research has been peformed in the field of
cement and concrete chemistry on the effects of organic chemical compounds on
cement hydration. This research is briefly reviewed below.
In cement chemistry, various types of organic compounds are used as
admixtures to change the physical and chemical properties of cement or to

40
control the hydration of cement (Collepardi, 1995, Ramachandran, 1995).
Organic compounds are generally most widely used as water reducers or
retarders. A water reducer is as admixture that reduces the amount of required
water for cement hydration for a given workability without reducing the strength of
cement. A retarder is an admixture that increases the setting time of cement and
workability time. The main compounds used as water reducers/retarders can be
divided into the following four main groups (Collepardi, 1995):
(1) Ca, Na, NH4 salts of lignosulfonic acids: polymer of a substituted
phenyl propane unit with hydroxyl (OH), methoxyl (OCH3), phenyl ring
(C6H5), and sulfonic acid (SO3H) groups.
(2) Hydroxycarboxylic acids: several hydroxyl groups (gluconic, citric,
tartaric, mucic, malic, salicylic, heptonic, saccharic, and tannic acids)
and carboxylate acids (COOH) attached to carbon chain.
(3) Carbohydrates: glucose, sucrose, or hydroxylated polymers obtained
by partial hydrolysis of polysaccarides.
(4) Other compounds.
Typical amounts of water reducers and retarders used in practice are in the
range of 200 400 ml per 100 kg of cement (Collepardi, 1995). When used
within this dosase, retarders or water reducers only affect the setting of cement
or strength of cement during the first three days of hydration. However, while the
cement mixed with retarders has a lower compressive strength than the
controlled cement prepared at the same water/cement ratio in the first three days
of cement hydration, the compressive strength starts increasing after three days
and becomes higher after 7 days.
Pollard et al (1991) stated that citric acid has the highest affinity to chelate to
aluminum and is a particularly effective retarder at high concentrations, but acts
as an accelerator at low concentrations. Phenol has an adverse effect on longterm strength of cement at volumetric concentrations as low as 0.2%. Phenol

41
substrates are readily leached from the cement matrix, and significantly change
physical characteristics such as setting times, compressive strength, matrix
morphology, and phase composition. However, non-polar organics do not hinder
strength development in cementitious or pozzolanic systems (Pollard et al, 1991).
The reaction between organic compounds and cement has also been widely
studied in the environmental area to understand the mechanisms for solidification
of waste material containing various types of organic matter (Spence, 1993,
Conner, 1990). Organic acids, organic acids chlorides, and phenol generally
have negative effects on the properties of solidified waste-cement system: high
ion exchange capacity of organic acids can inhibit or retard solidification process
(cement hydration) by removing calcium ions from the pore solution (Conner,
1990). Carbonyls have setting and curing retardation effects in the short term.
Chlorinated hydrocarbons have negative effects on the properties of cured
product at lower concentration but short-term setting and curing retardation
effects at higher concentration (Spence, 1990).
Organic admixtures can interfere with cement hydration through physio-chemical
mechanisms including adsorption, nucleation, complexation and precipitation
(Pollard et al. 1993). After the initial hydration of cement, the interaction of
organic admixture and cement hydration products occurs at the solid-solution
interface with a hydrated layer. The organic molecules adsorbed at the solidsolution interface can interfere further with the hydration of cement particles and
can inhibit crystal nucleation and growth.
Organic molecules with charge groups, such as SO3- and COO-, or polar
functional group, such as OH-, have high affinity toward the surface of cement
hydration products. Such organic molecules can be adsorbed on the surface of
cement hydrated products through electrostatic forces and/or hydrogen bonding
and retard hydration of cement particles. In this case, the surface of hydrated

42
cement becomes negatively charged by the adsorbed organic molecules.
Negatively charged cement particles push adjacent cement particles through
electrostatic forces, delaying or retarding formation or development of
cementitious products (Conner, 1990).
Many organic admixtures can help solubilize ionic species such as Ca2+ and Al3+,
through association and complexation. One example of such organic admixtures
is salicylic acid, which is a 2-hydroxybenxoic acid with a chemical formula of
C6H4(OH)COOH. The C3A phase of cement has the highest affinity for adsorption
of salicylic acid. The hydroxyl (-OH) and a carboxyl (-COOH) group forms
complexes with aluminum of C3A, and retard the hydration of C3A. At lower C3A
contents, only small amounts of salicylic acids are adsorbed, and larger amounts
of salicylic acids may affect the hydration of C3S (Ramachandran, 1995).
The low pH of organic acids can cause a negative effect on the cementing
process by preventing the pore water from reaching a sufficiently high pH level to
develop cementing products, such as C3S2H3 gel (Tremblay et al, 2002). For
example, the pH of acetic acid in the pure state is around 4, and thus acetic acid
can neutralize the alkaline pH of cement treated soils and inhibit the formation of
cementing products. Tremblay et al. (2002) observed that acetic acid, humic
acid, tannin acid, EDTA, and sucrose have low pH and are responsible for the
ineffectiveness of the cementing process. Sucrose can be easily adsorbed on
clay particles or on cement particle surfaces and can modify the hydration
surfaces. Sucrose also makes cement constituents soluble and can increase the
formation of CSH. Some organic compounds can delay the hydration and the
setting time of cement but not significantly affect the engineering properties (i.e.
compressive strength) of cement stabilized soils. Such organic compounds
include ethylene glycol, kerosene, and most of the hydrocarbons.

43
The paragraphs above have addressed the potential impact of the presence of
organic compounds on the hydration of cement and on the development of
strength. It is worth considering the reverse problem, i.e. the impact of cement
on the structure and nature of the organic matter. In the case of clay minerals it
was noted above that hydration of cement will lead to a reduction in the double
layer (and thus the generation of a more flocculated structure). The basic
environment (pH>12) created by the hydration of cement is likely, however, to
have a significant impact on the organic matter given that, as discussed in
Section 2.2, certain fractions of organic matter are soluble under alkaline
conditions.

2.8. 1-D consolidation behavior of chemically treated organic soils


Due to the successful implementations in the field and the proved effectiveness
of deep mixing methods as a soil improvement method for soft soils, significant
research has been aimed ta gaining an improved understanding of the physical
properties and the mechanical behavior of chemically stabilized soft soils
(hnberg, 1996; Rao and Rajasekaran, 1996; Tremblay et al, 2001; Holm, 2003;
Jeslic and Leppanen, 2003; Chew et al, 2004; Lorenzo and Bergado, 2004; Kang,
2007). Most of these research efforts have, however, addressed the effects of
the treatment on the strength characteristics, rather than the 1-D consolidation
behavior, often relying simply on measurements of the unconfined compressive
strength. In addition, most of the research has focused on the treatment of
inorganic clays, while data and experience with organic soils and peats has
remained limited.
This section presents a summary of the current knowledge on the 1-D
consolidation behavior of chemically treated organic soils based on the review of
publications which address the treatment of organic soils with a broad range of
characteristics (organic content, and degree of humification): soft and dredged
sediments with only 3-8% OC (Tremblay et al. 2001); three Swedish gyttjas

44
with OC~10% (Ahnberg 2003); two Italian peats with OC in the 60-75% range
and fiber content of 25% (Cortellazzo and Cola 1999); and two predominantly
organic (OC=94-98%) Irish peats, one fibrous, one amorphous (Hebib and Farell
(2003).
As discussed in section 2.8, the engineering properties and the mechanical
behavior of cement treated soil vary depending on the type and the amount of
binder, the characteristics of the soil, and the curing conditions. Therefore, the
discussion that follows provides, when available, details on the materials and
laboratory procedures used in each of the studies examined.
The effects of cement treatment on the 1-D consolidation behavior of soils with
small organic content (LOI = 0 8%) were investigated by Tremblay et al (2001).
The organic soils studied included Saguenay Fjord sediments, Port de Qubec
recent sediments, and dredged material from Qubec harbor. The index
properties of these three soils are summarized in Table 2.13.
To evaluate the effects of organic content on the compression behavior and
treatment with cement, the dredged material was treated with peroxide (H2O2) to
completely remove the organic matter. After treatment, the dredged material was
mixed with the natural dredged material to obtain organic contents in the 0 8%
range. The soils were treated with 5% and 10% of ordinary Portland cement by
the dry mass of soil, and were cured under a surcharge of 8 kPa for 30 to 100
days. This study is of particular interest as it explored soils with organic content
at the low end of the spectrum (it is possible that these materials could not be
recognized as organic based on available engineering classification methods). In
addition, the above described preparation procedures isolates the effect of
organic content, as, for each of the soils tested, the nature of the organic matter
mineral substrate remain constant.

45
The compression curves of the untreated and cement treated dredged materials
are shown in Figure 2.15. Focusing on the curves pertaining to the untreated
soils, it is seen that the compression curves of the dredged material with higher
organic contents lie above those of materials with lower organic content. This is
in agreement with data presented earlier in this chapter (see Figure 2.9).for
fibrous peats from northern Quebec. This effect is quite significant despite the
limited range in OC%. Given the low OC%, the initial void ratios of the dredged
materials are lower than 1.3 just higher that values typically measured on soft
clays, and significantly lower than the values cited earlier for fibrous peats (e.g.
see data for Middleton peat and fibrous peats from northern Quebec in Figure 2.5
and Figure 2.9, respectively). As shown in Figure 2.15, in addition to affecting the
position of the compression curve, the organic content influences also the
compressibility of the soil, with increasing compressibility being measured with
increasing organic content.
The compression curves of the dredged material treated with 10% PC lie above
those of the untreated materials with same organic content. This indicates that
the treated soils can sustain higher vertical effective stress than the untreated
soil at the same void ratio. According to Bergado (1990), this is due to the
structure of the cement treated soils. The cement treated dredged materials with
8% organic content were prepared at two liquidity indexes of 1.0 and 1.8. As
shown in Figure 2.15, the compression curves of the two dredged materials with
8% organic content merged together at a vertical effective stress of
approximately 4000 kPa, leading Tremblay et al. (2001) to conclude that the
virgin compression line of cement treated soil is independent of the initial curing
void ratio. The results presented in Figure 2.15 show that, within the range of
OC% investigated, the position of the virgin compression lines of the cement
treated dredge materials remained the same regardless of the organic content.

46
Tremblay et al. (2001) normalized the compression data shown in Figure 2.15,
for the 10% PC treated dredged materials with organic contents of 0, 3, and 8%
using the normalization method proposed by Burland (1990) (a similar approach
will be taken to analyze the data collected for this thesis in Chapter 6). The
method proposed by Burland (1990) uses a normalization parameter, Iv, called
the void index, which is defined as follows:
Iv =

*
e e100
*
*
e100
e1000

(2.3)

where e = void ratio of the soil


e*100 = void ratio on the intrinsic compression line at vertical
effective stress of 100 kPa
e*1000 = void ratio on the intrinsic compression line at vertical
effective stress of 1000 kPa.
The intrinsic compression line represents the normalized virgin compression line
for the reconstituted soil.
Considering the stress ranges for very soft clayey and organic soils, Tremblay et
al. (2001) modified Burlands void index as follows:
Iv ' =

e e10*
*
e10* e100

(2.4)

where e = void ratio of the soil


e*10 = void ratio on the intrinsic compression line at vertical effective
stress of 10 kPa
e*100 = void ratio on the intrinsic compression line at vertical
effective stress of 100 kPa.
Burlands void index (Iv) and the modified void index (Iv) are related as follows:
Iv ' = Iv 1

(2.5)

47
The compression curves of the 10% PC treated dredged material with organic
contents of 0, 3%, and 8% are presented in Figure 2.16 (a) in terms of modified
void index (Iv) versus vertical effective stress. The compression curves of
cement treated soils shown in this figure were normalized with the intrinsic
parameters (e*10 and e*100) of the untreated soils with the same organic contents,
as shown in Figure 2.15.
The figure shows that the normalized compression curve of the cement treated
dredged material with 3% organic content lies above that of the cement treated
dredged material with 0% organic content. This illustrates the fact that the
cement treated dredged material with 3% organic content can substain higher
vertical effective stress than the cement treated dredged material with 0%
organic content at the same void ratio. The cement treated 3% organic content
soil is also found to be more compressible, although the two compression curves
merge at a vertical effective stress of approximately 4000 kPa. On the other
hand, the normalized compression curve of the cement treated dredged material
with 8% organic content lies below both other curves, i.e. at a lower normalized
void index (or void ratio). Tremblay et al. (2001) suggest that this is a result of the
fact that when the organic content is higher than 3%, the presence of organic
matter interferes with cement hydration. As a result, the structure in the cement
treated dredged material with organic content of 8% is not as stable as the
structures in the cement treated dredged material with lower organic contents. In
addition, compressibility of the compression curve of the cement treated dredged
material with organic content of 8% is approximately the same or slightly lower
than that of the dredged material with 0% and 3% organic contents. The results
generally indicate that the cement treatment does not significantly improve the
compressibility of soils, in the presence of organic matter.
Figure 2.16 (b) presents the normalized compression curves of Louisville
sediments, dredged material, Saguenay Fjord sediments, and Port de Qubec

48
recent sediments with 0, 3%, 6%, and 8% organic contents, respectively, treated
with 5% of cement. The compression curves shown in this figure were
normalized with the intrinsic parameters of the same soil. As shown in the figure,
the compression curves of the soils treated with 5% cement lie above the curves
for the untreated soil, again illustrating the ability of the cement treated soil to
sustain higher void ratio than the untreated soil at the same vertical effective
stress. As observed in Figure 2.16 (a), the curves pertaining to the soils with
higher organic content (6 and 8%) lie below those for the lower OC (0 and 3%).
Based on these results, Tremblay et al. (2001) concluded that below an organic
content of 3 4% cementing clearly occurs, whereas the effect of cement is
smaller at higher percentages of organic content. Figure 2.17 presents a
schematic diagram which synthesizes the conclusions drawn Tremblay et al.
(2001)of the effects of organic content on the compression behavior of cement
treated soils.
The effects of cement treatment on the 1-D compression behavior of amorphous
and fibrous peats were investigated by Hebib and Farell (2003) and Cortellazzo
and Cola (1999).
Figure 2.18 (a) shows the compression curves of two peats obtained from raised
bogs (Raheenmore and Ballydermot) in the Irish Midlands (Hebib and Farell,
2003). The Raheenmore peat (RA) had a natural water content of 1200%,
organic content of 98 99%, and pH of 5.3. It was insignificantly decomposed
and classified as a H2 according to the Von Post classification system. The
Ballydermot peat (BA) had a natural water content of 850%, organic content of
94 98%, and pH of 4.9. The degree of decomposition of this peat was
moderately strong (H6) to nearly complete (H9). As shown in the figure, the
more fibrous Raheenmore peat (RA) exists at a much higher in-situ void ratio
(approximately 19.5) than the more highly decomposed Ballydermot peat (BA)
which had an in-situ void ratio of 11.5 13. In addition, the compression curves

49
of the Raheenmore peat lie above those of the Ballydermot peat at the same
vertical effective stress. The preconsolidation pressures of these two peats were
15 kPa for BA and less than 5 kPa for RA. The compression index (Cc) was
about 6.1 for BA and 6.5 for RA.
The Ballydermot peat was treated with 200 kg/m3 of cement, and cured for 28
and 240 days under a surcharge of 18kPa. The compression curves obtained by
testing the cement treated Ballydermot peat are shown in Figure 2.18 (b),
alongside the data for the untreated soil both in the intact reconstituted soil tested
in the reconstituted stat. The figure shows the following: in the virgin compression
region the curves of the cement treated peat lie above those of the unstabilized
peat and the remolded peat; the initial void ratio decreases with increasing curing
time from 5.9 at 28 days to 4.9 at 240 days; the preconsolidation pressure of the
cement treated peat increases to 210 kPa at 28 days and to 520 kPa at 240
days; the virgin compression curves for the 28 and 240 days cured specimens
are essentially coincident; the virgin compression index of the cement treated
peat is not significantly affected by curing time and exceeds the value measured
on the remolded untreated soil; the compression curves for the cement treated
peat eventually converge to the curve for the remolded peat.
Cortellazzo and Cola (1999) investigated the effect of cement treatment on the 1D consolidation behavior of two peats from North-Eastern Italy: the Adria peat
from the Po delta area and the Correzzola peat from an area between the Adige
and Bacchiglione rivers. The Adria peat had a natural water content (w) of 330
421%, organic content (LOI) of 68 75% and pH of 6.5 7.2, and the Correzzola
peat had w = 606 790%, LOI = 70 72%, and pH = 4.0 5.1.
The Adria peat had about 25% fiber content and was in a moderately strong
degree of decomposition with a Von Post classification designation H6-B2-F3-R1-2W0-N3-A1-P1-pH0-L. The preconsolidation pressure and the compression index of
the Adria peat were approximately 100 kPa and 2.75 4.5, respectively.

50
The fiber content of the Correzzola peat was about 25%, but the Correzzola peat
was slightly less decomposed than the Adria peat with a Von Post classification
designation of H5-B3-F2-R2-W0-N3-A1-P1-pH0-L. The preconsolidation pressure
and the compression index of the Correzzola peat were approximately 15 kPa
and 2.25 7.25, respectively.
Two Adria peat samples, E-2 with a natural water content of 334 % and I-7 with a
natural water content of 418%, and one Correzzola peat sample, S-7 with a
natural water content of 790%, were studied. The Adria peat was treated with
200kg/m3 of cement, and the Correzzola peat was treated with a combination of
150 kg/m3 of cement and 50 kg/m3 of gypsum. The treated peats were immersed
in water and cured under a surcharge of 18 kPa for 28 and 90 days. Figure 2.19
presents the compression curves of untreated and treated Adria and Correzzola
peats. The initial void ratios of the untreated Adria and Correzzola peats were 5
7 and 9.7 11.1, respectively. With treatment, the initial void ratios of both
peats decreased significantly to 3.6 3.8. The preconsolidation pressure
increased from 100 kPa to 150 kPa for Adria peat and from 15 kPa to 90 kPa for
Correzzola peat. Similar to the treatment of the Ballydermot peat the
compression curves of both treated peats converged to that of the untreated soil.
The increase in the preconsolidation pressure of these two peats seems to be
related to the decrease in the void ratio due to mixing and curing rather than to
the structure created by the cement treatment.
Figure 2.20 presents the recompression (Cr) and compression (Cc) indexes of the
untreated and treated Adria and Correzzola peats. Since these values are plotted
as a function of vertical effective stress, a direct comparison (i.e at the same
OCR) can not be made (arrows in the two figures indicate the values of the
preconsolidation stress). According to Cortellazzo and Cola (1999), the Cr and Cc
of the treated Adria peats were about 6.5% and 71% of those of the untreated
peats, and the Cr and Cc of the treated Correzzola peats were about 10% and 58

51
% of those of untreated peats. It should be noted, however, that the compression
indices of both treated peats were the same as those of the untreated peats at
the vertical effective stress of 800 1600 kPa for Correzzola peat. These results
indicate that the cement treatment decrease the compressibility of peats
significantly in the over consolidated region, but slightly or does not affect in the
normally consolidated region.
Figure 2.21 presents the coefficient of consolidation (Cv) of the untreated and
treated Adria and Correzzola peats. The Cv of the untreated peats in the over
consolidated region was in the range of 6 x 10-7 m2/sec (Adria) and 7 x 10-6
m2/sec (Correzzola) but decreased by approximately two orders of magnitude to
4 x 10-8 m2/sec (Arida) and 2 x 10-8 m2/sec with increasing vertical effective
stress. Figure 2.21 shows that in the over consolidated region the Cv of the
treated peats is slightly higher (Adria peat) or slightly lower (Correzzola peat)
than in the case of the untreated peats, up to a vertical effective stress of 100
200 kPa. In the normally consolidated region, however, the Cv values of the
treated peats are approximately 1 2 orders of magnitude higher than those of
the untreated peats at the same vertical effective stress. The coefficient of
consolidation of a soil depends on the hydraulic conductivity and the constrained
modulus. As shown in Figure 2.19 and 2.20, the effect of the treatment on the
compression curves is not very significant. Thus, the increase of Cv with
treatment indicates a significant increase in hydraulic conductivity.
As for natural soils, the hydraulic conductivity of chemically stabilized soil is very
important as it is used for estimation of the degree of consolidation with time.
Although a number of tests have been performed to evaluate the change in
hydraulic conductivity with chemical treatment, the current results show a wide
range, depending on the type and amount of binders and the type of soils
(Terashi and Tanaka, 1983, Locat et al,. 1996, hnberg et al., 1995).

52
The results of tests performed on chemically stabilized inorganic clays generally
show that the hydraulic conductivity of inorganic clays increase when treated with
lime-based binders (hnberg et al., 1995), but decreased when treated with
cement-based binders (Terashi and Tanaka, 1983). Locat et al. (1996) observed
that the hydraulic conductivity of a lime stabilized inorganic clay initially increases
when treated with small amount of lime (less than 2% by the dry mass of the soil),
but decreases with larger amounts of lime.
The effects of treatment on the hydraulic conductivity of organic soils and peats
were investigated by hnberg (2003). The soils studied included a peat with a
water content of 2000%, and three gyttjas with water contents of 220 370%
and organic content of 10 11%. Gyttja is a Swedish word for a nutrient-rich
sedimentary peat deposited in water consisting mainly of plankton, other plant
and animal residues, and mud. The soils were treated with 50 150 kg/m3 of
Portland cement and quicklime, and were cured for 28 days and one year without
applying any surcharge. Figure 2.22 presents the hydraulic conductivity of peat
and gyttja in the natural state (x-axis) and after treatment (y-axis). As shown in
the figure, the hydraulic conductivity of the soils generally decreased after
treatment with cement and lime. However, a slight increase in the hydraulic
conductivity with treatment was also observed on lime and cement stabilized
gyttjas (hnberg, 2003).
Based on the hydraulic conductivity tests performed on chemically stabilized
inorganic clays, gyttjas and peats, hnberg (2003) observed that the change in
the hydraulic conductivity is mainly related to the change in void ratio and the
change in water content resulting from chemical treatment. The relationship
between the change in hydraulic conductivity and the change in water content
with treatment is shown in Figure 2.23. The x-axis represents the ratio of the
water content of a soil after treatment (w) and in its natural state (wo). The y-axis
represents the ratio of the initial hydraulic conductivity of the soil after treatment

53
(Kstab, initial) and in its natural state (Ksoil). The figure shows that if the water
content of a soil decreased to about 53% of its natural water content after
treatment, the hydraulic conductivity of the soil decreases with treatment. It
should be noted, however, that the results shown in Figure 2.23 include various
types of soils (inorganic clays, gyttjas, and peats) treated with either cement or
lime with varying amounts of binders.
Based on the results of consolidation tests performed on soils stabilized with
cement, the effects of treatment on the 1-D consolidation can be summarized as
follows. The compression curves of soils with low organic content generally shift
to higher vertical effective stress when treated with cement, as shown in Figure
2.16 (a) and (b). The compression curves of the cement treated soils with higher
organic content generally lie at lower vertical effective stress than those of soils
with lower organic contents, as shown in Figure 2.17. This result implies that the
organic content is a very important factor affecting the improvement of the
mechanical behavior of organic soils with treatment, and that the effectiveness of
the treatment is inversely related to the organic content.
The consolidation tests results obtained for cement treated organic soils and
peats indicate that cement treatment does not significantly improve the 1-D
compression behavior of soils with organic content higher than 70%, at least up
to a cement factor of 200 kg/m3. As shown in Figure 2.18, the virgin compression
lines of cement treated Ballydermot peat does not show any shift with respect to
that of the remolded soils. Figure 2.19 also showed that the virgin compression
lines of Adria peat and Correzzola peat remained the same even when treated
with 200kg/m3 of cement. Figure 2.20 showed that while the compressibility of
peats decreased significantly in the over-consolidated region with cement
treatment, the compressibility in the normally consolidated region remained the
same. While cement treatment does not greatly modify the compressibility, it
does lead to an increase in the preconsolidation stress. However, this increase

54
appears to be due primarily to the decrease in the initial void ratio associated
with the mxing and curing operations. Overall, these results suggest that if the
organic matter is higher than 70%, the cement hydration. may be retarded and
even totally inhibited. This is confirmed by the fact that for peats with organic
content higher than 70%, the effects of curing time on the compression behavior
are almost negligible.
The values of the coefficient of consolidation measured on two treated Italian
peats generally indicate that this soil property increases with cement treatment,
and that this result is a reflection primarily of the increase in hydraulic
conductivity with treatment. The observations above imply that while the
mechanical behavior may not significantly affected, the soil structure (particle and
pore size), change as a result of the treatment.
With regards to the effects of hydraulic conductivity results opposite to those
discussed for the two Italian peats are reported by hnberg (2003). These
conflicting results can be interpreted based on the change in water content with
treatment.

55
Table 2.1: Major characteristics of 12 soil orders in Soil Taxonomy (after Brady
and Weil, 1999)
Name

Major Characteristics

Alfisols

Argillic, nitric, or kandic horizon; high to medium base saturation

Andisols

From volcanic ejecta, dominated by allophone or Al-humic complexes

Aridisols

Dry soil, ochric epipedon, sometimes argillic or nitric horizon

Entisols

Little profile development, ochric epipedon common

Gelisols

Permafrost, often with cryoturbation (from churning)

Histosols

Peat or bog; > 20% organic matter

Inceptisols

Embryonic soils with few diagnostic features, ochric or umbric epipedon, cambic
horizon

Mollisols

Mollic epipedon, high base saturation, dark soils, some with argillic or nitric horizons

Oxisols

Oxic horizon, no argillic horizon, highly weathered

Spodosols

Spodic horizon commonly with Fe, Al oxides and humus accumulation

Ultisols

Argillic or kandic horizon, low base saturation

Vertisols

High in swelling clays, deep cracks when soil dry

Table 2.2: Distribution of soil organic matter in Florida Histosols (after Zelazny
and Carlisle, 1974)
Series
Montverde

Pahokee
Okeetanta
Torry

Depth

Organic fraction

(cm)

Humin

Humic acid

Fulvic acid

HA/FA

0-13

75.4

13.3

11.3

1.2

13-36

67.9

15.3

16.9

0.9

36-81

86.2

11.0

2.9

3.8

0-18

57.0

7.2

35.8

0.2

18-86

66.8

12.2

21.1

0.6

0-21

58.7

22.4

18.9

1.2

21-66

70.1

12.5

17.4

0.7

0-21

55.4

37.7

6.9

5.4

21-53

35.1

54.6

10.3

5.3

53-109

65.3

19.6

15.1

1.3

109-137

78.4

16.1

5.5

3.0

56
Table 2.3: Elemental composition of humic acids extracted from soils from widely
different climates (after Schnitzer and Khan, 1978)
Element
(%)

Arctic

Cool, temperate
Acid soils

Neutral soils

Subtropical

Tropical

Carbon

56.2

53.8-58.7

55.7-56.7

53.6-55.0

54.4-54.9

Hydrogen

6.2

3.2-5.8

4.4-5.5

4.4-5.0

4.8-5.6

Nitrogen

4.3

0.8-2.4

4.5-5.0

3.3-4.6

4.1-5.5

Sulfur

0.5

0.1-0.5

0.6-0.9

0.8-1.5

0.6-0.8

Oxygen

32.8

35.4-38.3

32.7-34.7

34.8-36.3

34.1-35.2

Table 2.4: Elemental composition of fulvic acids extracted from soils from widely
different climates (after Schnitzer and Khan, 1978)
Element
(%)

Arctic

Cool, temperate
Acid soils

Neutral soils

Subtropical

Tropical

Carbon

47.7

47.6-49.9

40.7-42.5

42.2-44.3

42.8-50.6

Hydrogen

5.4

4.1-4.7

5.9-6.3

5.9-7.0

3.8-5.3

Nitrogen

1.1

0.9-1.3

2.3-2.8

3.1-3.2

2.0-3.3

Sulfur

1.6

0.1-0.5

0.8-1.7

2.5

1.3-3.6

Oxygen

44.2

43.6-47.0

47.1-49.8

43.1-46.2

39.7-47.8

57
Table 2.5: Elemental composition of Humic Acids (HA) and Fulvic Acids (FA)
from various tropical and temperate regions soils (after Tan, 2003)
Carbon

Hydrogen

Oxygen

Nitrogen

(%)

(%)

(%)

(%)

C/N

Tropical Region Soils


HA Alfisols

52.3

5.2

37.2

3.6

14.5

HA Andosols

58.5

3.9

34.8

3.8

15.4

HA Oxisols

54.5

4.4

38.0

3.1

17.5

FA - Andosols

48.9

4.3

44.5

2.3

21.3

Temperate Region Soils


HA Alfisols

56.8

5.0

33.6

4.6

12.3

HA Aridisols

54.5

4.1

36.4

5.0

10.9

HA Histosols

58.7

5.0

32.9

3.4

17.3

HA Inceptisols

51.4

5.4

39.1

4.1

12.5

HA Mollisols

53.7

4.3

36.3

3.7

14.5

HA Spodosols

53.1

4.9

39.2

3.5

15.2

HA Ultisols

50.5

5.2

40.0

3.9

12.9

FA Inceptisols

47.9

5.2

44.3

2.6

18.4

FA Mollisols

41.6

4.0

51.9

1.1

37.8

FA Spodosols

50.6

4.0

44.1

1.8

28.1

FA - Ultisols

45.1

4.5

48.0

2.2

20.5

Reference humic acid

56.2

4.7

35.5

3.2

17.6

Peat

55.5

5.4

33.1

3.0

18.5

58
Table 2.6: Elemental composition of humic substances of Florida Histosols (after
Zelazny and Carlisle, 1974)
Series

Montverde

Pahokee
Okeetanta
Torry

Depth

Humin

Humic Acid

Fulvic Acid

(cm)

0-13

56.9

4.4

3.1

0.2

35.4

56.8

5.0

3.4

0.3

34.5

53.3

4.2

2.3

1.9

38.3

13-36

56.2

4.2

3.0

0.2

36.4

56.2

4.6

3.4

0.3

35.5

49.6

6.4

2.4

1.4

40.2

36-81

56.8

6.8

3.1

0.3

33.0

58.8

4.4

3.8

0.1

32.9

51.2

6.4

2.6

1.1

38.7

0-18

56.1

4.8

3.1

0.1

35.9

56.3

4.0

3.6

0.1

36.0

54.8

4.6

1.5

0.2

38.9

18-86

58.5

4.2

2.3

0.2

34.8

56.0

5.2

2.6

0.3

35.9

55.3

4.4

1.3

0.3

38.9

0-21

56.4

4.4

2.7

0.3

36.2

56.6

3.8

3.1

0.3

36.2

56.6

4.6

2.2

0.6

36.0

21-66

57.5

4.8

2.6

0.3

34.8

57.8

4.2

3.3

0.4

34.3

59.6

4.4

2.5

0.3

33.2

0-21

60.7

4.4

3.4

0.5

31.0

56.6

4.4

3.9

0.6

34.5

46.9

7.2

1.2

0.4

44.3

21-53

61.6

4.2

3.4

0.3

30.5

58.0

4.0

3.8

0.3

33.9

52.2

6.8

1.2

0.4

39.4

53-109

58.2

4.6

2.6

0.5

34.1

56.6

5.0

2.8

0.4

35.2

60.8

5.4

2.1

0.6

31.1

109-137

56.6

5.2

2.7

1.1

34.4

57.4

5.2

3.0

0.7

33.7

59.4

4.2

1.6

1.3

33.5

59
Table 2.7: Distribution of oxygen-containing functional groups in humic and fulvic
acids extracted from soils of widely different climatic zones (all numbers in
meq/100g) (after Stevenson, 1994)
Functional Group

Arctic

Cool Temperate
Acidic Soils

Neutral Soils

Subtropical

Tropical

Range

Humic Acid
Total acidity

560

570-890

620-660

630-770

620-750

560-890

COOH

320

150-570

390-450

420-520

380-450

150-570

Acidic OH

240

320-570

210-250

210-250

220-300

210-570

490

270-350

240-320

290

20-160

20-490

10-180

450-560

80-150

30-140

10-560

40

30

30-50

60-80

30-80

Weakly acidic +
alcoholic OH
Quinone C=O

230

Ketonic C=O

170

OCH3

40

Fulvic Acid
Total acidity

1100

890-1420

640-1230

820-1030

640-1420

COOH

880

610-850

520-960

720-1120

520-1120

Acidic OH

220

280-570

120-270

30-250

30-570

380

340-460

690-950

260-520

260-950

Weakly acidic +
alcoholic OH
Quinone C=O

200

Ketonic C=O

200

OCH3

60

170-310
30-40

120-260

80-90

30-150
160-270
90-120

120-420
30-120

Table 2.8: Concentration of organic functional groups in humic substances from


11 Florida Histosols samples (average and standard deviation values expressed
in cmol/kg) (after Sparks, 2003)
Total acidity

Carboxyls

Phenolic OH

Alcoholic OH

Carbonyls

Humins

510 20

200 20

310 20

360 30

260 20

Humic acids

720 40

310 20

420 30

130 30

130 10

Fulvic acids

860 40

400 20

460 20

80 20

430 10

60
Table 2.9: Relative values of various peat properties for amorphous-granular,
fine-fibrous and coarse-granular peat (after MacFarlane and Williams, 1974)
Natural
Water
Unit weightShear strengthCompressiblity
Permeability
Void ratio
content
Amorphous - granular
3
3
2
1
3
2
Fibrous fine
1
2
1
3
2
1
(woody and nonwoody)
Fibrous course
2
1
3
2
1
3
(woody)

On relatively scale, with 1 being the greatest and 3 being the least

Table 2.10: Values of C/Cc for various types of geotechnical materials (Terzaghi
et al, 1996)
Material

C/Cc

Granular soils including rockfill

0.02 0.01

Shale ad mudstone

0.03 0.01

Inorganic clays and silts

0.04 0.01

Organic clays and silts

0.05 0.01

Peat and muskeg

0.06 0.01

61

Table 2.11: Values of natural water content (wo), permeability, and C/Cc for
various peat deposits (after Mesri et al, 1997)
Natural water
Peat

content
(%)

Permeability
(m/s)

C/Cc

Reference

Fibrous peat

850

4x10-6

0.06 0.10

Hanrahan (1954)

Peat

520

0.061 0.078

Lewis (1956)

500 1500

10-7 10-6

0.035 0.083

Rea and Brawner (1963)

200 600

10-5

0.09 0.10

Adams (1965)

705

0.073 0.091

Amorphous and
fibrous peat
Canadian muskeg
Amorphous to
fibrous peat

-3

0.075 0.085
0.052 0.072

Peat

400 750

10

Fibrous peat

605 1290

10-6

Fibrous peat

-6

613 886

10 10

600

Fibrous peat

660 1590

Dutch peat

370

Amorphous to
fibrous peat

Fibrous peat

610 850

-3

Keene and Zawodniak


(1968)
Weber (1969)
Samson and Larochell
(1972)

0.06 0.085

Berry and Vickers (1975)

10-6

0.042 0.083

Dhowian and Edil (1981)

5x10-7 5x10-5

0.06

Lefebvre et al. (1984)

0.06

Den Haan (1994)

0.052

Mesri et al. (1997)

-8

6x10 10

-7

Table 2.12: Composition of Ordinary Type I Portland cement manufactured in the


States (Lea, 1956)
Free CaO (%)

C3S (%)

C2S (%)

C3A (%)

C4AF (%)

0.4

50

22

12

0.2

42

32

1.6

50

26

10

2.0

33

36

14

62
Table 2.13: Index properties of Saguenay Fjord sediments, Port de Qubec
recent sediments, and dredged material from Qubec harbor (Tremblay et al,
2001)
Characteristic

Saguenay Fjord

Port de Qubec

Dredged material

Liquid limit (%)

86

65

60

Plastic limit (%)

34

34

Nonplastic

Plasticity index (%)

52

31

Organic content (%)

Specific gravity

2.75

2.67

2.65

Clay fraction (<2m)

62

12

63

Entisols
(recent)

Mild weathering on
volcanic ejecta

Andisols
(volcanic
Ejecta)

Deep accumulation
of organic materials,
wet conditions
Very cold,
permafrost

Mild weathering,
various conditions
Inceptisols
(beginning B)

Histosols
(organic)

Gellisols
(permafrost)

Desert shrubs,
grasses, dry

High base status,


high-activity clays,
dry season
Vertisols
(swelling
clays)

Andisols
(dry)

Moist, mildly acid


clay accumulation

Semiarid to moist
grassland, mollic
epipedon
Mollisols
(soft, dark)

Alfisols
(mildy acid
clays)
Wet tropical and subtropical
forests, acid silicate and Fe, Al
oxides

Cool, wet, sandy acid,


corniferous forest
Spodosols
(spodic
horizon)

Ultisols
(strongly
acid clays)

Wet, tropical forest, extreme


weathering, low activity clays,
Fe, Al oxides
Oxisols
(oxides)

Figure 2.1: General degree of weathering, climatic and vegetative conditions of


12 soil orders in Soil Taxonomy (after Brady and Weil, 1999)

64

Soil organic matter (Humus)

Nonhumic substances

carbohydrates

Humic substances

lipids

Fulvic acid

amino acids

Humic acid
Humin

Figure 2.2: Classification of soil organic matter (humus)

Compression index, Cc

10

Clay and Silt Deposits


Peats

0.1
10

100

1000

Initial water content, Wo (%)

Figure 2.3: Initial water content (wo) and compression index (Cc) of peats and soft
clay and silt deposits (after Mesri et al, 1997)

65

12
Fibrous Peat

Void ratio, e

10
8
6
4
2
0
100

Amorphous
Granuar Peat
101

102

103

Vertical effective stress, 'v (kPa)


12

Void ratio, e

10
Fibrous Peat

8
6
4

Amorphous
Granuar Peat

2
0
10-11

10-10

10-9

10-8

10-7

10-6

Hydraulic conductivity, k (m/sec)

Figure 2.4: Comparison of compression curves and curves of hydraulic


conductivity versus void ratio of fibrous and amorphous granular peat (after Berry
and Poskitt, 1972)

66

16
IL compression curves
Middleton Peat

14

Void ratio, e

12
10
8
6
4
2
0

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 2.5: Range in compression behavior of intact Middleton peats (after Mesri
et al, 1997)

67

16
Kv (vertical)

14

Kh (horizontal)
Kv (vertical) from CRS test

Void ratio, e

12
10
8
6
4
2
0
10-13

10-12

10-11

10-10

10-9

10-8

10-7

10-6

10-5

Hydraulic conductivity, k (m/sec)

Figure 2.6: Hydraulic conductivity of Middleton peat (after Mesri et al, 1997)

68

10
Clay and Silt Deposits
Peats

Ck = e/logk

Ck = eo/2

Ck = eo/4
0

10

15

20

25

Initial vodi ratio, eo

Figure 2.7: Comparison of Ck for peat and soft clay and silt deposits (after Mesri
et al, 1997)

69
0.05

C /(1+eo)

0.04
0.03
0.02

C/Cc = 0.052

0.01
0.00
0.0

0.2

0.4

0.6

0.8

1.0

Cc /(1+eo)

Figure 2.8: C/Cc for Middleton peat (after Mesri et al, 1997)

70

30

30
NB2-81-COE-04-U-GP-K
NB3-79-COE-04-K

25

25

20

20

Void ratio, e

Void ratio, e

NB2-80-COE-07
NB2-81-COE-04-U-GP-K
NB3-79-COE-04-K

15

15

10

10

10

100

Vertical effective stress, 'v (kPa)

0
1e-9 1e-8

1e-7 1e-6

1e-5

1e-4

1e-3

1e-2

Hydraulic Conductivity, k (cm/sec)

Figure 2.9: Compression curves and curves of hydraulic conductivity versus void
ratio for fibrous peats from northern Quebec, Canada (after Lefebvre et al, 1984)

71

0.9

Coefficient of secondary compression, C

0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
Compression index, Cc

Figure 2.10: C/Cc for fibrous peats from northern Quebec, Canada (after
Lefebvre et al, 1984)

72

Deep Mixing Methods

Dry (D)

Binder Type

Rotary + Jet (J)

Rotary (R)

Mixing Principle

End (E)

End (E)

Slurry (W)

Rotary (R)

Shaft (S)

End (E)

Mixing Location

DSM
(Geocon)

CDM and FGC-CDM


(Ratio)

SWING
(Raito)

DJM
(Raito)

SMW
(SMW Seiko/Raito)

SSM
(Geocon.)

GEOJET
(Condon Johnson)

Lime-Cement Columns
(Underpinning)

Multimix
(Trevisani)

SCC/Geocolumns
(SCC Technology)

HYDRAMECH
(Geocon)

Trevimix
(Trevi ICOS)

Mixed Wall
(Schnabel Foundation)

MECTOOL
(Millgard)

TurboMix
(Trevi ICOS)

KiMix
(Trevi ICOS/Hercules)

Hayward Baker Method


(Hayward Baker)

Rotomix
(Inquip)

Figure 2.11: FHWA classification of Deep Mixing Methods (after Bruce and Bruce,
2003)

73

Clay
Clay particles

Cement
(C3S, C2S, etc)

Pore water

Double layer
modification

Ca2+ + OH-

Water

Cement hydration
= C3S2Hx + Ca(OH)2

Increase in pH

Dissolution of
silica and Alumina

Pozzolanic
Reaction

Flocculation

Secondary
cementitious
products
(CAH & CSH)

Primary
cementitious
products
(hydrated gel
C3S2Hx)

Cement hardening
(pH > 12.6)

Aggregation of
clay particles and
cementation

Figure 2.12: Schematic diagram of inorganic clay - cement reactions (after Saitoh
et al., 1985)

74

100

Percent finer (%)

80

60

40
Untreated
Lime injection treated
Lime column treated (sea water)
Lime column treated (fresh water)

20

0
0.001

0.01

0.1

Particle size, (mm)


Figure 2.13: Increase in particle size due to pozzolanic reaction (after Rao and
Rajasekaran, 1996)

75

100
90

Pore volume finer (%)

80
70
60
50
40
30

Untreated
10% cement
30% cement
50% cement

20
10
0
0.001

0.01

0.1

10

100

1000

Entrance pore diameter (m)


Figure 2.14: Effect of cement treatment on pore size distribution of Singapore
marine clay (after Chew et al, 2004)

76

2.3
0 % OC (Untreated)
3 % OC (Untreated)
8 % OC (Untreated)
0 % OC (10% PC)
3 % OC (10% PC)
8 % OC (LL = 1.0, 10% PC)
8 % OC (LL = 1.8, 10% PC)

2.1
1.9

Void ratio, e

1.7
1.5
1.3
1.1
0.9
0.7
0.5
0.3

10

100

1000

10000

Vertical effective stress, 'v (kPa)


Figure 2.15: Compression curves of dredged materials untreated and treated
with 10% Portland cement (after Tremblay et al, 2001)

77

6
(a) 10% cement
4

Iv'

-2

-4

0 % OC
3 % OC
8 % OC
100

1000

10000

Vertical effective stress, 'v (kPa)


6
(b) 5% cement
4

Iv'

0
Louiseville (0 % OC)
Dredged naterial (3 % OC)
Saguenay (6 % OC)
Port de Quebec (8 % OC)

-2

-4

10

100

Untreated
1000

10000

Vertical effective stress, 'v (kPa)


Figure 2.16: Average normalized compression curves of: (a) dredged material
treated with 10% Portland cement, and (2) organic soils treated with 5% Portland
cement (after Tremblay et al, 2001)

78

6
Inorganic soils
(OC less than 3%)

Iv'

4
Increasing Orgacnic
Matter Content

-2

Organic soils
(OC more than 3%)

10

100

1000

Vertical effective stress, 'v (kPa)


Figure 2.17: Schematic diagram showing the effect of organic matter content on
the position of the normalized compression curves (after Tremblay et al, 2001)

79

20

(a) Untreated

18
16

Void ratio, e

14
12
10
8
6

BA peat : 1.00 m
BA peat : 0.60 m
BA peat : 0.25 m
RA peat : 0.20 m

4
2

10
100
Vertical effective stress, 'v (kPa)

1000

6.5

(b) Treated

6.0

Void ratio, e

5.5
5.0
4.5
4.0
3.5
3.0
2.5
10

remolded slurry
remolded
28 days
240 days

100
1000
Vertical effective stress, 'v (kPa)

Figure 2.18: Compression curves of (a) untreated Raheenmore peat (RA) and
Ballydermot peat (BA), and (b) Ballydermot peat (BA) treated with 200 kg/m3 of
cement (after Hebib and Farrell, 2003)

80

6.5

6.5
E-2 (untreated)
I-7 (untreated)

6.0

6.0

E-2 (28d treated)


I-7 (90d treated)

5.5
5.0

5.0

4.5

4.5

4.0
3.5

4.0
3.5

3.0

3.0

2.5

2.5

2.0

2.0

1.5

1.5

(a) Adria Peat


1.0

10

100

S3 (28d treated)
S3 (90d treated)

5.5

Void ratio, e

Void ratio, e

S3 (untreated)

(b) Correzzola Peat


1000

Vertical effective stress, 'v (kPa)

1.0

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 2.19: Compression curves of untreated and cement treated (a) Adria and
(b) Correzzola peats (after Cortellazzo and Cola, 1999)

81

7.5

5.0
4.5
4.0

E-2 (untreated)
I-7 (untreated)
E-2 (28d treated)
I-7 (90d treated)

6.5

S3 (28d treated)
S3 (90d treated)

5.5

Compression index, Cc

Compression index, Cc

S3 (untreated)

6.0

3.5
3.0
2.5
2.0
1.5

5.0
4.5
4.0
3.5
3.0
2.5
2.0
1.5

1.0

1.0

0.5

(a) Adria Peat


0.0
10

7.0

100

1000

Vertical effective stress, 'v (kPa)

0.5
0.0
10

(b) Correzzola Peat


100

1000

Vertical effective stress, 'v (kPa)

Figure 2.20: Effects of treatment on the compressibility of cement treated (a)


Adria and (b) Correzzola peats (after Cortellazzo and Cola, 1999)

82

S3 (untreated)
S3 (28d treated)
S3 (90d treated)

1e-5

Coefficient of consolidation, Cv (m /sec)

E-2 (28d treated)


I-7 (90d treated)

1e-5

Coefficient of consolidation, Cv (m /sec)

E-2 (untreated)
I-7 (untreated)

1e-6

1e-7

1e-8

1e-6

1e-7

1e-8

(b) Correzzola Peat

(a) Adria Peat


1e-9

10

100

1000

Vertical effective stress, 'v (kPa)

1e-9
10

100

1000

Vertical effective stress, 'v (kPa)

Figure 2.21: Effects of treatment on the coefficient of consolidation of (a) Adria


and (b) Correzzola peats treated with 200kg/m3 of cement (after Cortellazzo and
Cola, 1999)

83

1e-6

Kstabilized soil (m/sec)

1e-7

peat treated with cement


gyttjas treated with cement
gyttjas treated with lime, 28 days
gyttjas treated with lime, 1 year

1
1:

1e-8

1e-9

1e-10

1e-11
1.0e-10

1.0e-9

1.0e-8

1.0e-7

1.0e-6

Kunstabilized soil (m/sec)

Figure 2.22: Ranges in hydraulic conductivity measured on soils treated with lime
and cement (after hnberg, 2003)

84

100

Kstab, initial/Ksoil

10

Dolme Peat, 18kPa


Dolme Peat, 0kPa
Dolme gyttja
Lovstad gyttja
Fiskvik clay
Mellosa clay
Compacted clay

y = 0.043e

6.0048x

0.1

0.01
0.0

0.2

0.4

0.6

0.8

1.0

Wstab, initial/Wsoil
Figure 2.23: Relationship between the change in permeability and change in
water content of stabilized soils (after hnberg, 2003)

85

CHAPTER 3. LINDBERG ROAD SOIL

3.1. Introduction
This chapter presents the engineering classification and characterization of the
organic soil used in this research. This includes: water content, organic content,
specific gravity, particle size distribution, Atterberg limits, fiber content, soil
acidity, soil taxonomy, mineralogical composition of the inorganic fraction of the
soil and chemical composition of the organic fraction. This is done for the
following reasons: (1) gain an in-depth knowledge and understanding of the
characteristics of the soil; (2) compare the soil used in the research with other
organic soils based on their engineering characteristics and chemical
composition; and (3) link engineering classification and chemical composition of
the soil with its consolidation behavior in its natural state and treated with cement
(these two issues are fully discussed in Chapters 5 and 6).
The soil studied in this experimental program was obtained from the edge of the
Celery bog on the northern side of Lindberg road, in West Lafayette, Indiana,
thus is herein referred to as the Lindberg Road (LR) soil. The sampling site was
approximately 0.8 km west of the intersection of Lindberg road and Northwestern
Avenue.
The subsurface soil profile, obtained from the soil borings drilled by Earth
Exploration, Inc. (1996) along the center line of Lindberg road can be
summarized as follows: about 1 1.5m thick crushed limestone fill layer was
found beneath the pavement of the roadway. The crushed limestone was
underlain by a second fill layer described as silty clay or clay loam with

86
occasional sand seams, down to a depth between 1.37 and 2.74m. Below the
second fill layer, about 3.0 3.7 m thick highly organic soil layer, the soil
employed in this experimental program, was encountered. The highly organic soil
was underlain by a 1.4 1.7 m thick marl layer. The marl layer was underlain by
silty clay, clay loam and loam soils down to the maximum depth explored in the
borings. The highly organic soil and the marl layers were encountered only in the
boreholes drilled within the bog. The groundwater table observed after
completion of the drilling was about 3 m below the ground surface.
For this research, two different types of soil samples were obtained from the site:
intact block samples and disturbed soil samples. The details of the sampling
procedure were described by Humphrey (2001). Two block samples were
collected in April of 2001, employing a box sampler, specially designed for easy
sampling with minimal sample disturbance. The sampler was made of 1.3 cm
thick Lexan plates with dimensions 30 cm x 30 cm x 30 cm (Figure 3.1). The
block samples were collected at a depth of 1.7m from an excavation of about
4.6m in diameter. The sampler was pushed down by hand while the soil was
trimmed away. The bottom edge was beveled so that the sampler would cut
through the soil. The air inside the sample was allowed to escape during
sampling through holes located on the top. Once the sampler was filled, a cutting
blade was inserted at the bottom into the box to separate the sample from the
ground. After sampling, the sampler filled with soil was stored in a humid room
where relative humidity and temperature were kept constant at 95% and 10 C,
respectively. Approximately one month after sampling, the box sampler was
dismantled and the block sample was removed. The block was cut into smaller
pieces with a wire saw and a sharp knife blade. The samples were covered with
two layers of plastic warp and one layer of wax to prevent any moisture loss
(Figure 3.2). All the block samples were stored in the humid room

87
Disturbed samples were collected in April and July of 2001. The disturbed
samples collected in April were obtained from the same excavation where the
block samples were obtained. The July samples were obtained in the close
proximity of the April sampling site from a six-meter diameter hole excavated to a
depth of approximately two meters. The disturbed samples were kept in plastic
bins and stored in the humid room until testing.

3.2. Characterization and classification of LR soil


The index properties of organic soils lie in a very wide range as shown in Table
3.1 which summarizes index properties and unit weights of various types of peats
and organic soils collected in Wisconsin and Minnesota. In the table, soils with
organic content greater than 30% are classified as peats, and peats with more
than 30% fiber content are classified as fibrous peats. Peats have the natural
water content ranging from 150 to 655%, and specific gravity ranging from 1.40
to 2.23. Organic soils have the natural water content ranging from 50 to 367%,
and the specific gravity ranging from 2.29 to 2.63.
In this section, the index properties of LR soil and the methods for determination
of each index property are presented.
3.2.1. Water content
The water content of the LR soil was obtained in accordance with ASTM D2974.
The soil samples were kept in an oven at a temperature of 105 5 C for 24
hours. Note that British Standard (1377, Test 1(A)) recommends a drying
temperature of 60 C for peats and soils containing organic matter to prevent
oxidation of the organic content. The natural water content of the LR soil was
obtained with disturbed samples collected in sealed glass jars. The natural water
content of the LR soil showed high variation with average values in the range of
289 % 25 % for April samples and 188 % 31 % for July samples (Humphrey,
2001). Since the intact block samples were collected in April, the water content of

88
289% is taken as the in-situ natural water content. The water contents of the LR
soils employed in this experimental program are summarized in Table 3.2.
3.2.2. Organic content
The organic content of LR soil was taken as the loss on ignition in accordance
with ASTM D2974. The loss on ignition of each soil sample was measured during
or prior to the sample preparation procedure (see section 4.3 for details).
Following the water content determination, the oven dried soil specimen was
placed in a muffle furnace and was subjected to a temperature of 440 C until the
mass of the sample did not change upon further heating. In general, the samples
were kept in the furnace for three days to ensure that all the organic matter was
completely removed.
The average values of the loss on ignition of the block and of the disturbed
samples collected in April and July are summarized in Table 3.2. The loss on
ignition values showed high variation between samples. The block samples had
the highest value of loss on ignition, with an average value of 60.1 4.9 %. The
disturbed LR soil collected in April had slightly higher loss on ignition and also
higher variation than the soil collected in July. Note that the ASTM standard
(D4427) classifies a soil with more than 75% organic matter as peat. The Indiana
Department of Transportation (INDOT) standard classifies a soil with organic
content between 19 30% as organic soil, and a soil with more than 30%
organic content as peat.
3.2.3. Specific gravity
The specific gravity is essential for phase relationships and calculations of initial
void ratio. The specific gravity of most inorganic soils generally lies between 2.65
and 2.85. For soils containing substantial quantities of organic matter, the
specific gravity is considerably lower, usually below 2.0 (Table 3.1).
The specific gravity of LR soil was measured in accordance with ASTM D854.
For soils having solids that can dissolve or float in water, the ASTM standard

89
recommends using Gas pycnometer (ASTM D5550) or kerosene as the fluid.
However, LR soil did not contain any material that floats in water, so the test was
performed with deionized water. In the test, LR soil was de-aired by applying
both heat and vacuum. The pycnometer containing the soil sample was heated in
a water bath and vacuum was applied through a glass tube placed in the rubber
stopper. Using only vacuum was not sufficient and the test yielded values
significantly lower than the values measured using both heat and vacuum. During
the deairing process, the soil was agitated often to make sure that small air
bubbles could be removed easily and the soil would not stick to the wall of the
pycnometer.
Figure 3.3 presents the specific gravity and the loss on ignition of LR soil. The
results show that the specific gravity and the loss on ignition of LR soil are
linearly related through the equation:
Gs = -0.0104 x LOI (%) + 2.570

(3.1)

The linear relationship is more apparent for LR soil with loss on ignition lower
than 50%. This result is very similar to the relationship between specific gravity
and organic content of muskeg, as shown in Figure 3.4. The equation for the
best-fit of the data points shown in this figure is:
Gs = -0.012 x OC (%) + 2.7

(3.2)

The initial void ratios of LR soil specimens employed for CRS and IL
consolidation tests were calculated estimating the specific gravity with the loss on
ignition using equation (3.1).

3.2.4. Particle size distribution


In general, the particle size distribution of a soil is obtained by mechanical sieving
after completely drying the soil. However, this method can cause significant
errors for organic soils since organic matter can shrink or break when dried. To
prevent such problems, the particle size distribution of LR soil was obtained
through a combination of wet sieving and hydrometer test, in accordance with

90
ASTM D422. The size distribution of LR soil particles smaller than 0.075 mm was
obtained from the hydrometer test. The same soil samples employed for the
hydrometer test were washed with water (wet-sieving) to obtain the size
distribution of particle bigger than sand-size.
Figure 3.5 presents the particle size distribution of LR soil obtained by combining
results from wet sieve and hydrometer analysis. The test was completed after
four days when the hydrometer reading did not change further. About 28% of LR
soil particles were smaller than 0.8 m. The result indicates that Lindberg Road
soil contains approximately 4.9 % sand (>0.075 mm), 53.2% silt (> 0.075 and >
0.002 mm) and 41.5% clay (> 0.002mm) as defined by USCS.

3.2.5. Atterberg limits


The Atterberg Limits provide a useful method for identifying and classifying finegrained soils. The liquidity index describes the relative consistency of finegrained soils on the basis of the natural water content. The shear strength and
the compressibility of fine-grained soils also depend on the relative consistency,
or liquidity index (LI), of the soil.
The Atterberg limits of LR soil were measured in accordance with ASTM D4318.
The soil was treated with deionized water to 15 drop consistency prior to testing
(wet method). When tap water is used, the Atterberg limits, especially the liquid
limit, can be affected due to the possibility of ion exchange between the soil and
impurities in the water. In general, the liquid limit obtained from the soil treated
with tap water is lower than the value obtained using deionized water. The liquid
limit increases if excessive amount of NaCl is present in the water. Therefore, it
is recommended to use deionized or distilled water to determine the Atterberg
limits. After allowing the soil to hydrate for 16 hours, the test proceeded from the
wet side to the dry side while drying the soil with a fan.

91
The Atterberg limits of LR soil are summarized in Table 3.2. Similar to the in-situ
natural water contents and the organic contents, the Atterberg limits of LR soil
vary between samples. The block sample has the highest liquid and plastic limits,
followed by the disturbed samples collected in April and in July. The Atterberg
limits of LR soil are plotted in Figure 3.6 as a function of organic content. As
shown in the figure, the Atterberg limits of LR soil generally increase with organic
content. Hobbs (1986) noted that Atterberg limits of peats depend mainly on the
degree of decomposition of organic matter, the parent material, and ash content.
He also observed that Atterberg limits of peats increase with organic content and
degree of humification or decomposition of organic matter.
The Plasticity Index (PI) of LR soil also increases with organic content, as shown
in Figure 3.7 (a). The relationship between organic content and plasticity for an
organic soil (Juturnaiba organic clay from Brazil) is shown in Figure 3.7 (b). As
shown in the figure, the plasticity index and the organic content of the soil are
linearly related; the relationship can be expressed as:
PI(%) = 2.21 x OC (%) + 21.0

where OC>10%

(3.3)

Equation (3.3) is plotted as a solid line in Figure 3.7 (a). The figure shows that
the change in plasticity index with organic content, thus the influence of organic
matter on the plasticity index, of LR soil is similar to that of other organic clays.
The Atterberg limits of LR soil are plotted on a plasticity chart in Figure 3.8. As
shown in the figure, LR soil plots slightly lower than the A-line, and falls in the
region for highly plastic inorganic silts (MH) or organic clays (OH). The Atterberg
limits of LR soil are compared with other peats in Figure 3.9. Amorphous peats
and fibrous peats also plot slightly lower than the A-line, but with liquid limit
higher than about 300 %. The figure shows that the Atterberg limits of LR soils
are slightly lower than those of various types of peats, but have a similar trend.
For the disturbed LR soil samples collected in April of 2001, another set of liquid
limits was determined after drying at 105 5 C for 24 hours. When soils are

92
completely dried, many changes can occur: the fabric of the soil can change, soil
particles can be subdivided or agglomerated and the absorbed water can
evaporate, a process that can not be reversed by rewetting. These changes are
significant for organic clays and silts. It was observed that LR soils were
hardened after oven drying. Therefore, the soil was finely grinded before the
determination of the liquid limit. As summarized in Table 3.2, the liquid limit of LR
soil significantly decreased after oven drying to 24 36% of the liquid limit of the
non-oven dried soil. According to the USCS classification system, if the ratio
between the liquid limits of the oven dried and non-oven dried soil is smaller than
75%, the soil is classified as organic.
Based on the results from particle size distribution analysis in conjunction with
the Atterberg limits, LR soil can be classified as organic silt with sand (OH),
according to the USCS (ASTM D2488).
3.2.6. Fiber content
The fiber content of LR soil was approximately 2.29% (Humphrey, 2001). This
would categorize LR soil as sapric, an organic soil with fiber content less than
33%, according to the ASTM standard D1997. Sapric organic soils are classified
from strongly humidified (H7) to completely humidified (H10) according to the Von
Post System. The low fiber content of LR soil indicates that the organic matter of
LR soil is highly decomposed.
3.2.7. Soil acidity
The pH of LR soil was determined in accordance with ASTM D4972. The factors
that may influence the pH are: (1) the nature and type of inorganic and organic
constituents of the soil; (2) the soil to solution ratio; (3) the salt or electrolyte
content; and (4) the CO2 content (SSSA book series 9, 1982). The ASTM
standard (D4972) recommends measuring pH in deionized H2O and 0.01 M
CaCl2 solutions. Most naturally deposited soils contain salts from fertilizer or
irrigation water. The salts displace exchangeable Al which increases the

93
concentration of H+ ion in the solution when hydrolyzed. The result is that the pH
of soils containing salts is generally low. The effect of salt content on the pH
becomes smaller when measured in 0.01 M CaCl2 solution, since the sodium
ions of the salt in the soil are replaced with Ca2+. The pH values of LR soil
measured in deionized H2O and 0.01 M CaCl2 were 5.93 and 5.60, respectively.
The results indicate that LR soil is slightly acidic (pH between 5.5 and 7.0).
Figure 3.10 presents the pH of organic soils and peats as a function of the
organic content. MacFarlane (1969) noted that the pH of organic soils gradually
decreases with organic content. However, as shown in the figure, the pH of
peats (organic content higher than 75%) varies significantly. The pH of LR soil
measured in H2O and 0.01M CaCl2 solutions plots close to the linear best-fit
curve, indicating that the pH value of LR soil is similar to the value of other
organic soils with similar organic content.
3.2.8. Soil taxonomy classification
Soil taxonomy is a soil classification system that is most widely used in soil
science and soil survey. Soil taxonomy classifies a soil into 12 orders based on
the presence or absence of major diagnostic characteristics such as horizon
morphology, wetness, climate, mode of deposition, and texture (USDA, 2003).
The 12 soil orders of soil taxonomy and their characteristics are described in
Chapter 2.
The classification of soils in soil taxonomy is based primarily on the field
description of a soil to a depth of about 2 m. The lower boundary of soil is
arbitrarily set to 2 m for classification purposes, as biological activity and
pedogenic process (soil formation process) generally extends only to 2 m.
Therefore, soil taxonomy has a strong agricultural bias (USDA, 2003). Since the
soils employed for this research were sampled at a depth ranging from about 1 2 m (July sample) to about 5 - 6m (April sample), as described in Section 6.1, soil
taxonomy classification is not applicable to all the soil samples. However, soil

94
taxonomy classification can be used as a reference to compare the properties
and possibly the mechanical behavior of the soil to other similar soils
encountered in different regions.
The soil is mapped as Houghton muck in the Tippecanoe County Soil
Survey (USDA, 1998). Houghton is the name of the soil series, and muck is
the texture of the surface horizon (layer). In soil taxonomy, soils that contain
more than 20% of organic matter are called peat or muck. The Houghton series
is predominantly found in the Midwest region of the U.S., including Michigan,
Iowa, Illinois, Minnesota, Wisconsin, and Indiana (NRCS, 2004). The organic
fibers in the Houghton series are originated from herbaceous plants and some
layers contain as much as 30 % of woody material. The Houghton series is
generally highly decomposed and is classifies as sapric. In soil taxonomy
organic soils can be classified into three types: fibric, hemic and sapric, based on
the degree of decomposition of the plant matter from which the organic materials
are derived. Sapric soils are the most decomposed among the three types and
have the smallest amount of plant fiber, the highest bulk density, and the lowest
water content. Sapric soils are generally very stable which means that the
chemical and physical properties of the soils change very little with time. ASTM
standard D1997 classifies an organic soil with fiber content less than 33% as
sapric. As described in section 3.2.6, the fiber content of the soil was
approximately 2.29%.
The soil is classified in soil taxonomy as Euic, Mesic Typic Haplosaprists. Euic
indicates that the pH of the soil is generally above 4.5 in 0.01M CaCl2 and is rich
in nutrients. Mesic indicates that the mean annual temperature of the soil is
generally between 8 and 15. Typic Haplosaprists indicates that the soil is a
typical Histosol, a soil group that contains more than 20% organic matter, and
that the organic soils are sapric. Haplo means that the soil does not have
additional features that define other sapric soils.

95
3.3. Mineral composition of LR soil
Understanding the mineralogy of certain deposits can provide useful information
to predict or understand the behavior of the soil. For example, a clay deposit
abundant in Texas, which is classified as Vertisol, has very high shrink-swell
potential and has caused serious foundation problems (Reid-Soukup and Ulery,
2002). The problems are caused mainly by smectite, which is the most abundant
clay mineral in Vertisol.
The mineral composition of an organic soil is may be used to better understand
the interaction between the organic and the inorganic matter of the soil. In
general organic substances can be retained by clay minerals in two ways: (1) by
attachment to clay mineral surfaces through cation and anion exchange, H
bonding or van der Waals forces, and (2) by penetration into the interlayer
spaces of expanding-type clay minerals, such as smectite and vermiculite
(Stevenson, 1994).
The mineral composition of LR soil was obtained from X-ray diffraction (XRD)
analysis from the inorganic components of the soil, since soil organic matter is
amorphous and thus cannot be investigated by XRD analysis. A detailed XRD
analysis requires pre-treatment of the specimen for removal of cementing agents
and separation of the soil into different size fractions since many minerals are
found in distinct size ranges. For example, kaolinite, the most abundant 1:1 type
phyllosilicate mineral, forms larger minerals through very strong hydrogen
bonding and is most commonly found in the coarse clay (2 0.2m) to fine silt
size fractions. Smectite and vermiculite are usually found in the fine clay size
fraction (<0.2m). In this experimental program, LR soil was separated into six
size fractions (>50m, 50 20m, 20 5m, 5 2m, 2 0.2m and <0.2m)
following the procedure recommended by Schulze et al (2000).

96
3.3.1. Fractionation of LR soil
Fractionation of a soil sample consists of two major steps: (1) dispersion of soil
particles, and (2) separation of soil particles into different size fractions.
Dispersion of soil particles is required so that particles can be easily separated.
During dispersion, the following materials were removed: cementing agents,
divalent and trivalent cations and salts.
Cementing agents such as calcium carbonate, organic matter, iron oxide, and
silica can cement individual soil particles together and form aggregates of sand
or silt size. The calcium carbonate was removed with pH 5 NaOAc. The organic
matter was removed with 30% hydrogen peroxide (H2O2).
Divalent and trivalent cations such as Ca2+, Mg2+ and Al3+ decrease the double
layer thickness of phyllosilicate minerals, which promotes flocculation. Such
cations in LR soil were replaced with Na+ during the treatment of the soil with pH
5 NaOAc.
A high salt content also promotes flocculation. During the dispersion step,
calcium carbonate and organic matter were removed using a solution of high salt
content. Therefore, after removal of cementing agents, salts were washed with
deionized water from the soil sample.
Following dispersion, the soil particles were separated into the six size fractions
mentioned by wet-sieving, gravity sedimentation and centrifugation. Soil particles
larger than >50m were separated by wet-sieving. The soil particles of clay size
fraction were separated into coarse clay (2 0.2 m) and fine clay (< 0.2 m)
fractions by centrifugation. The soil particles between 50 m and 2 m were
separated into three fractions, 50 20 m, 20 5 m, and 5 2m, by gravity
sedimentation. The details of each step are described below.

97
3.3.1.1. Procedure for dispersion of soil particles
The specimens used for XRD analysis were prepared with LR soil from the bin
LR-B9, which was collected in July 2001 (see Table 3.2 for average index
properties of the soil in LR-B9 bin). LR soil equivalent to 15g dry mass was
placed in a 250ml polypropylene centrifuge bottle. About 200ml of pH 5 NaOAc
solution was added to remove carbonates from the soil. The treatment with
NaOAc solution also makes the negatively charged sites (atoms) at the surface
to be saturated with Na+, so the individual soil particles get dispersed. After
treatment, the sample was centrifuged at 1500 rpm for 5 minutes and the clear
supernatant was discarded.
The organic matter was removed by treating the soil with 30% hydrogen peroxide
(H2O2). About 10ml of 30% H2O2 was added to the soil. The bottle with the soil
was placed in a water bath at a controlled temperature of about 100 C. The
reaction is auto catalytic, which means that the released heat accelerates the
reaction. When the reaction became very violent, the samples were stirred
vigorously or the bottle was placed in a cold water bath to slow down the reaction.
For LR soil, about 140 ml of 30% H2O2 was required to remove the organic
matter. After treatment, the sample was suspended in a 200ml pH5 NaOAc
solution and centrifuged at 1500 rpm for 5 minutes to separate the solution. The
clear supernatant was discarded.
The excess salts in the soil were removed by washing the soil with 150 ml of
deionized H2O. The sample was suspended into deionized H2O by shaking and
then centrifuged at 1500 rpm for 5 minutes to separate the solution. If the
supernatant was clear after centrifuging, which indicated that the soil particles
were not fully dispersed, the washing and centrifuging was repeated. The
procedure was repeated until the supernatant became cloudy, which indicated
dispersion of soil particles.

98
After the treatment, about 150ml of deionized H2O was added and the soil
particles were mechanically dispersed by sonicating the samples for 10 minutes.
Following dispersion, the pH of the suspension was adjusted to 8.0 using 1M and
0.2M NaOH solution. The edges of the clay minerals and the iron and aluminum
oxides minerals become positively charged at pH about neutral or below. In this
case, the positively charged minerals are attracted to the negatively charged clay
mineral surface causing flocculation. At pH higher than 10, the iron and
aluminum oxides are negatively charged, promoting dispersion. However, silica
becomes soluble at pH higher than 8.5. Therefore, the pH of the solution was
adjusted to 8.0 preventing flocculation and dissolution of silica.

3.3.1.2. Procedure for fractionation


The LR soil particles were fractionated following the procedures recommended
by Jackson (1969). The fractionation procedure is shown schematically in Figure
3.13.
The particles greater than 50 m were separated by wet sieving. The soil was
placed on a 50 m sieve and was washed with deionized H2O which pH was
adjusted to 8 with a Na2CO3 solution. The solution containing soil particles
smaller than 50 m was retained into a beaker. After washing the particles
thoroughly, the sieve and the particles greater than 50 m were dried in an oven
at 105 C.
The solution containing particles smaller than 50 m was centrifuged to separate
soil particles into two size fractions: 50 2 m and < 2 m. The time required for
centrifugation for the separation at 2 m was calculated using Stokes law as
follows:

99

tmin =

63.0 108 n log10

(N )

(s

R
S

sl ) D 2

(3.4)

where n = viscosity of the liquid (poises)


R = radius of the top of the sediment (cm)
S = radius of rotation of the surface of the suspension (cm)
N = centrifuging velocity (rpm)
sp = specific gravity of particles
sl = specific gravity of water
D = equivalent spherical diameter of soil particles (m)
R and S are defined as shown in Figure 3.11. The time (t) calculated with
equation (3.4) represents the time required for soil particles of an equivalent
spherical diameter D to travel from S to R at velocity N.
Two lines were marked on the bottle at 14.5 cm (S) and 23.1 cm (R) from the
axis of rotation. The temperature of the centrifuge chamber was kept at 21 C.
The viscosity and the specific gravity of water at 21 C were used in equation
(3.3). The bottle was filled with the solution containing soil particles smaller than
50 m up to the top line drawn at S = 14.5 cm. The bottle was centrifuged at 500
rpm for 7.5 minutes. During centrifuging, all the particles in the 50 2 m size
sedimented below the bottom line drawn at R = 23.1cm. After separation, the
solution containing particles smaller than 2 m was poured into a new beaker.
The particles smaller than 2 m were further separated by centrifugation into two
size fractions: particles in the range of 2 0.2 m and particles smaller than 0.2

m. Due to the smaller particle size, the separation at 0.2 m required higher
centrifuging speed. Using equation (3.3), the time required to separate particles
smaller than 0.2 mm at 3000 rpm was calculated as 35.9 minutes. An IEC PR
7000 centrifuge machine equipped with an IEC model 949 centrifuge head was
employed for this procedure. After centrifugation, the particles of 2 0.2 m were

100
sedimented on the bottle, and the solution containing particles smaller than 0.2
mm was placed into a new beaker.
The pH of the solution containing particles smaller than 0.2 m was adjusted to
6.0 with dilute HCl so that the particles could flocculate. At this pH, the edges of
phyllosilicate minerals are positively charged and thus soil particles flocculate.
After all particles smaller than 0.2 m flocculated, the clear supernatant was
discarded. The particles smaller than 0.2 m were suspended into a small
amount of deionized water. The solution was poured into a dialysis tube and was
heavily dialyzed against deionized water to remove excess salts.
The particles 50 2 m were separated into three size fractions: 50 20 m, 20
5 m and 5 2m, by gravity sedimentation. The time (t) required for a particle
of equivalent spherical diameter D to travel a distance h was calculated using
Stokes law as follows:

t=

18nh
g ( s p sl ) D 2

(3.5)

where n = viscosity of the liquid (poises)


h = travel distance (cm)
g = gravitational constant
sp = specific gravity of particles
sl = specific gravity of water
D = equivalent spherical diameter of soil particles (cm)
The particles with 50 2 m size were suspended into 50ml of pH=8 H2O. Two
lines were drawn on the outside of a 600ml beaker at 6cm and 1cm from the
bottom. In this case, the travel distance h was set to 5 cm. Using equation (3.5)
the sedimentation time required for separation at the particle size of 5m was
calculated as 33 minutes and 20 seconds. The solution containing the particles of
50 2 m was poured into the beaker up to the top line (6cm from the bottom)

101
and was allowed to stand undisturbed. Exactly after 33 minutes and 20 seconds,
the particles of 50 5 m had sedimented below the bottom line (1cm from the
bottom) and the particles of 5 2 m still remained in the solution. The solution
was poured into a 250ml bottle and was centrifuged for 5 minutes at 1500 rpm.
The 50 5 m sediments were suspended into deionized water which filled the
beaker up to the top line. The suspension was allowed to stand for exactly 2
minutes and 20 seconds. At the end of this time, the 50 20 m size particles
were sedimented below the bottom line. The solution containing 20 5 m size
particles was poured into a new beaker. The particles in these two fractions were
oven dried at 105 C.
The three solutions containing particles 5 2 m, 2 0.2 m, and the particles
smaller than 0.2 m were frozen with liquid nitrogen, which was freeze-dried to
remove the moisture.
The particle size distribution of the inorganic components of LR soil is
summarized in Table 3.3. The total mass of the inorganic components was
58.9% of the total soil. The result showed good agreement with the average loss
on ignition of LR soil from bin LR-B9 (40.0 4.0). The particle size distribution of
the inorganic components of LR soil is compared with the particle size
distribution of the total LR soil in Figure 3.12. The particle size distribution curve
of the inorganic component of LR soil lies above the curve of the whole soil
(inorganic and organic). The difference between these two curves indicates the
distribution of organic matter in the soil. What is interesting to note is that the
inorganic component of LR soil contains about 62% of clay-size particles
(<0.002mm) while the entire soil contains only 41.5% of clay-size particles. This
indicates that the organic matter is adsorbed by fine-grained particles. For the LR
soil the organic matter is adsorbed by silt- and clay-size particles that are smaller
than 0.02 mm and forms bigger soil particles.

102
3.3.2. Mineral composition of LR soil
X-Ray Diffraction (XRD) analysis was performed on the three inorganic soil
fractions: 5 2 m, 2 0.2 m and <0.2 m. X-ray Diffraction patterns were
obtained employing a Phillips 3520 vertical diffractometer equipped with a fixed
1o divergence slit and a 0.2 mm receiving slit. Cobalt target was used to generate
X-ray of a wavelength of 1.79026 .
For each fraction, two specimens were prepared. One specimen was saturated
with Mg and glycerated, and the second specimen was saturated with K. These
steps are essential to identify smectite and vermiculite. Smectite is a 2:1
phyllosilicate mineral with a layer charge per unit formula in the range of 0.25
0.6. Because of the lower layer charge, the smectite can have four layers of
water molecules. When the interlayer exchange sites are saturated with Mg2+
ions and glycerol molecules, smectite expands fully to a d-spacing of 18 . On
the other hand, when the inter layer exchange sites are saturated with K+ at room
temperature, smectite collapses to a d-spacing of 10 12 . Vermiculite is also a
2:1 phyllosilicate mineral but has a higher layer charge than smectite, in the
range of 0.6 0.9. When saturated with Mg2+ ions and glycerol molecules,
vermiculite fully expands to a d-spacing of 14 . When saturated with K+ ions,
vermiculite also collapses to a d-spacing of 10 .
The K-saturated specimens were re-analyzed after heating at 300 C and 550 C
for two hours to collapse hydroxyl-interlayered smectite or vermiculite to a dspacing of 10 . The heat treatment is also useful for identification of kaolinite.
Kaolinite is the most common 1:1 phyllosilicate, and has a d-spacing of 7.2 .
When heated at 550 C, kaolinite is de-hydroxylated and the 7.2 peak
disappears.
The Mg-glycerated and the K-saturated specimens were kept at room
temperature at 21 C and were step-scanned for 2range from 2 to 40o with a

103
0.5o increment and two-second count time. After heat treatment, the K-saturated
specimens were scanned for 2-range of 2 to 28. The 2 of 28 corresponds to
3.7 .
The minerals identified by XRD analysis in the three size fractions of LR soil are
summarized in Table 3.4. The results of XRD analysis were used only to identify
the major minerals of the inorganic components of LR soil smaller than 5 m.
Measurements of the quantities of minerals require calibration of XRD patterns
with a specific mineral of known quantity. As discussed previously, comparison of
the particle size distribution of inorganic components and the combination of
inorganic and organic components of LR soil indicates that organic matter is
adsorbed by inorganic particles smaller than 0.02 mm (20 m). Thus, the results
of XRD analysis can be used to identify the major minerals that have adsorbed
the organic matter.
The XRD patterns of the inorganic components of LR soil for the 5 2m size
fraction are shown in Figure 3.14. The soil in this size fraction is mineralogically
the most complex. Quartz is indicated by the sharp first order peak at 3.36 and
the second order peak at 4.27 . The mica peaks are shown at 10.06, 5.00 and
3.33 . The third order peak of the mica at 3.33 is hindered by the first order
peak of quartz at 3.36 . The mica is probably dioctahedral (muscovite). For
trioctahedral mica (biotite), the second order peak at 5.00 is usually not
observed. The 14.34 peak, which disappears after K-saturation, indicates the
presence of vermiculite. Kaolinite is indicated by the 7.12 peak which
disappears upon heating at 550 C, and the second order peak at 3.55 .
Gibbsite is indicated by the peak at 4.74 , which disappears after heating at 300
C. A small peak at 3.26 indicates the presence of rutile (TiO2). The peak at
3.20 may indicate the presence of feldspar: anorthite and oligoclase have the
first peak at 3.20 , andesine and anorthoclase at 3.21 .

104
Figure 3.15 presents the XRD patterns of the inorganic components of LR soil in
the 2 0.2 m size fraction. The X-ray diffraction pattern of the K-saturated
specimen after heating at 300 C has high noise due to unexpected technical
difficulties. The sharp peak at 3.35 and the 4.26 peak indicate quartz. The
peak at 7.20 , which disappears after heating at 550 C, indicates kaolinite. A
small peak at 3.58 is the second order peak of kaolinite. The peak at 14.29 ,
which shifts to 10.11 upon K-saturation, indicates vermiculite. The vermiculite
in this size fraction may be dioctahedral vermiculite weathered from dioctahedral
mica. Vermiculite is a 2:1 type clay mineral. As the XRD analysis results indicate,
the double layer thickness of vermiculite can change from about 10 to about 14
. The double layer thickness of vermiculite changes significantly depending on
the type and amount of ions present in the pore water and the ionic strength of
the pore water. In general, vermiculite is considered as a highly reactive clay
mineral. The change in the particle size with organic matter in 2 0.2 m size, as
shown in Figure 3.12, is mainly due to the aggregation of vermiculite with organic
matter. The peaks at 10.11 and 5.01 in Mg-glycerated specimen indicate
dioctahedral mica.
Figure 3. presents the XRD patterns of the inorganic components of LR soil
smaller than 0.2 m. The peak at 19.09 , which shifts to 10.08 upon
saturation with K at room temperature, indicates smectite. Smectite is also a 2:1
type clay mineral, in which the double layer thickness can change from about 10
to about 18 19 . Smectite is also a very highly reactive clay mineral. The
change in the particle size with organic matter in < 0.2 m fraction is mainly due
to the reaction of smectite with organic matter. The second and third order peaks
of smectite are shown at 5.03 and 3.34 . A very weak peak at 10.08 in the
Mg-glycerated specimen indicates a small amount of mica, probably
dioctahedral. Kaolinite is indicated by peaks at 7.18 which disappear at 550 C,
and by the second order peak at 3.59 .

105
As described in section 2.3, the reaction between soil organic matter and soil
mineral occurs through attachment of organic matter on fine particle surfaces and
penetration of organic matter into the interlayer spaces of expandable clay
minerals. The results of XRD analysis indicate the presence of vermiculite in the
2 0.2 m size fraction, and smectite in the smaller than 0.2 m size fraction.
These two clay minerals are very reactive. The significant increase in particle
size of the clay-size fraction, as shown in Figure 3.12, is mainly due to the
interaction between vermiculite and smectite and organic matter.

3.4. Characterization of the organic matter in LR soil


Similar to the inorganic components, organic matter plays a significant role in
affecting the chemistry of a soil. As shown in section 3.2, organic soils have very
high water content. This is mainly due to the high water holding capacity of
organic matter. Soil organic matter combines with clay minerals to form
aggregates; the increase in particle size causes increase in hydraulic conductivity
of the soil. Soil organic matter has a very high specific surface of 800 900 m2/g
and a very high cation exchange capacity in the range of 300 1400 cmoles/kg.
The cation exchange capacity of organic soil is mainly attributed to soil organic
matter (Stevenson, 1994).
The humic substances of soil organic matter have three main fractions: humin,
humic acid, and fulvic acid. Humin is the fraction that is insoluble in alkali. Humic
acid is the dark colored organic material that is soluble in alkali but is insoluble in
acid. Fulvic acid is the fraction of soil organic matter that is soluble in both alkali
and acid. Among the three fractions, fulvic acid has the highest molecular weight,
cation exchange capacity and reactivity.

106
In this section, the soil organic matter present in LR soil is characterized and the
chemical composition and the organic functional groups that exist in each fraction
of LR soil are determined.
3.4.1. Extraction and fractionation of the humic substances in LR soil
Fractionation of humic substances is necessary for analysis to reduce
heterogeneity of the isolated material (Stevenson, 1994). The extraction and
fractionation of organic matter were carried out using solubility characteristics:
humic acid is insoluble in acid whereas fulvic acid is soluble in acid. The humic
and fulvic acids were dissolved in a highly basic solution (pH >12) and were
separated from the humin and inorganic components of the soil. The humic acid
was separated from the fulvic acid by precipitating humic acid at low pH (pH <1).
The procedure for fractionation of the organic matter of LR soil is shown
schematically in Figure 3.17.
LR soil equivalent to 15g dry mass was added to a 250ml polypropylene
centrifuge bottle. LR soils from two bins (LR-B9 and LR-B11) collected in July,
2001 were used. About 150ml of 0.1M HCL-HF solutions were added to soil
samples with a solution to solid ratio of 10:1. The treatment with HF dissolves
silicates in the soil and reduces the ash content of the extracted organic matter.
The HF solution also dissolves inorganic carbon and heavy metal ions in the soil.
The treatment lasted for three days with the bottle being shaken in a mechanical
shaker. After treatment, the bottle was centrifuged at 2000 rpm for 15 minutes to
remove the solution.
A 0.5N NaOH solution was used to extract the humic and the fulvic acids from LR
soil. About 150ml of the solution was added to soil samples with a solution to
solid ratio of 10:1. The pH of the solution increased above 12.74 after adding the
NaOH solution. The extraction was carried out under a blanket of N2 gas to
minimize chemical changes due to autoxidation. The extraction step lasted two
days. After two days, the bottle was closed tightly with a cap and was shaken in a

107
mechanical shaker for two hours to suspend all the particles into the solution.
The bottle was centrifuged at 3000 rpm for 30 minutes to separate the dissolved
humic and fulvic acids from humin and the inorganic components of the soil. The
solution was poured into a 500ml polypropylene centrifuge bottle. 150ml of 0.5N
NaOH solution was added to the sediments again and the bottle was shaken for
2 more hours. After centrifuging, the solution was added to the 500ml bottle
containing what has been previously extracted, and the humin and the inorganic
components of the soil were air-dried.
The pH of the solution was adjusted to 1.0 with 5M HCl to precipitate humic acid.
After 24 hours, the bottle was centrifuged at 4700 rpm for 45 minutes to separate
the fulvic acid from the precipitated humic acid. The temperature of the centrifuge
chamber was kept at 2 C. The solution containing fulvic acid was poured into a
500ml bottle. The precipitated humic acid was dissolved into 50ml of a 0.5N
NaOH solution. The bottle was shaken in a mechanical shaker for 24 hours. The
pH adjustment and the centrifugation were repeated one more time to separate
humic and fulvic acids.
The humic acid fraction almost always contains a considerable amount of
inorganic materials, since humic acids are very strongly bonded to mineral
matter. Therefore, a purification procedure was applied to the extracted humic
acid fraction to reduce the ash (inorganic material) content. The most widely
used method for purification is to treat humic acid with acid solution containing
HF. The procedure recommended by Stevenson (1994) was applied for
purification of the humic acid.
50ml of 0.5N NaOH solution was added to redissolve the precipitated humic acid.
The ionic strength of the solution was adjusted to 0.1 with respect to KCl, and the
bottle was centrifuged at 4700 rpm for 45 minutes to separate inorganic material.
The inorganic material was added to the humin and the inorganic components of

108
LR soil and was air dried. The humic acid was treated with 50ml of 0.3N HF:
0.1N HCl solution to dissolve any inorganic material that might still be present in
the humic fraction. The bottle was shaken for two hours in a mechanical shaker.
The bottle was centrifuged at 4700 rpm for 45 minutes and the clear supernatant
was discarded. The treatment with the 0.3N HF: 0.1N HCl solution and
centrifugation were repeated three more times.
The purified humic acid was redissolved into 50ml of 0.5N NaOH solution. The
humic and fulvic acid fractions were poured into a dialysis tube and were heavily
dialyzed with deionized water for one week to remove excess salts. After dialysis
was completed, the humic and fulvic acids were put into a polypropylene
centrifuge bottle and were frozen with liquid nitrogen. The opening of the bottle
was covered with cloth and tightly sealed with a rubber band. The bottle was
placed in a freeze-drier to completely remove water from the humic and fulvic
acids.
The masses of the three fractions of the LR soil are summarized in Table 3.5.
The results show very good repeatability. The humic and fulvic acids count for
approximately 23.5% and 2.4%, respectively, of the total mass of LR soil. The
ash content was determined by combusting the humin and inorganic components
and the humic acid fraction in a furnace at 440 C. The ash content of the humic
acid fraction is very low, in the range of 4.26 6.94%. This indicates that the
treatment with HCl-HF solution was very effective in purifying the humic acid.
The results also show that the ratio of the humic acid to fulvic acid of LR soil is in
the range of 9.7 0.7. Kononova (1996) reported that the humic acid to fulvic
acid ratio for various surface soils was in the range of 0.3 2.5. However, direct
comparison cannot be made since the ratio changes depending on the
procedures and types of solvents applied for extraction, fractionation, and
purification. For instance the ratio of the extracted humic to fulvic acids from a H+

109
exchanged Sapric Histosol changed from 0.8 (with sulfolane) to 29.0 (with 0.5M
NaOH) (Hayes, 1985). The 0.5M NaOH solution yielded the highest amount of
humic acid but relatively smaller amount of fulvic acid among the nine different
solvents employed. In general the humic substances of forest soils (Alfisols,
Spodosols, and Ultisols) have lower ratios of humic acid to fulvic acid, which are
in the range of 0.3 1.0, than the humic substances of peat and grassland soils
(Mollisols), which have a range of 1.0 2.0 (Stevenson, 1994).

3.4.2. Elemental composition of the organic matter in LR soil


The elemental composition of each fraction of LR soil was evaluated employing
the LECO CHN-2000 analyzer available in the USDA soil erosion lab. The
organic carbon, hydrogen, and nitrogen contents were measured. The LECO test
is in principle similar to thermal analysis. In the LECO test, approximately 0.2g of
soil is placed in an aluminum container. The specimen is dropped into the
combustion chamber where the temperature of the furnace and the flow of
oxygen gas cause the specimen to combust. During the combustion process,
carbon, hydrogen and nitrogen are converted into CO2, H2O, N2 and NOx. The
gases are passed through infra red (IR) cells to determine the carbon and
hydrogen content and to a thermal conductivity (TC) cell to determine the
nitrogen content. (CHN-2000 Elemental Analyzer Instruction manual).
The results of the elemental analysis are summarized in Table 3.6 . The humic
acid of LR soil contains about 51.4% carbon, 4.65% hydrogen and 3.85%
nitrogen. The fulvic acid contains about 44.4% carbon, 2.85% hydrogen, and
1.42% nitrogen. The humin and inorganic components of LR soil contain
significantly less carbon than the other two fractions, because of the inorganic
nature of this fraction. The C/H and C/N ratios for the humic acid are 11.1 and
13.4, respectively. The C/H and C/N ratios for the fulvic acid are 15.7 and 32.7,
respectively.

110
According to Tan (2003), the C/N ratio for humic acid indicates the degree of
decomposition of organic matter. In general, the C/N ratio for humic acid falls
between 10 and 15 when decomposition is virtually completed, and the C/N ratio
for fulvic acid ranges from 18.5 to 38.0. The higher C/N ratio for the fulvic acid is
due to the fact that the N content is about 2 3 times higher in the humic acid
than in the fulvic acid. The C/N ratios for the humic and fulvic acids of LR soil fall
within the typical ranges of highly decomposed soil, which indicates that the
degree of decomposition or humification of the organic matter in LR soil is high.
This is in good agreement with the fact that LR soil has a very low fiber content
(about 2.29%, see section 3.2.6), which is another indication of the high degree
of decomposition of the organic matter in LR soil.
As described in section 2.2, the elemental composition of the three humic
substances of organic soils does not vary much. The typical elemental
compositions of the three humic substances of various organic soils and four
Florida Histosols are shown in Tables 3.7 and 3.8, respectively. Carbon is one of
the major elements in humic and fulvic acids. In general, humin has the highest
carbon content, usually more than 62%, followed by humic acid (51 62%) and
fulvic acid (42 47%). The average contents of hydrogen and nitrogen are 3
7% and 1 4 %, respectively (Sparks, 2003). Approximately 90% of the nitrogen
in most soils occurs in the following types: (1) free amino group (-NH2-), (2) open
chain group (-NH-, =N-), (3) part of a heterocyclic ring (Stevenson, 1994).
The chemical composition of the humic and fulvic acids of LR soil are generally in
the average range of these other organic soils, except for the humic and fulvics,
which have a carbon content on the lower bound of those found in typical organic
soils.
Compared to other organic soils, such as the four Florida Histosols shown in
Table 3.8, the humic acid of LR soil contains about 5 9 % less carbon. The

111
fulvic acid of LR soil also contains less carbon (about 2 16 % less) and less
hydrogen (about 1 5 % less) than the fulvic acids of the four Florida Histosols
and the other organic soils (shown in Table 3.7).
As described in section 2.2, the most abundant carbon-containing organic
functional group in organic matter is the carboxyl acid (COOH). Carboxyl acid is
also the most highly reactive organic functional group. The low carbon contents
of the humic and fulvic acids of LR soil suggest that the LR soil contains less
carboxyl acid in humic and fulvic acids than other organic soils.

3.4.3. FT-IR analysis of the humic substances of LR soil


The humic substances of LR soil were characterized by employing Fourier
Transform Infrared Spectroscopy (FT-IR) analysis. XRD analysis can only be
used to identify materials of distinctive atomic structure. Since humic substances
are amorphous, XRD analysis is not adequate to study them. The principle of the
FT-IR analysis is based on the fact that certain molecules vibrate at discrete
energies in the IR region of the electromagnetic spectrum. There are two general
types of vibration modes: (1) stretching and (2) bending. The stretching vibration
can occur symmetrically or asymmetrically. The vibration modes of water
molecules are schematically shown in Figure 3.18. The O-H stretching vibrations
(v1 and v3) occur in the 3200 3700 cm-1 region, while the O-H bending vibration
(v2) occurs in the 1600 1650 cm region. When the frequency of the incident IR
radiation matches the frequency of the particular vibrational mode of certain
molecule, IR absorption occurs. The vibration of most organic molecules occurs
in the mid-IR spectral region (400 4000 cm-1).
One of the shortcomings of the FT-IR analysis is that it can not identify specific
organic functional groups, as most organic functional groups are similar in their
chemical composition. As discussed in section 2.2, most of the organic functional
groups consist mainly of C, H, O and N, and thus vibration of several organic
functional groups occurs in the same frequency regions. The characteristic

112
infrared adsorption frequencies of various organic functional groups are
summarized in Table 3.9.
The main objectives of FT-IR analysis are to obtain the FT-IR spectrum of the
organic matter that is present in LR soil and to compare the FT-IR spectrum with
other data obtained from the literature. Other analytical techniques, such as 13C
NMR method, may be required to identify specific organic functional groups such
as salicylic acid and acetic acid that are known to interfere with cement reactions
(see section 2.6). It should be emphasized however that the interpretation and
comparison of the FT-IR analysis results should be performed very carefully
because of the complex nature of the humic substances of soil organic matter,
soil variability, differences in extraction and fractionation methods (including the
solution employed), and testing error (Tan, 2003, Stevenson, 1994, Schnitzer
and Khan, 1972).
The FT-IR spectrum of each of the three fractions was obtained employing a
Perkin-Elmer 1600 series Fourier transform infrared spectrophotometer. The
specimen was scanned 64 times from 370 4000 cm-1 with 0.5 cm-1 interval and
2 cm-1 resolution. The specimens were prepared in solid form using potassium
bromide (KBr).
KBr pellets were grinded into fine powder, and the powder was oven dried at 105
C for 4 5 hours to completely remove H2O. The freeze-dried LR soil humic
substances were grinded into fine powder. About 2mg of the humic substances
were thoroughly mixed with 250 mg of KBr powder. The mixture was placed
inside the KBr pellet apparatus. The air inside the apparatus was removed by
applying vacuum for five minutes. Approximately 100kPa of vertical stress was
applied to the apparatus with a hydraulic jack for 10 15 minutes while applying
vacuum. During this process, the KBr power solidifies and a KBr pellet is

113
prepared. The apparatus was disassembled, and the KBr pellet was stored in a
glass jar until testing.
The organic functional groups identified with FT-IR analysis in the three fractions
of the organic matter of LR soil are summarized in Table 3.10.
The FT-IR spectrum of the humin and inorganic component of LR soil is shown in
Figure 3.19. The weak broad band in the 3300 3400 cm-1 region indicates O-H
stretching of phenolic OH or NH. The two peaks in the 2902 and 2850 cm-1
regions indicate asymmetric CH stretch of CH2- (2930 cm-1) and symmetric CH
stretch of CH2- (2840 cm-1). The peak in the 1590 cm-1 region indicates
presence of carboxyl acid (asymmetric COO- stretching in 1450 1360 cm-1), NH deformation or +C=N stretching. The peak in the 1450 cm-1 region may be
related to symmetric COO-stretching (1465 1440 cm-1) or the presence of
phenolic OH and aliphatic C-H (1460 1450 cm-1). The very strong peak in the
1085 1032 cm-1 region is related to C-O stretching of polysaccharides (1170
950 cm-1). However the high ash content in the humic and inorganic fractions of
LR soil indicates that these two strong bands are related to the Si-O vibration of
the quartz and aluminosilicate layers (clay minerals).
Among the organic acids in the humin and inorganic component of LR soil
identified from the FT-IR spectrum, the organic compounds that are known to act
as retarders, water reducers, or inhibitors to cement reactions (see section 2.6)
include phenolic OH, carboxyl acid, and possibly polysaccharide.
The FT-IR spectrum of the humic acid of LR soil is shown in Figure 3.20. The
same peaks in the 3350, 2917 and 2850 cm-1 regions observed in humin and
inorganic fraction (Figure 3.19) were also observed in the humic acid, but the
intensities of these peaks are two to three times higher than in the humin and
inorganic fraction. This is due to the low ash content, or higher organic content of
the humic acid. The strong and broad band at 3350 cm-1 is due to the stretching
vibration of H bonded OH in aliphatic polyols, H bonded phenolic OH as well as

114
aliphatic OH. The two peaks at 2917 and 2850 cm-1 indicate aliphatic CH groups.
The two peaks at 1720 and 1630 cm-1 indicate presence of carboxyl acids,
amides and ketones. The peak at 1380 cm-1 indicates the presence of phenolic
OH, CH2 and CH3 (1400 1390 cm-1). The peak at 1235 cm-1 is due to
symmetric stretching of COO- (1450 1360 cm-1) and C-O stretching and O-H
bending of COOH (1250 1200 cm-1). The very strong peak in 1032 cm-1 region
is related to C-O stretching of polysaccharides (1170 950 cm-1) or O-CH3
vibration. Even though the humic and fulvic acids of LR soil were treated with
HF-HCL solution during extraction to reduce ash content, the humic acid of LR
soil still contains about 5 - 6 % of ash, as shown in Table 3.5. Thus, the strong
peak at 1032 cm-1 may be caused by the presence of clay minerals. The organic
compounds identified from the FT-IR spectrum that are known to act as retarders,
water reducers, or inhibitors (see section 2.6) include phenolic OH, carboxyl acid,
polysaccharide, and methoxyl (OCH3).
The FT-IR spectra of three different humic acids and IR spectra of humic acids
from four different sources are shown in Figure 3.21 and Figure 3.22,
respectively. The FT-IR spectrum of the humic acid of LR soil is similar to the
spectra of other humic acids, especially the humic acid of peat standard (H4),
which in Figure 3.21 shows two sharp peaks at 2920 and 2850 cm-1, a low
intensity response in the 2200 to 1800 cm-1 region, and two sharp peaks at 1725
and 1620 cm-1. This indicates that the humic acid of LR has a chemical
composition similar to that of other organic soils. The intensity of the broad band
at 3350 cm-1 for the humic acid of LR soil is smaller than that of the humic acids
from other organic soils and peats. This was also observed for the humin fraction.
This may suggests that the humic acid of LR soil has lower phenolic OH or NH
contents than other organic soils or peats. The IR spectra of humic acids from
various organic soils, shown in Figure 3.22, also show the characteristic peaks
observed in the humic acid of LR soil.

115
The FT-IR spectrum of the fulvic acid of LR soil is shown in Figure 3.23. The
spectrum is similar to that of the humic acid. However, the intensity of the band at
3390 cm-1 increased further with respect to the humic acid. This indicates a
higher concentration of phenol OH or NH in the fulvic acid. The intensities of the
peaks at 1630 and 1400 cm-1 for the fulvic acid are lower than the intensity of the
band at 3390 cm-1. This is in contrast to what was observed with the humic acid
in Figure 3.21, where the intensities at 1630 and 1400 cm-1 where higher. Since
these two intensities are associated with the carboxyl group, it follows that the
fulvic acid of LR soil has lower carboxyl content than the humic acid. This is in
good agreement with the fact humic acid has that low carbon content than fulvic
acid (Table 3.6). The adsorption band at 1400 - 1380 cm-1 indicates the presence
of phenolic OH (OH deformation and C-O stretching), CH2 and CH3 group (CH
deformation), and carboxyl acid (asymmetric COO- stretching). The low intensity
at 1400 cm-1 indicates that the concentration of phenolic OH, CH2 and CH3, and
carboxyl acid is lower in the fulvic acid than in the humic acid.
The FT-IR spectrum of the fulvic acid of LR soil also confirms the presence of
phenolic OH, carboxyl acid, and polysaccharides. As mentioned before, the
results indicate that the fulvic acid has lower carboxyl acid and phenolic OH
contents than the humin and the humic acid of LR soil.
IR spectra of fulvic acids from various organic soils are shown in Figure 3.24.
The spectrum of the fulvic acid of LR soil is similar to the spectra shown in this
figure: a very strong peak at 3390 cm-1, a weak peak at 2935 cm-1, and a strong
peak at 1630 cm-1. However, the intensity of the peak at 1400 cm-1, which is
associated with phenolic OH and CH, is higher than in the other spectra,
indicating that the fulvic acid of LR soil has a higher phenolic OH content than
other typical organic soils.

116
Table 3.1: Summary of index properties and unit weights of peats and organic
soils (after Edil and Wang, 2000)
Fiber

OC

(kN/m3)

(%)

9.110.1

83-95

64

1.40
1.70

10.3

58-65

20-30

1.60
1.90

600

9.6

81

31

1.72

240

10.2

60

20

1.94

9.8

7484

7592

1.62
1.85

12.9

361

66

1.82

Organic Soil

321

10

2.56

STH 29, Shawano, WI

Fibrous
Peat

9.612.1

3566

1.82
2.23

STH 29, Shawano, WI

Organic Soil

18.4

610

2.55

Hoyt Lakes, MN

Fibrous
Peat

250
516
125
367
270
470

10.4

5085

3676

Hoyt Lakes, MN

Organic Soil

50100

12.5

Richfield, MN

Amorph.
Peat

11.6

3137

3745

2.02

Richfield, MN

Organic Soil

175
300
150
160

25

2.29

Soil

Description

Middleton, WI

Fibrous
Peat

Middleton, WI

Amorph.
Peat

Portage, WI
Fond du Lac, WI
Nine Springs, WI
USH 12, Middleton, WI
USH 12/18, Cambridge,
WI
USH 12/18, Cambridge,
WI

Fibrous
Peat
Amorph.
Peat
Fibrous
Peat
Fibrous
Peat
Fibrous
Peat

wn (%)
500600
43052
0

150
655
157
165

content

Gs

(%)

1.59
1.70
2.47
2.63

117
Table 3.2: Summary of the index properties of LR soil
Bin

Sampling

OC

LL (%)

PL

PI

number

Time*

(%)

(%)

Not-dried Oven-dried

(%)

(%)

LR-B5

April

211.6 16.5 43.6 6.6

LR-B6

April

266.6 7.7 43.7 2.0

364.8

86.2

166.1 198.7

LR-B7

April

234.0 15.9 53.8 3.1

325.2

102.1

161.6 163.6

LR-B8

July

155.5 9.8

LR-B9

July

158.6 10.4 40.0 4.0

227.6

82.2

113.5 114.1

LR-B10

July

173.2 3.2 43.9 1.0

314.9

148.6 166.3

LR-B11

July

195.0 3.6 43.9 1.3

260.6

129.3 131.3

Block sample

April

405.7

252.7 153.0

n/a

n/a

60.1 4.9

*: all the samples were obtained in 2001

Table 3.3: Particle size distribution of the inorganic component of LR soil (LR-B9)
Size

Mass (g)

Mass %

% finer

> 50 m

0.21

1.4

1.4

50 20 m

0.91

6.1

7.5

20 5 m

2.62

17.5

24.9

5 2 m

1.94

12.9

38.9

2 0.2 m

1.42

9.5

47.3

< 0.2 m

1.73

11.5

58.9

Organic matter

6.17

41.1

Total

15.0

100.0

118
Table 3.4: Mineral composition of each of three size fractions of LR soil
Fraction size (m)

Mineral compositions

52

quartz, mica (muscovite), kaolinite, gibbsite, rutile, feldspar

2 0.2

quartz, kaolinite, vermiculite (dioctahedral), mica (muscovite)

< 0.2

smectite, mica (muscovite), kaolinite

Table 3.5: Mass of the three organic matter fractions of LR soil


Bin

Sample

number

number

Humin and inorganic

Humic acid

components

Fulvic acid

Total

(g)

(g)

HA/FA

Mass (g)

Ash (%)

Mass (g)

Ash (%)

11.46

68.89

3.94

5.08

0.37

15.77

10.6

11.41

65.85

3.38

5.88

0.37

15.16

9.1

11.01

64.81

3.33

6.94

0.34

14.88

9.8

10.77

61.19

3.59

4.26

0.39

14.65

9.2

Average

11.16

65.19

3.56

5.54

0.37

15.12

9.68

Standard Deviation

0.33

3.18

0.28

1.14

0.02

0.48

0.69

LR-B9
LR-B11

Table 3.6: Elemental composition of each fraction of LR soil


Humin and
Bin

Sample

number

number

Humic acid

inorganic components

Fulvic acid

(%)

(%)

(%)

(%)

(%)

(%)

(%)

(%)

(%)

13.75

2.49

0.98

47.91

4.83

3.48

41.83

2.63

1.16

14.28

3.66

1.04

52.73

4.41

3.81

40.54

2.46

1.01

15.16

3.80

1.11

52.16

4.72

4.03

47.74

3.16

1.74

17.69

3.44

1.28

52.83

4.64

4.06

47.64

3.13

1.75

Average

15.22

3.35

1.10

51.41

4.65

3.85

44.44

2.85

1.42

Standard Deviation

1.75

0.59

0.13

2.35

0.18

0.27

3.79

0.35

0.39

LR-B9

LR-B11

119
Table 3.7: Elemental compositions of the humic substances from soils and other
sources (after Tan, 2003)
Carbon

Hydrogen

Oxygen

Nitrogen

(%)

(%)

(%)

(%)

C/N

Tropical Region Soils


HA Alfisols

52.3

5.2

37.2

3.6

14.5

HA Andosols

58.5

3.9

34.8

3.8

15.4

HA Oxisols

54.5

4.4

38.0

3.1

17.5

FA - Andosols

48.9

4.3

44.5

2.3

21.3

Temperate Region Soils


HA Alfisols

56.8

5.0

33.6

4.6

12.3

HA Aridisols

54.5

4.1

36.4

5.0

10.9

HA Histosols

58.7

5.0

32.9

3.4

17.3

HA Inceptisols

51.4

5.4

39.1

4.1

12.5

HA Mollisols

53.7

4.3

36.3

3.7

14.5

HA Spodosols

53.1

4.9

39.2

3.5

15.2

HA Ultisols

50.5

5.2

40.0

3.9

12.9

FA Inceptisols

47.9

5.2

44.3

2.6

18.4

FA Mollisols

41.6

4.0

51.9

1.1

37.8

FA Spodosols

50.6

4.0

44.1

1.8

28.1

FA - Ultisols

45.1

4.5

48.0

2.2

20.5

Reference humic acid

56.2

4.7

35.5

3.2

17.6

Peat

55.5

5.4

33.1

3.0

18.5

120
Table 3.8: Elemental composition of humic substances of Florida Histosols (after
Zelazny and Carlisle, 1974)
Series

Depth

Humin

Humic Acid

Fulvic Acid

(cm)

0-13

56.9

4.4

3.1

0.2

35.4

56.8

5.0

3.4

0.3

34.5

53.3

4.2

2.3

1.9

38.3

13-36

56.2

4.2

3.0

0.2

36.4

56.2

4.6

3.4

0.3

35.5

49.6

6.4

2.4

1.4

40.2

36-81

56.8

6.8

3.1

0.3

33.0

58.8

4.4

3.8

0.1

32.9

51.2

6.4

2.6

1.1

38.7

0-18

56.1

4.8

3.1

0.1

35.9

56.3

4.0

3.6

0.1

36.0

54.8

4.6

1.5

0.2

38.9

18-86

58.5

4.2

2.3

0.2

34.8

56.0

5.2

2.6

0.3

35.9

55.3

4.4

1.3

0.3

38.9

0-21

56.4

4.4

2.7

0.3

36.2

56.6

3.8

3.1

0.3

36.2

56.6

4.6

2.2

0.6

36.0

21-66

57.5

4.8

2.6

0.3

34.8

57.8

4.2

3.3

0.4

34.3

59.6

4.4

2.5

0.3

33.2

0-21

60.7

4.4

3.4

0.5

31.0

56.6

4.4

3.9

0.6

34.5

46.9

7.2

1.2

0.4

44.3

21-53

61.6

4.2

3.4

0.3

30.5

58.0

4.0

3.8

0.3

33.9

52.2

6.8

1.2

0.4

39.4

53-109

58.2

4.6

2.6

0.5

34.1

56.6

5.0

2.8

0.4

35.2

60.8

5.4

2.1

0.6

31.1

109-137

56.6

5.2

2.7

1.1

34.4

57.4

5.2

3.0

0.7

33.7

59.4

4.2

1.6

1.3

33.5

Montverde

Pahokee
Okeetanta
Torry

Table 3.9: Characteristic infrared adsorption frequencies of various organic


functional groups (after Stevenson, 1982)
Assignment

Frequency (cm-1)
3400-3300

O-H stretching, N-H stretching (trace)

2940-2900

Aliphatic C-H stretching

1725-1720

C=O stretching of COOH and ketones (trace)

1660-1630

C=O stretching of amide groups (amide I band), quinine C=O and/or


C=O of H-bonded conjugated ketones

1620-1600

Aromatic C=C, strongly H-bonded C=O of conjugated ketones

1590-1517

COO- symmetric stretching, N-H deformation, +C=N stretching (amide II


band)

1460-1450

Aliphatic C-H

1400-1390

OH deformation and C-O stretching of phenolic OH, C-H deformation of


CH2 and CH3 groups, COO- antisymmetric stretching

1280-1200

C-O stretching and OH deformation of COOH, C-O stretching of


arylethers

1170-950

C-O stretching of polysaccharide or polysaccharide-like substances, Si-O


of silicate impurities

121
Table 3.10: Organic functional groups identified with FT-IR analysis in the three
fractions of the organic matter of LR soil
Fraction

Humin and
Inorganic
component

Humic acid

Frequency
(cm-1)

Organic functional group

3300 - 3400 O-H stretching of phenolic OH or NH


2902
Asymmetric CH stretch of CH22850
Symmetric CH stretch of CH2Asymmetric COO- stretching, N-H deformation, or
1590
+C=N stretching
1450
Symmetric COO- stretching, aliphatic C-H
1085 - 1032 C-O stretching of polysaccharides
3350
O-H stretching of phenolic OH or NH
2917
Asymmetric CH stretch of CH22850
Symmetric CH stretch of CH21720, 1630 Carboxyl acids, amides and ketones
Symmetric stretching of COO-, OH deformation and
1380
C-O stretching of phenolic OH, C-H deformation of CH2
and CH3 groups

1235
1032
3390
2935
1630
Fulvic acid

1400

C-O stretching and O-H bending of -COOH


C-O stretching of polysaccarides
O-CH3 vibration
O-H stretching of phenolic OH or NH
Asymmetric CH stretch of CH2Carboxyl acids, amides and ketones
Symmetric stretching of COO-, OH deformation and
C-O stretching of phenolic OH, C-H deformation of CH2
and CH3 groups

1255

C-O stretching and O-H bending of -COOH

1040

C-O stretching of polysaccarides

122

Figure 3.1: The 30 cm x 30 cm x 30 cm block sampler made of Lexan plates

Figure 3.2: Intact block sample sealed with plastic wrap and wax

123

2.25
LR-B7(April)
LR-B8(April)
LR-B9(July)
LR-B10(July)
LR-B11(July)

2.20

Specific gravity, Gs

2.15
2.10
2.05
2.00

Gs = -0.0104 x LOI(%) + 2.570


R2 = 0.8821

1.95
1.90
1.85
35

40

45

50

55

60

65

70

Loss on ignition (%)

Figure 3.3: Linear relationship between the specific gravity and the loss on
ignition of LR soil

2.6

Specific Gravity, Gs

2.4
2.2
2.0
1.8
1.6
1.4
Muskeg (McFarlane, 1969)
LR soil (current study)

1.2
1.0

10

20

30

40

50

60

70

80

90

100

Organic Content, OC (%)

Figure 3.4: Relationship between specific gravity and organic content (after
MacFarlane, 1969)

124

100
90
80

% finer

70
60
50
40
30

Sand
20

0.1 0.075

Clay

Silt
0.002

0.01

0.001

Particle size (mm)

Figure 3.5: Particle size distribution of LR soil (LR-B9)

600

Atterberg Limits (%)

500

400

LL [April, Disturbed Samples]


LL [July, Disturbed Samples]
LL [April, Block Sample]
PL [April, Disturbed Samples]
PL [July, Disturbed Samples]
PL [April, Block Sample]

300

200

100

0
10

20

30

40

50

60

70

Organic Content, OC (%)

Figure 3.6: Influence of organic content on the Atterberg limits of LR soil

125

(a)

(b)

300
April, Disturbed Samples
July, Disturbed Samples
April, Block Sample
Juturnaiba organic clay

200

180

Plasticity Index, PI (%)

Plasticity Index, PI (%)

250

150

100

50

0
10

20

30

40

50

60

70

Organic Content, OC (%)

PI = 2.21 x OC (%) + 21.0


OC > 10%

160
140
120
100
80
60
40
20
0

20

40

60

Organic Content, OC (%)

Figure 3.7: Influence of organic content on the Plasticity index of (a) LR soil and
(b) Juturnaiba organic clay from Brazil (after Mitchell, 1996)

300

April, Disturbed Samples


July, Disturbed Samples
April, Block Sample

e
-li
n

200

Plasticity Index, PI (%)

250

150

li
A-

ne

100

50

50

100

150

200

250

300

350

400

450

Liquid limit, LL (%)

Figure 3.8: Atterberg limits of LR soil on the plasticity chart

500

126

800

Amorphous Peaty Soils


From Germany

Plasticity Index, PI (%)

700
600

Slightly Fibrous Peat


From Massachusetts

500
400

LR Soil

300
200

Various Types of Peat


(probably fibrous)

100
0

100

200

300

400

500

600

700

800

900 1000 1100

Liquid limit, LL (%)


Figure 3.9: Comparison of the Atterberg limits of LR soil with other amorphous
and fibrous peats (after MacFarlane, 1969)

127

10
9
8

Acidity, pH

7
6
5
4
3
2

H2O
0.01M CaCl2

1
0

20

40

60

80

100

Organic Content, OC (%)


Figure 3.10: Acidity (pH) vs. organic content (after MacFarlane, 1969)

Figure 3.11: Fractionation of particles using centrifuging method

128

100
Whole soil
Inorganic component

90
80

% finer

70
60
50
40
30
20

Sand
1

Clay

Silt
0.1 0.075

0.01

0.002 0.001

0.0001

Particel size (mm)

Figure 3.12: Particle size distribution of the inorganic component of LR soil (LRB9)

129

Inorganic Components

Size fraction
method

wet sieving

> 50 m

< 50 m

centrifugation

50 - 20 m

50 - 2 m

< 2 m

gravity
sedimentation

centrifugation

20 - 5 m

5 - 2 m

2 - 0.2 m

< 0.2 m

Figure 3.13: Fractionation procedure for XRD specimen

130

Figure 3.14: XRD patterns of the inorganic components of LR soil particles in 5


2 m size fraction: from the bottom (a) Mg-glycerated, (b) K-saturated at 21 C,
(c) K saturated and heated at 300 C, and (d) K-saturated and heat treated at
550 C

131

Figure 3.15: XRD patterns of the inorganic components of LR soil particles in 2


0.2 m size fraction: from the bottom (a) Mg-glycerated, (b) K-saturated at 21 C,
(c) K saturated and heated at 300 C, and (d) K-saturated and heat treated at
550 C

132

Figure 3.16: XRD patterns of the inorganic components of LR soil particles


smaller than 0.2 m: from the bottom (a) Mg-glycerated, (b) K-saturated at 21 C,
(c) K saturated and heated at 300 C, and (d) K-saturated and heat treated at
550 C

133

LR soil

fraction
solution

0.1M HCl-HF
Pre-treatment
0.5N NaOH
under N2 blanket
Extraction
(1) Humin and Inorganic
component

Humic and Fulvic acids

5M HCl
Adjust pH to 1.0
Precipitation

In solution

Humic acid

(2) Fulvic acid

0.3N HF:0.1N HCl


Purification
(3) Ash-free Humic acid

Figure 3.17: Organic matter extraction procedure

134

V2
bending

V3
symmetrical
stretching

V1
asymmetrical
stretching

Figure 3.18: Three vibration modes: bending (v2), symmetrical (v3) and
asymmetrical stretching (v1) vibration of water molecules (after Johnston, 1996)

135

Figure 3.19: FT-IR spectrum of the humin and inorganic component of LR soil

136

Figure 3.20: FT-IR spectrum of humic acid of LR soil

137

1725
3275
1620
1200

Suwannee
2960 River HA
IHSS Ref

1395

850

1620
1725
2920
2850

Peat H4
1032

1630
1720
3350

1235
1380

2917
2850

LR soil

4000

3000

2000

1000

Wavenumber (cm-1)
Figure 3.21: Comparison of the FT-IR spectra of humic acid from a Suwannee
River humic acid (after Niemeyer et al., 1992), a peat standard (H4) (after
Niemeyer et al., 1992), and LR soil

138

3000

1600 1200 1000 400

3000

1600

1000

3000

2000 1600

1000

4000

3000

600

1500 1000
1032

E. LR Soil

1630
1720

3350

4000

1235
1380

2917
2850

3000

2000

1000

Wavenumber (cm-1)
Figure 3.22: IR spectrum of the humic acids from four different sources (after Tan,
2003), and LR soil

139

Figure 3.23: FT-IR spectrum of fulvic acid of LR soil

140

A
4000

3000 2000

1500

1000

3000

3000

3000

4000

2000

1500

1000

1500

1000

1000 500

2000

E. LR Soil

1040

3390
1630

1400

2935

4000

3000

1250

2000

1000

Wavenumber (cm-1)
Figure 3.24: IR spectrum of fulvic acids from four different sources (after Tan,
2003) and LR soil

141

CHAPTER 4. EXPERIMENTAL METHODS

4.1. Introduction
The first part of this chapter describes the procedure employed to prepare
reconstituted and treated soil samples. The second part describes the overview
of the Constant Rate of Strain (CRS) consolidation tests and the procedure for
CRS consolidation test set up.

4.2. Experimental Program


In this experimental program, the effects of treatment on the 1-D consolidation
behavior of LR soil were evaluated through constant rate of strain (CRS) and
incremental loading (IL) consolidation tests. These tests yielded compression
curves, compressibility parameters, coefficient of consolidation and hydraulic
conductivity as a function of vertical effective stress. The effects of treatment on
the creep behavior of LR soil were evaluated by performing long-term creep tests
followed by consolidating the specimens at different vertical effective stress
levels in the IL consolidation tests.
The effects of treatment on the strength of LR soil were evaluated by Humphrey
(2001) by performing unconfined compressive strength tests. Humphrey (2001)
observed that among the various types of additives investigated such as ordinary
Portland cement, lime (CaO), High Calcium Flue Dust (HC), Marblehead
Buffington Dust (MB), and Bentonite (BEN), ordinary Portland cement was the
most effective binder to improve the strength of LR soil. Following his results,
type I ordinary Portland cement was selected as the binder to evaluate the
effects on the 1-D consolidation behavior of LR soil.

142
The variables investigated in this experimental program included:
1) Amount of PC: 8, 20, 50 and 100% by dry mass of LR soil
2) Curing time: 14 and 28 days
3) Curing surcharge: 48, 96, 192 kPa.
As mentioned in Chapter 3, three types of LR soil samples were employed for
testing. The intact block soil samples were used to evaluate the 1-D
consolidation behavior of the natural LR soil. Reconstituted soil samples were
prepared with the two disturbed LR soil samples. The disturbed LR soil samples
collected in July, 2001, were used to prepare PC treated soil samples. The
results of the tests performed on the PC treated LR soil were compared to the
results of the reconstituted LR soil to evaluate the effects of treatment. The range
of the variables investigated in the experimental program is summarized in Table
4.1.

4.3. Laboratory sample preparation procedure


The procedure employed in the experimental program to prepare reconstituted
and PC treated LR soil samples was developed originally to produce
homogeneous and reproducible samples for unconfined compression tests
(Humphrey, 2001). The sample preparation procedure consists of four steps:
mixing, compaction, curing and extrusion. The details of each step of the sample
preparation procedure are described in this section
4.3.1. Mixing
The mixing procedure was specially designed to simulate the kneading action
applied to soil in one of the most widely used soil stabilization methods, the Deep
Mixing Method (DMM). The basic concept of the DMM is to improve the
deformation properties and the strength of the soil in situ by mixing it with binding
agents, usually lime or Portland cement. The binding agent can be introduced to
the soil in slurry or grout form (wet method) or in dry form (dry method). In both

143
methods, the binder can be mixed with the soil either by pure rotation of the
mixing tool at relatively low pressure or by the combination of the rotation of the
mixing tool and injection of the binding slurry or grout into the soil at high
pressure (Bruce and Bruce, 2003). In either case, the structure of the soil is
completely destroyed by the rotation of the mixing tool. Although the water
content of LR soil was very high, in the range of 190% for July samples and
289% for April samples, the cement was introduced as slurry (wet method) to
ensure that sufficient amount of water was available for hydration of cement.
Figure 4.1 shows the devices employed to mix soil samples: a KitchenAid stand
mixer and two mixing tools. Two different mixing tools, a flat beater and a hookshaped tool, were used to apply kneading action to the soil during mixing.
Before samples were prepared, the water content of each soil bin was
determined. About 1500 grams of disturbed soil was placed in a mixing bowl, and
deionized water was added to bring the water content of the soil up to the natural
water content (289%). A water content of 289% was used as the natural water
content since this was the water content of the intact block soil samples. The soil
and water were mixed thoroughly for two minutes in the mixer using the flat
beater, and then mixed for five minutes with a hook-shaped mixing tool. The
water contents of the samples were determined after mixing was completed.
Portland cement was added to the soil mixture in a slurry form to prepare treated
soil samples. A water/cement ratio of 0.5 was used to prepare the slurry. After
pouring the cement slurry into the soil, the mixture was prepared following the
same procedure described above: with a flat beater for two minutes and with a
hook-shaped mixing tool for five minutes. Two additional water contents of the
samples were determined after the mixing was completed.

144
4.3.2. Compaction
Following mixing, the soil/cement mixture was compacted into plastic cylinders of
7.62 cm in diameter and 15.24 cm in height (Figure 4.2), employing a dynamic
compaction method with a modified mechanical compactor (Figure 4.3).
Considering the small size of the cylinders used to manufacture the specimen,
the diameter of the hammer was reduced to 2.54 cm. A 6.35 cm extension collar
was attached so that the cylinder could be completely filled for the full length of
the cylinder. A thin layer of vacuum pump oil was applied to the inner wall of the
cylinder so that the sample specimen could be extruded easily. The soil was
compacted into the cylinder using three lifts and applying 40 blows per lift. Once
compaction was completed, the top surface of the sample was leveled with a
metal spatula and the mass of the sample was measured.
To minimize the effects of side friction during the curing stage (see section
4.3.3.), the plastic cylinder was cut in half with a bench saw immediately after
compaction was completed (Figure 4.4).

4.3.3. Curing
A curing surcharge was applied to the compacted soil sample to simulate the
overburden stress and to remove any air bubbles that may have entrapped
during compaction. 48 kPa of surcharge corresponding to approximately 2.5 m of
granular soil fill was used as the base case. For evaluation of the effect of the
curing surcharge, curing was also performed with 96 and 192 kPa surcharge.
A piece of wax paper was placed inside the cylinder on the top surface of the
sample. The sample cylinder was placed at the center of a plastic bin. A concrete
cap of 7.62 cm of diameter and about 3.5 cm high was placed on top of the
sample. The mass of the concrete cap was approximately 500 g. For the curing
surcharge of 48 kPa, one 16 kg, one 4 kg and two 1kg dead weights were placed
on top of the concrete cap. Due to the high water content of the soil samples, the

145
dead weights were balanced with two duct tapes as shown in Figure 4.5. For
reconstituted soil samples, the duct tapes were removed after one day so that
the surcharge can be applied on the soil sample. For PC treated soil samples,
the tapes were removed approximately 2 4 hours after applying the surcharge
so that the surcharge could be applied to the soil before the samples were
hardened. After applying the curing surcharge, the samples were completely
immersed in tap water to ensure continuous access to water as would occur in
the field. A curing time of 14 days was used as the base case. For evaluation of
the effect of curing time, additional tests were performed on 50 and 100% PC
treated LR soil samples cured for 28 days.

4.3.4. Extrusion
After the curing process was completed, the cylinder was removed from the
water bath. Hot water was run over the sides of the plastic cylinder for easier
extrusion of the soil sample. The cylinder was then inverted on a piece of wax
paper, and air pressure was applied carefully through the opening on the bottom
of cylinder to extrude the soil samples from the cylinder. The mass and the height
of the soil sample were measured and the unit weight of the sample was
calculated to check the quality of the sample.
4.4. CRS consolidation test
4.4.1. General overview
The 1-D consolidation behavior of a soil may be investigated by performing
incremental loading (IL) consolidation tests. In IL consolidation tests, the load
doubles with a load incremental ratio (LIR) of one, and each load increment lasts
for 24 hours. From the time settlement curve obtained in each load increment,
the end of primary (EOP) is determined and the hydraulic conductivity and the
coefficient of consolidation are calculated employing Terzaghis consolidation
theory. Although the IL consolidation test has been widely used, there are a few

146
disadvantages. Depending on the loading schedule, a test can take several
weeks to complete. In addition, since only a few discrete data points are obtained
from an IL consolidation test, the preconsolidation pressure can not be estimated
accurately, especially for highly structured soils that exhibit S-shaped
compression curves.
To overcome some of the disadvantages of the conventional incremental loading
(IL) oedometer test, the CRS consolidation test was first introduced by Smith and
Wahls in 1969. Wissa et al. (1971) published the solutions for calculating the
coefficient of consolidation (Cv) and the hydraulic conductivity (k) for CRS
consolidation tests based on both linear and non-linear consolidation theories,
and since then this method has been widely used among researchers. In 1989,
the CRS test was adopted by ASTM as one of the standard methods for 1-D
consolidation (ASTM D4186).
Compared to the IL test, which is a stress-controlled test, the CRS test is a
strain-controlled test, in which the stress-strain relationship of the soil specimen
is obtained by imposing a constant rate of displacement on the soil. The height
and the diameter of soil specimen is the same as those of oedometer specimen
(2.54cm in height and 6.35cm in diameter). The specimen can be saturated
maintaining its volume constant by applying a back pressure from the top and the
bottom. During consolidation, drainage is allowed only at the top and the pore
pressure is measured at the bottom of the specimen with a pore pressure
transducer. One of the advantages of the CRS test is that the test provides
continuous data. The compression curve can be defined as a continuous function
of the vertical effective stress, and thus, the preconsolidation pressure can be
easily defined with Casagrandes graphical construction. The coefficient of
consolidation (Cv) and the hydraulic conductivity (k) can be directly calculated
from the excess pore pressure as a continuous function of the vertical effective
stress or axial strain. While an IL oedometer test may take several weeks to

147
complete, a CRS consolidation test can be completed in much a shorter period of
time, usually within one to two weeks depending on the strain rate employed.
However, since the specimen is continuously consolidated without allowing it to
creep, the creep information may be limited.
In this experimental program, three computer controlled CRS apparatuses: CRS1
- 3, available at Purdues Bechtel Geotechnical Engineering Laboratory were
employed. The main differences of the systems are: (1) the type of seal installed
on the loading piston to prevent leakage from the cell pressure: O-ring for CRS1
and rolling diaphragm for CRS2 and CRS3; and (2) the location of the cell
pressure: at the middle height of the specimen in CRS1 and at the top of the cell
in CRS2 and CRS3. Figure 4.6 shows a schematic of a CRS apparatus which
includes the following four main components: 1) the CRS cell and four sensors
incorporated to measure axial load, axial displacement, cell pressure, and pore
pressure; 2) the load frame used to impose a constant rate of displacement; 3)
the air-water interface system through which the cell pressure is applied; and 4)
the data acquisition system and personal computer.
The soil specimens for CRS testing were prepared employing a procedure and
devices designed for trimming very soft soils with minimal disturbance. The
trimming devices consist of a ring characterized by a sharp cutting edge, a
specimen trimming frame, a thin metal spatula, a cutting blade and a wire saw
(Figure 4.7). The ring was attached to the vertical rod of the trimming frame
through a collar. After curing was completed and the sample was extruded, the
soil sample was placed on the rotating plate of the trimming frame and was
aligned along the centerline of the frame. Trimming was performed by advancing
the ring in small increments. A thin metal spatula was used to trim the soil ahead
of the ring, so that no load could be exerted on the soil as the ring advanced.
After trimming was completed, the excess soil above and below the specimen
was cut roughly with a wire saw. The top and bottom ends of the specimen were

148
carefully trimmed again with a sharp metal knife blade. Two water contents were
determined with samples roughly cut above and below the portion used for the
specimen.
The trimmed specimen was placed on the base of the CRS cell. The CRS cell
incorporates a loading piston, low friction ball bearings, a top cap, and two
porous stones within a Plexiglas chamber that can withstand up to 700kPa of cell
pressure. Figure 4.8 shows a schematic of one of the CRS cells (CRS1). A 8900
N capacity load cell is bolted to the cross head of the loading frame and
measures the axial load transmitted though the loading piston. A DCDT is
connected to the loading piston and measures the displacement of the piston.
The cell and the pore pressure are monitored with 1400 kPa capacity diaphragm
pressure transducers.
Once the CRS cell was assembled, it was mounted on the platen of the load
frame. A constant rate of displacement can be imposed on the specimen by
moving the platen both in the upward and downward directions. The platen is
driven by a high-precision double gear-type motor which can be controlled within
1.175 x 10-6 cm using the PC an operating software. The CRS consolidation test
was performed using the operating software called Sigma-1 CRS.
During the test, signals from sensors were recorded by the data acquisition
system and stored into a data file in the personal computer. The analog signal
(voltage) from each of the sensor was converted to digital signal in the ADIO
(Analog Digital Input Output) module of the 22-bit data acquisition system. The
ADIO module provides excitations (5 or 10 volts) to the sensors connected to it.
With full utilization of its 22-bit resolution, the data acquisition system is able to
discriminate up to 4.77 x 10-3 mV for +/- 10 V input range. To get the best
resolution, the input range was selected to 10.0V for the DCDT and 0.1V for
the Load Cell and two pressure transducers. The resolutions of the data

149
acquisition system for the sensors of CRS1 are shown in Figure 4.9, which
shows the change of the output voltage during 1000 seconds. The Load Cell and
the pressure transducers have resolutions of 0.001 mV (corresponding to 0.287
N, 0.003% FSO and 0.0138 kPa, 0.001% FSO, respectively). The resolution of
the DCDT is 0.1 mV (.232 m, 0.001% FSO). The output signal from each sensor
can be integrated over a time span of 12.5 ms (80Hz) to 3.125 ms (320 HZ). For
the highest stability, the sampling rate was set to 80 Hz. The stability of each
sensor was checked by monitoring the outputs for 24 hours (Figure 4.10). The
range, resolution and stability of each sensor used in the three systems are
summarized in Table 4.2 4.4.
Each of the tests presented in this report was performed following the same
procedure. First the soil was saturated for 24 hours by applying back pressure
through the bottom of the specimen. The back pressure line was hydraulically
connected to the cell pressure line, and thus the same pressure was applied to
the top and bottom of the specimen. The back pressure was increased in small
increments of 7 kPa per minute by manually controlling the air-water interface
pressure regulator. For reconstituted specimens, about 300 kPa of back pressure
was applied, while the magnitude of the back pressure was increased to 450kPa
for PC treated specimens. 0.3% of seating strain was applied to ensure contact
between the loading piston and the specimen. After the seating strain was
applied, the volume of the specimen was kept constant during back pressure
saturation. During the consolidation phase, the back pressure line was closed to
monitor the pore pressure generated at the base of the specimen with the pore
pressure transducer. The excess pore pressure was calculated as the difference
of the pore pressure and the cell pressure at any time t. The maximum load was
maintained for two to three days until at least 95% of excess pore pressure was
dissipated. Then, the specimen was unloaded to 10% of the maximum load at
10% of the displacement rate employed during loading. The rate reduction was
required to prevent generation of excessive negative pore pressure during

150
unloading. The load was again maintained for two to three days until 95% of
negative pore pressure was dissipated. Finally, the specimen was reloaded with
the same displacement rate employed for the consolidation phase until the
maximum target stress was reached.

4.4.2. Procedure for CRS consolidation test


All the CRS consolidation tests were set up following the procedure described
below:
1) Saturate two porous stones and two filter papers for one hour by
sonication
2) Measure the mass of the ring
3) Set up the CRS cell with the ring, two porous stones and two filter
papers. Measure the depth from the top of the cell to the top of the
porous stone three times with a micrometer
4) Finish trimming following the procedure described before
5) Saturate all the lines in the CRS cell and the two transducers with
deionized water
6) Place one porous stone and one filter paper at the bottom of the CRS
cell
7) Set the reading of the pore pressure transducer to zero and record it
into the software
8) Place the specimen ring on top of the filter paper
9) Place one filter paper and the porous stone on top of the specimen
10) Measure the depth from the top of the CRS cell to the porous stone
three times with a micrometer. The specimen height is determined by
subtracting the height measured in step 3 from the height measured in
this step
11) Thoroughly apply vacuum grease to all the O-rings
12) Place the Plexiglas cell, and fully assemble the CRS cell.

151
13) Open all the valves of the CRS cell and fill the CRS cell with deionized
water. In CRS1, set the reading of the cell pressure transducer to zero
when the water is filled to the middle of the specimen. In CRS2 and
CRS3, set the reading of the cell pressure transducer to zero when the
cell is completely filled with deionized water
14) Lock the loading piston after lowering it until it slightly touches the top
of the specimen
15) Set the readings of the DCDT and the load cell to zero
16) Mount the CRS cell on the load platen and align it vertically with the
load cell
17) Raise the load platen until there is about 0.3 mm gap between the
piston and the load cell button
18) Input the testing schedule and the sample height into the operating
software
19) Initiate the seating process by clicking Make Contact button on the
operating software. The load platen moves up until the load increases
to 8 -9 N (or 2 lbs)
20) Once the contact is made, unlock the piston. The reading of the DCDT
measured at this point is used as the actual zero.
21) Start the back pressure saturation stage by clicking the start button.
The load platen moves up until the specimen is strained to the
selected value (0.3% in this experimental program)
22) Increase the back pressure in small increments
23) Once the back pressure saturation stage is completed, click Done on
the operating software and the consolidation tap will be activated
24) Close the valve to the bottom of the specimen so that excess pore
pressures can be measured
25) Start the consolidation test by clicking the Start button

152
Once the test is completed, the water is removed through the opening through
which the cell pressure is applied. The CRS cell is then dissembled, and the
height of the specimen is measured again. The specimen is placed into a steel
container to determine its dry mass. The two filter papers and the two porous
stones are thoroughly washed in a steel container to collect all the soil. Following
the dry mass determination, the oven dried specimen is placed in a furnace at
440 C to determine the loss of ignition.

4.4.3. CRS apparatus calibration


During back pressure saturation, a pressure of the same magnitude is applied to
the top and bottom of the specimen. However, since the cell pressure cannot be
applied to the area occupied by the loading piston at the top of the specimen, a
higher force is applied to the bottom of the specimen (Figure 4.11). The
additional force acting on the bottom of the specimen, called the uplift force, is
transmitted through the loading piston to the load cell. Therefore, the load
measured with the load cell is not the actual load applied to the specimen, and
needs to be corrected to account for the uplift force. In addition, since the loading
piston and the top cap are imposing loads, which can not be measured with the
load cell, their weights should also be added to the axial load to obtain the actual
load.
To evaluate the uplift force, the CRS cell was set up but without a specimen. The
load platen was adjusted until about 5.0 N of load was applied before applying
the cell and the back pressure to ensure the contact between the piston and the
load cell. The cell pressure was increased in small increments to about 500 kPa
while monitoring the axial load and the cell pressure.
As shown in Figure 4.12, there is a linear relationship between the cell pressure
and the axial load. The slope of the linear best fitting curve of these data points
represents the piston area and the y-intercept represents the piston weight. The

153
piston areas and the piston weights of the three systems are summarized in
Table 4.5. With these two parameters, the actual load can be calculated as
follows:

Actual load = Axial load Piston area * Cell pressure + Piston weight
During a consolidation test, the load frame, the loading piston, the filter papers
and the porous stones deflect as a function of the actual load. In the current
configuration of the apparatus, the deformations of the specimen are measured
at the top of the CRS cell as the relative change in the position of the DCDT
clamped onto the loading piston. Therefore, the DCDT measurements include
the deformation of the specimen as well as the deflection of the whole system
due to compression. To obtain the actual deformation of the specimen, the
compliance (machine deflection) must be quantified and the data obtained from
the DCDT should be corrected.
To evaluate the compliance of the system, the cell was set up with a steel
dummy specimen with the same dimensions as the soil specimens. The cell was
filled with de-aired and de-ionized water but no cell pressure was applied. In this
way, the load measured with the load cell corresponds to the actual load applied
to the specimen. Once the cell was filled, a seating load of about 0.5 N was
applied. The CRS cell was loaded and unloaded at a constant rate while
measuring the load and the displacement. Given its stiffness it is calculated that
the steel specimen deforms by about 3.39 x 10-4 mm (0.001% axial strain) under
the maximum applied load (8900 N), thus it is assumed as incompressible. The
displacement measured during calibration corresponds to the systems
deformation.
Figure 4.13 shows the compliance of the system during loading and unloading as
a function of the actual load, and the best fitting curves obtained using a power
function. The coefficients of the power function used to estimate the machine

154
deflection of each system are summarized in Table 4.5. Under the maximum
load, the system deflects about 0.15 0.75 mm (0.6 3.0% axial strain).
Although the machine deflection curve has hysteresis during loading and
unloading, as shown in Figure 4.13, the difference between the measured
machine deflection and the value calculated with the power function is very small
in the range of 0.02 0.04 mm (0.08 0.15% axial strain). The best fitting curve
for loading was used to correct the vertical displacement for the compliance of
the systems during loading and unloading.

4.4.4. Data reduction


As described in the previous section, the readings (voltages) from each sensor
and the corresponding reading time were recorded by the data acquisition
system and stored into a data file during CRS test. After completion of the test,
the data was reduced for analysis using a data reduction program written in
Visual Basic and embedded in a Microsoft Excel file. Two options are available
for reduction of the CRS data: the linear and the non-linear solution (e.g. Wissa
et al., 1971). These solutions are described in the following section
The sequence of data reduction is shown as a flow chart in Figure 4.14. First, the
actual load is calculated by correcting the axial load for uplift as a function of the
cell pressure using the theoretical values of piston area and piston weight. The
actual displacements of the specimen (H) are calculated by correcting the
displacements for the compliance as a function of the actual load. The axial
strain, total stress and vertical effective stress are computed from the actual load,
actual displacement and excess pore pressure. These calculations are based on
the initial specimen height and the specimen area.
During consolidation, three strain rates were calculated: the axial strain rate
based on the initial specimen height, the current strain rate based on the current

155
specimen height, and the encoder strain rate based on the encoder that monitors
the relative movement of the load platen. The starting and the ending points of
each phase of a CRS consolidation test were determined from the encoder strain
rate.

4.4.5. CRS consolidation theory


In 1971, Wissa et al. developed solutions for CRS tests based on both linear and
non-linear stress-strain relationships including the initial transient and steady
states. The basic assumption of the solutions for CRS tests is that infinitesimal
strains can be used and that the coefficient of consolidation (Cv) is constant and
is independent of the depth at any time, t.
Based on the assumptions, the basic equation of consolidation is expressed as

2
Cv 2 =
z
t

(4.1)

where Cv = k mv w ; t = time; z = the vertical coordinate of a point from the top of


the specimen.
The solution of equation (4.1) for the strain at any point, z and time, t is:

( z , t ) = t 1 + F ( z , t )
2
1
z
2
z
in which F ( z , t ) =
2 6 + 3 2
6Tv
H
H Tv

and, Tv =

(4.2)

cos n z
exp ( n 2 2Tv )
2
n
H
n =1

Cv t
H2

The first part of function F represents the deviation from the average strain in the
steady state, and the second part describes the decay of the initial discontinuities
in the transient state. The transient component becomes insignificant by the time
Tv becomes 0.5.
The data obtained from a CRS tests include, as described in the previous
section, the total stress, v, measured at the top of the specimen (z = 0), which is

156
constant throughout the specimen, the excess pore pressure, uh, measured at
the bottom of the specimen (z = H), and the average strain rate, .
In order to interpret the CRS test data, it is necessary to assume a relationship
between stress and strain, and the following two possibilities are considered:
linear and non-linear stress-strain relationship.

4.4.5.1. Linear theory


The basic assumption of the linear theory is that the coefficient of volume
compressibility, mv, is constant. In other words, the change of effective stress is
linearly proportional to the change of strain, and the relation can be expressed
as:

'v =

mv

(4.3)

The transient state occurs at the start of loading or throughout the duration of a
test performed at higher strain rates in which large excess pore pressure is
generated. The transient conditions can be interpreted from equation (4.2). At
any time t, the ratio of the strain at the bottom (z = H) and at the top (z = 0) of
the specimen can be expressed as:

( H ,t ) 1+ F ( H ,t )
=
= F3 (Tv )
( 0, t ) 1 + F ( 0, t )

(4.4)

For a linear material, since the change of strain is proportional to the change of
effective stress from time 0 to t, equation (4.4) can be expressed as

F3 =

'v ( H , t ) 'v ( H , 0 ) ( v uh ) v ,t =0
=
'v ( 0, t ) 'v ( 0, 0 )
( v ) v ,t = 0

(4.5)

The function F3 is dependent only on Tv. and F3 can be plotted as a function of


Tv (Figure 4.15). Sheahan et al. (1997) proposed an equation for Tv after
performing a regression analysis on the plot of Tv versus F3 as follows:

157

Tv = 4.78 ( F3 ) 3.21( F3 ) + 1.65 F3 + 0.0356


3

(4.6)

If Tv is smaller than 0.5 at any time t, the specimen is assumed to be in the


transient state. Once the value of F3 is obtained from CRS tests data and
equation (4.5), Tv can be calculated from equation (4.6), and the coefficient of
consolidation and the hydraulic conductivity in the transient state can be
calculated using Terzaghis theory as follows:

Cv =

Tv H 2
t

(4.7)

k = Cv mv w

(4.8)

If Tv becomes greater than 0.5, the soil is assumed to be in the steady state, and
the second term of function F(z ,t) in equation (4.2) can be neglected. In the
steady state, the strain is parabolically distributed through the depth of the
specimen. Therefore, the average strain can be obtained as follows:

ave = 2 ( H , t ) + (0, t ) = t
3

(4.9)

and the average effective stress corresponding to the average strain is

'vave = 2 'v ( H , t ) + 'v ( 0, t ) = v uh


3
3

(4.10)

The hydraulic conductivity of a linear soil in the steady state can be directly
calculated from the difference of the stress and strain at the top and bottom of
the specimen. The strain and the effective stress at the top and bottom of the
specimen at any time t are:
At top: ( 0, t ) = t +

H2

'v = v

3Cv

At bottom: ( H , t ) = t

H2
6Cv

'v = v uh

Using equations (4.3) and (4.8), the difference of strain and stress at the top and
bottom of the specimen is expressed as:

158

( v u h ) v =
uh =

H2
2Cv mv

( 0, t ) ( H , t )
mv

H 2 w

(4.11)

2k

Rearranging equation (4.9) for the hydraulic conductivity, k:

k=

H 2 w

(4.12)

2uh

Also from equation (4.11), the coefficient of consolidation is:

Cv =

H2

(4.13)

2mv uh

4.4.5.2. Non-linear theory


The assumption for non-linear theory is that the difference of strain is related to
the difference of the logarithm of effective stress by a constant C called the
strain index as:

d
= C
d ( log 'v )

(4.14)

Since the change in strain is proportional to the change in logarithms of the


effective stress, the function F3 is expressed as:

F3 =

log 'v ( H , t ) log 'v ( H , 0 )


log 'v ( 0, t ) log 'v ( 0, 0 )

log ( v uh ) log ( v ,t =0 )
log ( v ) log ( v ,t =0 )

(4.15)

From the value F3, Tv can be estimated using equation (4.8). If a transient
condition exists (Tv < 0.5) at any time t, k and Cv can be calculated from the value
Tv using equations (4.7) and (4.8), respectively.
In the steady state, using effective stresses and strains at the top and bottom of
the specimen at any time t, equation (4.14) can be approximated as:

159

( 0, t ) ( H , t )
=
log ( v ) log ( v uh )

H 2
= C
v uh
2Cv log

(4.16)

Similarly, by comparing the stress and the strain at two different times, t1 and t2,
at the top of the specimen yields:

( 0, t1 ) ( 0, t2 )
=
log ( v1 ) log ( v 2 )

t
= C
v2
log

v1

(4.17)

Combining equations (4.18) and (4.19), the coefficient of consolidation can be


obtained as:

H 2 log v 2
v1
Cv =
u
2t log 1 h
v

(4.18)

The coefficient of volume compressibility, mv, can be computed from equation


(4.17) as:

mv =

0.434C
'v

(4.19)

in which v is the average effective stress between time t1 and t2.


From equations (4.18) and (4.19), the hydraulic conductivity can be calculated
as follows:

k = Cv mv w =

0.434 H 2 w
u
2 'v log 1 h
v

(4.20)

For a non-linear soil, the distribution of strain in steady state is parabolic and the
average strain can be calculated using equation (4.9). Then, the average
effective stress corresponding to the average strain is calculated from the
following equation:

'vave = ( v 3 2 v 2uh + v uh 2 )

(4.21)

160
4.4.5.3. Linear vs. Non-linear theory
The difference of the results obtained from the linear and non-linear theories can
be evaluated by comparing the equations for the vertical effective stress, the
hydraulic conductivity and the coefficient of consolidation.
From equations (4.10) and (4.21), the ratio of the vertical effective stresses from
the linear and non-linear theories is

'v linear
'v non linear

2
3

v uh

3
v

2 v uh + v uh
2

=
3

2 uh
3 v

u u
1 2 h + h
v v

(4.22)
3

The ratio of the vertical effective stresses from the linear and non-linear theories
is plotted against uh/v in Figure 4.16. The plot shows that for positive excess
pore pressure, the linear theory yields greater vertical effective stress than the
non-linear theory, and that the divergence between the two theories increases
with the ratio of uh/v.
Similarly, the ratios of the hydraulic conductivity and the coefficient of
consolidation can be derived as, using v = v for small uh:

H 2 w
klinear
knon linear

2uh
0.434 H 2 w

u
2 'v log 1 h
v

u
log 1 h
v
=
u
0.434 h

(4.23)

and,

H2
Cv linear
Cv non linear

2mv uh
0.434 H 2

u
2 'v mv log 1 h
v

u
log 1 h
v
=
u
0.434 h

(4.24)

161
Since the hydraulic conductivity and the coefficient of consolidation are related to
each other through equation (4.8), the ratios of these are the same. The ratio of
the coefficient of consolidations obtained from the linear and non-linear theories
is plotted against uh/v in Figure 4.17. The ratio of the coefficient of consolidation
is similar to the ratio of the vertical effective stress. The linear theory produced a
greater value of the coefficient of consolidation than the non-linear theory. The
difference between the two theories diverges as the ratio uh/v increases, but the
divergence of the coefficient of consolidation is greater than that of vertical
effective stress. It should be noted that this comparison is valid only for small
values of uh/v so that the constrained modulus from the linear and non-linear
theories can be cancelled out and the vertical effective stress can be replaced
with the total stress equations (4.23) and (4.24).
In this experimental program, the data was reduced employing the non-linear
CRS consolidation theory. Wissa et al. (1971) recommended to use linear CRS
Consolidation theory when the strain rate is slow enough to yield a pore pressure
ratio smaller than 0.05. As will be discussed in the following chapters 5 and 6, the
strain rate employed in this experimental program yielded pore pressure ratio
values significantly higher than 0.05 but smaller than 0.30.

162
Table 4.1: Ranges of the testing variables investigated
Soil sample
type

Amount of PC

Curing time

Curing surcharge

(%PC)

(days)

(kPa)

April, 2001

April, 2001

14

48

14

48

14

48

20

14

48

50

14, 28

48, 96, 192

100

14, 28

48

Sampling time

Intact, block

Disturbed

July, 2001

Table 4.2: Resolution and stability of the sensors used in CRS1 system
Sensor

I.D.

DCDT

0243-0141

Load Cell
Pore Pressure

IC18824
8162-014

Transducer
Cell Pressure
Transducer

8162-024

Range

Resolution

Stability (24 hour)

0.232 m

3.016 m

(0.1mV)

(1.3mV)

0.287 N

2.296 N

(0.001 mV)

(0.008 mV)

1379 kPa 0.0138 kPa

0.1104 kPa

(0.001mV)

(0.008mV)

1379 kPa 0.0138 kPa

0.114 kPa

(0.001mV)

(0.008mV)

2.54 cm
8900 N

163
Table 4.3: Resolution and stability of the sensors used in CRS2 system
Sensor

I.D.

DCDT

0244-0177

Load Cell
Pore Pressure

102116
PS2036

Range

Resolution

Stability (24 hour)

0.308 m

4.004 m

(0.1mV)

(1.3mV)

0.479 N

3.382 N

(0.001mV)

(0.008mV)

1379 kPa 0.0138 kPa

0.1104 kPa

(0.001mV)

(0.008mV)

1379 kPa 0.0138 kPa

0.1104 kPa

(0.001mV)

(0.008mV)

5.08 cm
8900 N

Transducer
Cell Pressure

PS2037

Transducer

Table 4.4: Resolution and stability of the sensors used in CRS3 system
Sensor
DCDT
Load Cell
Pore
Pressure

I.D.

Resolution

Stability (24 hour)

0.232 m

3.016 m

(0.1mV)

(1.3mV)

0.479 N

3.382 N

(0.001mV)

(0.008mV)

PS2149 1379 kPa 0.0138 kPa

0.1104 kPa

(0.001mV)

(0.008mV)

PS2150 1379 kPa 0.0138 kPa

0.1104 kPa

(0.001mV)

(0.008mV)

D127
164028

Range
2.54 cm
8900 N

Transducer
Cell
Pressure
Transducer

164
Table 4.5: Coefficients for CRS system calibration
Uplift force

Machine deflection*

Piston area (m2)

Piston weight (N)

CRS1

0.1514

-8.712

0.0072

0.4154

CRS2

0.3141

-6.829

0.0021

0.4749

CRS3

0.2916

-5.691

0.0078

0.5028

*Machine deflection = a x (net load)

165

Figure 4.1: Mixer with two different mixing tools employed to prepare samples for
CRS consolidation tests.

Figure 4.2: A plastic cylinder and modified compaction mold employed for
compaction.

166

Figure 4.3: Modified mechanical proctor compactor employed for treated LR soil
samples: (1) original and (2) modified compaction hammer

Figure 4.4: Cutting of plastic cylinder with Wells metal band saw for preparation
of soil samples for CRS consolidation tests

167

Figure 4.5: Curing of soil specimen in water bath under a surcharge of 48 kPa

168

Pressure
regulator

CRS Cell
Load Platen

Load Frame

PC
DAS

Figure 4.6: Schematic of the four components of a CRS consolidation test


apparatus; CRS Cell, Load Frame, Air-Water Interface Pressure Regulator and
Data Acquisition System (DAS) (connecting wires behind PC, DAS, and load
frame not shown)

169

Figure 4.7: Specimen trimming apparatus: trimming frame, a thin spatula, cutting
blade and a wire saw.

170

Load Cell
Low friction
Ball bearing

DCDT

Air Vent

O-ring

Top Cap

Pore Pressure
Transducer

Pressure
regulator

Cell Pressure
Transducer

Porous Stone

Soil Specimen
Porous Stone
Cell Pressure Line
Pore Pressure Line

Back Pressure Line

Figure 4.8: Schematic of CRS consolidation cell (CRS1)

171

Figure 4.9: Resolution of the Data Acquisition System and sensors (CRS1)

172

Figure 4.10: Stability of the sensors

173

Piston weight
Cell pressure

Cell pressure

Backpressure

Soil Specimen

Figure 4.11: Uplift calibration for CRS cell (The area subjected to cell or
backpressure is indicated with a thick line)

174

70

Axial Load (N)

60
50
40
30

CRS1
Piston Area = 0.151 m2
Piston Weight = -8.712 N
r2 = 0.9976

20
10
0

100

200

300

400

500

Cell Pressure (kPa)


Figure 4.12: Calibration for uplift force

0.00

CRS1

Displacement (mm)

0.05
0.10

Loading
Deflection = 0.0072*(Load)0.4154

0.15

r2 = 0.9815
0.20
0.25

Unloading
Deflection = 0.0154*(Load)0.3306

0.30

r2 = 0.9802
0.35

2000

4000

6000

Actual Load (N)


Figure 4.13: Calibration for machine deflection (compliance)

8000

175

DCDT

Load Cell

CPT

PPT

Displacement

Appled Load

Cell Pressure

Pore Pressure

Uplift Force

uh

v
Machine Deflection

Time

Strain Rate

Figure 4.14: Flow chart for data reduction

'v

176

1.0
0.9
0.8

Time Factor, Tv

0.7
0.6
0.5
0.4
0.3
0.2
Tv = 4.78F33-3.21F32+1.65F3+0.0356
0.1
0.0
0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

F3
Figure 4.15: Plot of the function F3 as a function of Tv

177

3.5

'v linear / 'v non-linear

(a)
3.0

2.5
2.0

1.5

1.0
0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

Pore Pressrue Ratio, uh/v

1.06
(b)
'v linear / 'v non-linear

1.05
1.04
1.03
1.02
1.01
1.00
0.0

0.1

0.2

0.3

0.4

0.5

Pore Pressrue Ratio, uh/v

Figure 4.16: Comparison of vertical effective stress computed using linear and
non-linear CRS theory (a) 0<uh/v<1.0 and (b) 0< uh/v <0.5

178

3.5

Cv linear / Cv non-linear

(a)
3.0

2.5
2.0

1.5

1.0
0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

Pore Pressrue Ratio, uh/v

1.4

Cv linear / Cv non-linear

(b)
1.3

1.2

1.1

1.0
0.0

0.1

0.2

0.3

0.4

0.5

Pore Pressrue Ratio, uh/v

Figure 4.17: Comparison of coefficient of consolidation computed using linear


and non-linear CRS theory (a) 0<uh/v<1.0 and (b) 0< uh/v <0.5

179

CHAPTER 5. 1-D CONSOLIDATION BEHAVIOR OF LINDBERG ROAD SOIL

5.1. Introduction
This chapter presents and discusses the results of consolidation tests (CRS and
incremental loading tests) performed on both reconstituted and intact LR soil.
The main three objectives of this chapter are: (1) to characterize the 1-D
consolidation behavior of the intact LR soil; (2) to compare the 1-D consolidation
behavior of the intact LR soil with that of other soils (i.e. fibrous peats to inorganic
clays); and (3) to characterize the 1-D consolidation behavior of the reconstituted
LR soil as a base case for evaluating the degree of structuring of the intact soils
as well as to measure the effects of cement, as discussed in the following
chapter. The test results for the reconstituted specimens serve as a base case
for comparison with the results for the PC treated soil specimens due to the fact
that in field applications of deep mixing, the existing soil structure is destroyed by
rotation of the augers. Any subsequent increase strength and stiffness originates
from the chemical reactions of the binding agents. It is this contribution that the
comparison of the data for the treated soil and the reconstituted soil provides the
means to isolate.
Section 5.2 presents the results of CRS consolidation tests performed on
reconstituted LR soil specimens prepared from disturbed soil samples collected
in April of 2000. One CRS consolidation test is selected to describe the typical
results (compression curve, coefficient of consolidation, hydraulic conductivity)
obtained for this material. Following that, the results of three CRS consolidation
tests are compared to evaluate the repeatability of the results.

180
Section 5.3 presents the results of CRS consolidation tests performed to assess
the strain rate dependency of the reconstituted soil. The section presents results
obtained from tests performed on both the April (0.25%/hr 1.0%.hr) and July
(0.l%/hr 1.0%/hr) specimens. The effects of strain rate on the preconsolidation
pressure, compressibility, generation of excess pore pressures, pore pressure
ratio, hydraulic conductivity and coefficient of consolidation are discussed.
Section 5.4 presents the results of CRS consolidation tests on intact LR soil. The
1-D consolidation behavior of this soil is compared to the results obtained for the
reconstituted material (both April and July specimens) to evaluate the role and
the effects of the structure existing in the natural intact soil. The destructuring
process of intact LR soil during consolidation is investigated based on the model
proposed by Liu and Carter (1999, 2000).
Throughout the chapter the results of intact and reconstituted LR soil are
compared with those of peat deposits and other selected soils chosen for
reference (e.g. Boston blue clay, a typical inorganic marine clay) to highlight the
effects of organic matter on the 1-D consolidation behavior.
Section 5.5 presents the data from incremental loading consolidation tests
performed to assess the secondary consolidation index (C) and the C/Cc ratio
of LR soil.
5.2. 1-D consolidation behavior of reconstituted LR soil
A critical factor in performing CRS consolidation tests is the strain rate selection
since preconsolidation pressure, compressibility and the magnitude of excess
pore pressures depend on the strain rate. To select the most appropriate strain
rate, five pilot tests were performed on reconstituted LR soil specimens during
the early stages of the experimental program with two different strain rates:
0.5%/hr and 1.0%/hr. Of these two strain rates, only the 0.5%/hr strain rate
yielded pore pressure ratios smaller than 0.3 and satisfied the ASTM criteria.
Based on the results of these pilot tests, the 0.5%/hr strain rate was selected for

181
the experimental program. The strain rate dependency of LR soil was further
evaluated by performing additional CRS consolidation tests with strain rates
ranging from 0.1%/hr to 0.25%/hr, 0.5%/hr and 1.0%/hr. The results of these
tests will be further discussed in section 5.3.
Data for the reconstituted LR soil samples used to manufacture the specimens
for this part of the experimental program are summarized in Table 5.1, including
bin number, sampling date, curing surcharge, curing duration, initial water
content (wi), loss on ignition (LOI), and initial void ratio (ei). Table 5.2 summarizes
the testing parameters (strain rate and stress ranges) for each of the CRS
consolidation tests performed on reconstituted LR soil. For example, during test
CRS007, the soil specimen was consolidated up to a vertical effective stress of
998.9 kPa with a strain rate of 0.5%/hr. Once the target stress was reached, the
specimen was allowed to dissipate under constant total stress more than 95% of
the excess pore pressure developed. After creep, the specimen was unloaded to
a vertical effective stress of 85.7 kPa with a strain rate of 0.5 %/hr (unload rate
factor = 1). The specimen was again allowed to dissipate the negative excess
pore pressure. Finally, the specimen was reloaded up to the final stress of
2005.8 kPa with a strain rate of 0.5%/hr. For CRS030 and CRS046, the
unloading strain rates were reduced to 0.1%/hr (unload rate factor = 0.1) to
prevent generation of excessive negative excess pore pressures, which
influenced swelling and reloading indexes. The effects of high negative excess
pore pressures on unloading and reloading will be described in detail later.
5.2.1. Compression behavior and generation of excess pore pressure
Figure 5.1 a-b present the compression curves obtained from CRS007 performed
on reconstituted LR soil with the vertical effective stress plotted on (a) log-scale
and (b) arithmetic scale, respectively. As shown the CRS consolidation test
consisted of five main stages. The starting and ending points of each stage of
CRS consolidation test are indicated in Figure 5.1 a-b with letters denoting:
loading (A-B), first dissipation (B-C), unloading (C-D), second dissipation (D-E),

182
and reloading (E-F). Similar stages were performed in all other CRS tests (see
Table 5.2). The unloading and reloading were applied to assess the unloading
and reloading indices of LR soil. In CRS consolidation tests, positive excess pore
pressures are generated during loading. If pore pressures are not allowed to
dissipate before unloading, the positive excess pore pressures prevent the soil
specimens from swelling during unloading. This would cause a flatter shape of
the unloading curve during the early stage of unloading and an underestimation
of the swelling index. Similarly, the specimens are allowed to dissipate negative
excess pore pressures generated during unloading prior to reloading so that the
specimens can deform in compression from the beginning of reloading.
The compression curve of reconstituted LR soil exhibits the characteristics of
non-structured soft soils: stiffer response in the recompression range, a well
defined break corresponding to the preconsolidation pressure (p), and steeper
and linear slope in log scale along the virgin compression line (Figure 5-1 a).
Figure 5-1 b indicates significant strain hardening behavior in the normally
consolidated (NC) range.
The preconsolidation pressure of each specimen was calculated employing the
strain energy method (Becker et al, 1987). In this method, the work per unit
volume is used as a yield criterion to define the preconsolidation pressure, at
which the change from small strain response to large strain response takes
place. As shown in Figure 5.2, the preconsolidation pressure for CRS007 test,
estimated as the intersection of pre-yield and post-yield lines, was 57.0 kPa. The
preconsolidation pressure obtained with the strain energy method was consistent
with the value obtained using Casagrandes graphic construction as shown in
Figure 5.3.
When strain energy method is employed, however, care should be exercised to
prevent any systematic error. This is especially true for LR soil, which exhibits

183
strong strain hardening behavior in the NC region as shown in Figure 5-1 b. If the
entire NC range is used for post-yield line, the strain energy method would cause
over-estimation of preconsolidation pressure, up to 90 kPa.
For reconstituted LR soil, it was observed that using vertical effective stress in
the range of 100 200kPa (~2 4 p) for the post-yield line and vertical effective
stress in the range of 0 20 kPa (OCR >2 3) for the pre-yield line provided
preconsolidation pressure values that were in good agreement with
Casagrandes method.
It should be noted that the preconsolidation pressure of CRS007 is greater than
the curing surcharge of 48 kPa applied to the soil specimen during 14 days.
According to Mesri and Castro (1986), soft soils develop preconsolidation
pressure as a result of secondary compression, also known as aging effect. The
overconsolidation resulting from secondary compression depends mainly on the
aging time and the ratio of C/Cc of the soil and can be estimated using the
following equation:

t
'
OCR = vc =
'vi t p

( C

Cc )(1 Cr Cc )

(5.1)

where vc = preconsolidation pressure developed from secondary


compression

vi = consolidation pressure at which secondary compression takes place


t = total duration of time that the load is maintained
tp = time required for completion of primary compression
C = coefficient of secondary compression
Cc = compression index, and Cr = recompression index.
For reconstituted LR soil, vi = 48 kPa, the total curing time (t) = 14 days, C/Cc=
0.102 for July soil and 0.095 for April soil (see discussion later in section 5.5.3),

184
and Cr/Cc = 0.2 (this value is reported later in this section). The tp value is not
measured directly during curing, but can be estimated with the Cv data at vi =
48 kPa (i.e. Cv = 2 - 4 x 10-4 cm2/sec from Figure 5.27 and Figure 5.28) using
Terzaghis consolidation theory. The height of the soil sample prior to curing was
approximately 7 8 cm. With the Cv value of 2 - 4 x 10-4 cm2/sec and the
specimen height, the estimated time tp is approximately 7 18 hours. Using
these values in equation (5.1), the preconsolidation pressure for reconstituted LR
soil due to secondary compression can be estimated in the range of 60 65 kPa,
which compares well with the stress value observed.
Reconstituted LR soil exhibits a slightly S-shaped compression curve. The
compression index of the soil increased from 1.492 (in the 2 4 p effective
stress range) to 1.676 (4 8 p) followed by slight but continuous decrease to
1.603 (8 16 p) and 1.472 (16 32 p). Compared to typical soft soils, the
compression index of reconstituted LR soil lies between that of typical inorganic
marine clays (e.g. Cc = 0.342 0.399 for Resedimented Boston Blue Clay,
Sheahan et al, 1997; Cc = 0.20 0.30 for undisturbed Boston Blue Clay, Mesri
and Ali, 1999) and fibrous peats (e.g. Cc = 4.41 14.93 for a fibrous Muskeg
from Quebec, Canada, with organic content = 70 100%, Lefebvre et al, 1984).
The excess pore pressure developed at the base of the specimen during loading
is shown in Figure 5.4. The magnitude of the excess pore pressures depends on
the strain rate, hydraulic conductivity (k), void ratio and stiffness of the soil
specimen. In the over consolidated region, generation of excess pore pressures
is negligible due to the high void ratio, k, and stiffness. The excess pore
pressures measured at the preconsolidation pressure in CRS007 was 6.26 kPa.
In the normally consolidated region, the rate of excess pore pressure generation
increases significantly as the hydraulic conductivity and stiffness decrease. For
CRS007 at the end of the loading stage when the vertical effective stress is
998.9 kPa, the excess pore pressure has reached a value of 220.3 kPa.

185
The substantial increase in excess pore pressures in the normally consolidated
region is in part also a result of the increase in the actual strain rate. In the CRS
consolidation test apparatus employed in this experimental program, as
mentioned in Chapter 4, the load platen advances upward during loading at a
constant displacement rate, which is calculated based on the initial specimen
height and the strain rate selected prior to testing. This implies that the actual
strain rate imposed on the specimen increases as the specimen height
decreases. The increase in actual strain rate with decrease in specimen height is
illustrated in Figure 5.5. It should be noted that the platen advance rate was
maintained constant (corresponding to a nominal rate of 0.5%/hr) throughout the
loading stage. As shown in the figure, at the beginning of loading, the actual
strain rate is slower than the platen advance rate because the axial deformation
was corrected for machine deflection (see section 4.4.2.). The effects of the
increase in actual strain rate on generation of excess pore pressure will be
discussed more in detail in section 5.2.
The pore pressure ratio, which is the ratio between the change in excess pore
pressures (uh) and the change in total stress (v), during loading is shown in
Figure 5.6. In the over consolidated region (v < 57.0 kPa), the pore pressure
ratio remains constant or slightly decreases within the range of 0.08 and 0.10. As
the soil specimen is loaded into the normally consolidated region, the pore
pressure ratio increases continuously reaching the maximum value of 0.2 at a
vertical effective stress of 998.9 kPa.
Once the target stress was reached, the soil specimen was allowed to dissipate
approximately 95% of excess pore pressures. The dissipation stage lasted from
24 to 96 hours depending on strain rate and target stress. More time was
required when the strain rate was fast or the target stress was high. As will be
discussed in section 5.3, a faster strain rate yield generation of higher excess
pore pressures since pore pressures accumulated faster. The void ratio and

186
hydraulic conductivity of a soil decreases with increasing vertical effective stress.
Thus, if a soil is loaded to a higher vertical effective stress level, the dissipation of
excess pore pressures requires more time due to lower hydraulic conductivity.
Figure 5.7 presents time settlement and excess pore pressure dissipation curves
obtained from CRS007. At a vertical effective stress of 998.9 kPa and a void ratio
of 1.770, 72 hours were required to dissipate excess pore pressures from 220.3
kPa to 11.3 kPa (95% of dissipation). During dissipation, the load platen moves
continuously up and down to keep the total stress constant. Therefore, as the
excess pore pressures dissipate the effective stress increases. In CRS007 during
the dissipation stage, the effective stress increased from 998.9 kPa to 1129.6
kPa
After dissipation of the excess pore pressure, the soil specimen was unloaded to
a vertical stress level to 10% of the target loading stress. In CRS007, the
specimen was unloaded at 0.5%/hr, the same strain rate employed for loading.
The ASTM standard recommends to reduce the unloading strain rate to 10% of
the loading strain rate, which may be related to the fact that the swelling index of
most soil is about 1/5 1/10 of the compression index. The fast unloading rate
used in CRS007 caused generation of significant negative excess pore
pressures, with the maximum negative excess pore pressure and minimum pore
pressure ratio being -227.3 kPa and -12.2, respectively.
The swelling index (Cs) of CRS007 increased from 0.059 (1 2 OCR), to 0.106
(2 4 OCR) and 0.143 (4 8 OCR), followed by a decrease to 0.113 (8 12
OCR). The decrease in swelling index at OCR > 8 is not considered to represent
true soil behavior but is instead an artifact of the fast unloading rate. As will be
discussed later, CRS consolidation tests with reduced unloading rate showed
continuous increase in swelling index with increasing OCR.

187
After unloading, the total stress was kept constant and the soil specimen was
allowed to dissipate negative excess pore pressures. Figure 5.8 shows timesettlement and dissipation of negative excess pore pressures. In CRS007, the
excess pore pressure dissipated to -6.6 kPa (97% of dissipation) after 72 hours
of creep. During this creep stage, the effective stress decreased from 85.7 kPa to
17.7 kPa.
After the second dissipation, the soil specimen was reloaded to the final vertical
effective stress of 2000 kPa, at the same strain rate employed for loading. The
reloading index (Cr) increased with decreasing OCR from 0.071 (12 8 OCR), to
0.191 (8 4 OCR), 0.297 (4 2 OCR) and 0.428 (2 1 OCR). As shown in
Figure 5.1 a), reconstituted LR soil developed a quasi-preconsolidation pressure
due to mechanical unloading/reloading. The level of quasi-preconsolidation could
not be measured with the strain energy method, since the specimen was not
further loaded into the normally consolidated region. However, a rough estimate
using Casagrandes method indicates that the quasi-preconsolidation pressure is
approximately 1600 1650 kPa. Similar results were observed in peats
(Middleton peat, Mesri et al, 1994) and soft clays (Nagasaka clay ,Liu and Carter,
2000). The reconstituted soil merged into the virgin compression line, shown as a
dashed line in Figure 5.1 a), at a vertical effective stress of 2000 kPa.
The generation of excess pore pressures during reloading is shown in Figure 5.9.
The magnitude of excess pore pressures during reloading was significantly
higher than during loading. This is due to the combined effects of the lower
hydraulic conductivity and the higher actual strain rate imposed on the soil
specimen during reloading. Figure 5.10 presents the actual strain rate during
loading and reloading as a function of vertical effective stress. As shown in the
figure, the actual strain rate imposed on the specimen during reloading increased
quickly from 0.45%/hr to 0.7%/hr. From a vertical effective stress of 200 to about
1650 kPa, the actual strain rate increased relatively slowly from 0.7 %/hr to

188
0.8%/hr, followed by a faster increase to 0.85%/hr in the normally consolidated
region. Once the soil specimen was reloaded into the normally consolidated
region, the excess pore pressures began to increase more quickly. In the
normally consolidated region, the excess pore pressures during reloading
followed the trend extrapolated from the first loading, as indicated with a dotted
line in Figure 5.9.
The pore pressure ratio during reloading is plotted in Figure 5.11. At the
beginning of reloading, the pore pressure ratio increases very rapidly to 0.55 due
to the higher actual strain rate, its rapid increase, and also due to fast generation
of excess pore pressures. Following the sharp increase, the pore pressure ratio
decreases continuously to 0.25 until reaching the normally consolidated region.
This is related to the high stiffness of the soil in the over consolidated region,
where the rate of total stress (or vertical load) increase is higher than the rate of
excess pore pressure generation. The pore pressure ratio increased again in the
normally consolidated region as excess pore pressures began to build up more
rapidly as shown in Figure 5.9.

5.2.2. Constrained Modulus


The constrained modulus is defined as the inverse of the coefficient of volume
compressibility (mv). In this experimental program, the constrained modulus was
calculated as the tangent modulus using vertical effective stress and axial strain
at two adjacent data points i-1 and i. Thus, the equation for constrained modulus
can be expressed as:

Di =

1 'v ( 'v )i ( 'v )i 1


=
=
mv
a
( a )i ( a )i 1

(5.2)

Figure 5.12 presents the constrained modulus (D) as a function of the vertical
effective stress during the loading, unloading and reloading stages. The

189
constrained modulus shows a large scatter in the over consolidated region during
loading, with maximum variation of three orders of magnitude. Figure 5.13 shows
the stress-strain curve in the vertical effective stress range of 0 40 kPa. As a
result of the lack of smoothness in the stress strain curve in the OC region, the
constrained modulus calculated between any two successive points
(corresponding to readings every five minutes) shows substantial scatter.
To reduce the scatter, the values of the constrained modulus during loading were
processed using the moving average method with a window size of nine. The
moving average curve of the constrained modulus is shown in Figure 5.12 with a
solid line. The moving average curve better captures the degradation of the
constrained modulus in the over consolidated region, which indicates strain
softening behavior in this range. According to Janbu et al. (1981), the decrease
in the constrained modulus in the over consolidated region is related to
structural breakdown, and the constrained modulus reaches the minimum value
at the preconsolidation pressure. In CRS007, the constrained modulus reached
its minimum value at the vertical effective stress of 50 55 kPa. This value is in
general agreement with the preconsolidation pressure obtained using the strain
energy method and Casagrandes graphical construction. In the normally
consolidated region, the constrained modulus increased by one order of
magnitude per log cycle of vertical effective stress. The gradual increase in
constrained modulus with vertical effective stress in the normally consolidated
region is a reflection of the strain hardening behavior of reconstituted LR soil in
this stress range.
Upon unloading, the constrained modulus of reconstituted LR soil decreases
quickly. This is related to the concave upward shape of the unloading curve. At
the beginning of reloading, the constrained modulus decreased very sharply,
followed by a steady increase. However, as the specimen reaches the new

190
preconsolidation stress, the constrained modulus decreased again, similar to the
behavior during loading.

5.2.3. Hydraulic conductivity (k)


Figure 5.14 plots the hydraulic conductivity of reconstituted LR soil (CRS007)
versus void ratio during the loading and reloading stages of the test. The
hydraulic conductivity shown in this figure was calculated based on the measured
excess pore pressures employing non-linear CRS consolidation theory.
As the void ratio decreases from 4.0 to 2.5, the hydraulic conductivity of
reconstituted LR soil decreased from 9.36 x 10-8 cm/sec to 2.16 x 10-9 cm/sec,
yielding a value of Ck (=e/logk) of 0.972 ( or 1/1.029). At lower void ratios, for
e between 2.5 and 1.6, the rate of decrease in hydraulic conductivity increased,
with a Ck value of 0.641 (or 1/1.559). The average value of Ck is approximately
0.453 (or 1/2.21) (Force, 1998) for Resedimented Boston Blue Clay (RBBC) and
2.28 (or 1/0.439) for Middleton peat (Mesri et al, 1997).
Based on these data the Ck/eo ratio can also be computed. For the test
described here, Ck/eo decreases from 1/4.38 to 1/6.64 in the 4.0 2.5 and 2.5
1.6 void ratio ranges, respectively. These values are slightly lower than the
average value of 1/4 reported for peats (Mesri et al.1997). The average value of
Ck/eo for soft clays and silt deposits is close to 1/2. According to Mesri et al.
(1997), the low values of Ck/eo and high values of Ck for peats as compared to
clays and silts suggests that for peats only the macropores serve as flow
channels.
Figure 5.15 presents the change in the hydraulic conductivity with vertical
effective stress during loading and reloading. The hydraulic conductivity of LR
soil deceases by three orders of magnitude with an increase of two orders of
magnitude of effective stress. The inverse of the slope of the k- v curve in the

191
normally consolidated region corresponds to Ck/Cc, which indicates the change of
the logarithm of hydraulic conductivity for one log cycle change of vertical
effective stress:

e
Ck
log 'v
log k
=
=
Cc e
log k
log 'v

(5.3)

The value of Ck/Cc for reconstituted LR soil in the vertical effective stress range
of 40 550 kPa (void ratio of 4.0 2.5) is approximately 1/1.83, and decreases
slightly to 1/1.88 in the vertical effective stress range of 550 2000 kPa. These
values lie between the average values for clay and silt (near one) and peats (1/3
~ 1/2). According to Mesri et al. (1997), the low value of Ck/Cc compared to many
clays and silts indicates that only macropores serve as flow channels, whereas
the compression of both micropores and macropores contributes to the total
volume of flow.

5.2.4. Coefficient of Consolidation (Cv)


Figure 5.16 presents the variation of the coefficient of consolidation (Cv) with
vertical effective stress during the loading, unloading and reloading stages of a
CRS test. The coefficient of consolidation was calculated using the actual data of
the constrained modulus (D) and the hydraulic conductivity (k) obtained with nonlinear CRS consolidation theory. Thus:

Cv =

k
mv w

kD

(5.4)

where w = unit weight of water (9.81 kN/m3)


The scatter in the coefficient of consolidation is related to the high variation in the
constrained modulus. The moving average values calculated with nine data
points are also shown in Figure 5.16.

192
The coefficient of consolidation decreases very rapidly in the over consolidated
region. This is due to the decrease in both constrained modulus (Figure 5.12)
and hydraulic conductivity (Figure 5.15). The coefficient of consolation of
reconstituted LR soil continues to decrease beyond the preconsolidation
pressure, albeit less markedly with increasing vertical effective stress. From
vertical effective stresses of 50 to 1000 kPa, the decrease in the coefficient of
consolidation is approximately one order of magnitude. This is in contrast to the
behavior observed for most inorganic clays (e.g. Terzaghi et al. 1996; Sheahan
et al, 1997; Force, 1998) where the coefficient of consolidation remains constant
or increases moderately in the normally consolidated region. However, fibrous
peats data show continuous decrease of the coefficient of consolidation with
effective stress (Terzaghi et al, 1996). The continuous decrease in the coefficient
of consolidation of reconstituted LR soil in the normally consolidated region is
due to the substantial decrease in the hydraulic conductivity. As indicated in
Figure 5.12 and Figure 5.15, the increase in constrained modulus with vertical
effective stress was one order of magnitude in the normally consolidated region
while the decrease in hydraulic conductivity was two orders of magnitude.
During unloading, the coefficient of consolidation decreased by two orders of
magnitude from 7.2 x 10-3 cm2/sec to 1.3 x 10-4 cm2/sec due to the continuous
decrease in constrained modulus with increasing OCR. The change of the
coefficient of consolidation during reloading was a bit different than the behavior
during loading. At an OCR of 4 12, there was substantial decrease in
coefficient of consolidation but remained constant from 4 1. As the soil
specimen was reloaded into the normally consolidated region, the coefficient of
consolidation decreased faster and eventually appeared to follow the trend
extrapolated from the previous virgin loading (loading stage).

193
5.2.5. Repeatability
Three CRS consolidation tests, CRS006, 007 and 008, were performed to
evaluate the repeatability of the results. All the specimens were prepared using
disturbed soil samples from the same bin (LR-B5), sampled in April of 2000.
Although the same curing condition was applied, the initial void ratios and the
water contents of these three specimens were slightly different (Table 5.1). This
variation may be related to the differences in loss on ignition (36-45%) between
the specimens. All three tests were performed employing the same strain rate of
0.5%/hr. The details of the testing schedule of each test are summarized in Table
5.2.
The compression curves from these three tests are compared in Figure 5.17. The
compression curves in the over consolidated region show slight variation due to
the differences in the initial void ratios. In the normally consolidated region,
however, the three compression curves fall in a very narrow band and show good
repeatability. The preconsolidation pressures varied from 57.0 kPa (CRS007) to
61.21 kPa (CRS006) and 67.29 kPa (CRS008), with the average value of 61.83
5.17 kPa.
The compression index also shows consistent results, with an increase in the
average value from 1.567 0.100 (2 4 p) to 1.659 0.025 (4 8 p), followed
by a decrease to 1.585 0.026 (8 16 p). The swelling index showed relatively
higher variation. The average swelling index showed gradual increase with
increasing OCR from 0.068 0.014 (1 2 OCR) to 0.116 0.015 (2 4 OCR).
The average reloading index increased with decreasing OCR from 0.305 0.034
(2 4 OCR) to 0.454 0.024 (1 2 OCR).
The generation of excess pore pressures showed good repeatability as shown in
Figure 5.18. The excess pore pressures of these three specimens were almost
identical at the same void ratio.

194
Figure 5.19 presents a comparison of the constrained modulus of the three tests.
The repeatability of the constrained modulus in the over consolidated region is
hard to evaluate due to the high scatter. However, the constrained modulus
curves of all three tests followed a unique curve in the normally consolidated
region.
The hydraulic conductivity also showed excellent agreement in the normally
consolidated region as shown in Figure 5.20. The average value of Ck/eo for the
three reconstituted specimens (all prepared from the April specimens) was
1
.
4.49 0.15
The coefficient of consolidation curves are compared in Figure 5.21. As a result
of the consistency in the constrained modulus and hydraulic conductivity data,
with the exception of the behavior at very small stresses (<10 kPa), the
coefficient of consolidation of the three specimens shows very good repeatability,
with a maximum standard deviation of approximately 15%.

5.3. Strain rate selection and rate effects


As mentioned above, selection of the appropriate strain rate is a critical aspect of
CRS testing because of the inherent strain rate dependent nature of soil behavior
and because this is the factor that ultimately determines testing productivity.
At high rates CRS tests have been shown to cause a shift of the compression
curve towards higher stresses, ultimately leading to overestimation of the
preconsolidation pressure. According to Leroueil et al. (1983), the
preconsolidation pressure increases by approximately 10% per log cycle
increase in strain rate. This effect, attributed to the soils structural viscosity
(Jamiolkowski et al. 1985) is especially significant in the case of sensitive clays
such as those found in Canada and Scandinavia.

195
Excessively low rates, on the other hand, are a concern particularly when testing
soils that exhibit significant tendency to creep. If the rate applied is not sufficiently
fast to override the deformation of the soil arising from creep, the test leads to an
underestimation of the preconsolidation pressure, and to an overestimation of the
compressibility. The magnitude of the excess pore pressures generated at the
base of the specimen depends on the applied strain rate. To determine the
coefficient of consolidation and the hydraulic conductivity, the strain should be
fast enough to generate sufficiently large excess pore pressures so that they can
be accurately measured with available devices: pore pressure transducer and
data acquisition system. However, the strain rate should be slow enough to
prevent generation of an excessive hydraulic gradient across the soil specimen
which is not representative of in situ conditions.
There are two criteria available for selection of appropriate strain rates for CRS
consolidation tests, and both criteria are based on the absolute value of the pore
pressure ratio (uh/v). Wissa et al (1971) suggested a strain rate that yields a
pore pressure ratio of 2 5%, while the ASTM standard (D 4186-89)
recommends a strain rate that causes a pore pressure ratio between 3 and 30%.
Instead of using a pore pressure ratio as the strain rate selection criteria, Force
(1996) recommended a strain rate that yields a hydraulic gradient between 20
and 215. Based on the hydraulic conductivity results obtained from CRS
consolidation tests performed on RBBC with a strain rate in the range of
0.07%/hr 12.71%/hr, Force (1996) observed that the hydraulic conductivity
obtained with faster strain rate deviated from that obtained with slower strain rate.
With a faster strain rate, the hydraulic gradient applied across the specimen was
higher than 250. In this case, the flow cannot be assumed as laminar and
consolidation theory cannot be applied.
In the case of a highly compressible soil such as LR soil, the strain rate selection
is further complicated by the fact that the specimen height changes significantly

196
with time. In the CRS consolidation test apparatus employed for this
experimental program, the load platen on which the CRS cell rests advances at a
constant displacement rate, which causes the soil specimen to strain at a
constant rate with respect to the initial specimen height. As a result, the actual
strain rate, which is calculated with respect to the current specimen height at any
given time, increases continuously as the height of the soil specimen decreases.
The increase in actual strain rate with decrease in specimen height was shown in
Figure 5.5.
To evaluate the effect of the strain rate on the 1-D consolidation behavior of LR
soil, additional CRS consolidation tests were performed on reconstituted
specimens prepared with soil sampled in both April and July of 2000.
Reconstituted specimens were used to eliminate other variables, such as soil
structure, which may influence the evaluation of strain rate effects. Given the
slight difference in index properties (see Chapter 3) between the April and the
July samples, the results for these two samples need to be discussed separately.
Using soil obtained from the April 2000 sample, tests were conducted using the
following two strain rates: 0.25%/hr (CRS016) and 1.0%/hr (CRS017). Additional
tests were conducted using soil from the July samples employing rates of
0.1%/hr (CRS044) and 1.0%/hr (CRS046). The data for these specimens are
summarized in Table 5.1, and the schedules of all tests are summarized in Table
5.2. Note that in tests CRS044 and CRS046, the unloading strain rates were
reduced to 0.1%/hr to prevent generation of substantial negative excess pore
pressures.

5.3.1. Strain rate effects on compression behavior


Figure 5.22 presents the compression curves of April and July specimens
obtained employing strain rates of 0.25 and 1.0%/hr (April specimens), and 0.1
and 1.0%/hr (July specimens). The compression curves of these specimens lie in
two distinctively different bands, with the compression curves of the July soil lying

197
below those for the April soil with the difference in void ratio at a given vertical
effective stress ranging between 0.25 and 0.5 lower than the April specimens at
the same vertical effective stress. Note that as shown in Table 5.1, the July soil is
characterized by lower LOI, liquid and plastic limits compared to the soil sampled
in April. The effects of organic content and index properties on the compression
behavior of LR soil will be further discussed in section 5.4.
The values of the preconsolidation pressure determined from all the compression
curves shown in Figure 5.22 are shown in Table 5.3. The average
preconsolidation pressures ranged from 59.48 kPa (0.25%/hr) to 61.83 5.17
kPa (0.5%/hr) and 64.2 kPa (1.0%/hr) for April specimens. The preconsolidation
pressure of July specimens changed from 44.26 kPa (0.1%/hr) to 49.72 kPa
(1.0%/hr). No clear trend was observed between the strain rate employed and
the preconsolidation pressure for April specimens. The preconsolidation pressure
of July specimens increased by 12% with one log cycle increase in strain rate.
This is in agreement with the observation by Leroueil et al (1983).
No significant effect of strain rate was observed on the compressibility of both the
April and July specimens. It is very apparent in Figure 5.22 that the compression
curve of CRS016 (April, = 0.25%/hr) is parallel to the compression curve of
CRS017 (April, = 1.0%/hr), and that the compression curves of CRS044 (July,
= 0.1%/hr) and CRS046 (July, = 1.0%/hr) are also parallel. As summarized in
Table 5.3, the compressibility of the slowest and fastest tests of both April and
July specimens are very close to each other.

5.3.2. Strain rate effect on excess pore pressure generation


As mentioned in section 5.2.1., the magnitude of the cumulative excess pore
pressures generated in a CRS consolidation test depends mainly on the strain
rate and the hydraulic conductivity of the soil. In general, the magnitude of the

198
excess pore pressure increases with an increase in the strain rate and a
decrease in the hydraulic conductivity. The excess pore pressures in the April
and July specimens are plotted versus axial strain in Figure 5.23 a) and b),
respectively. In the April specimens, the strain rate of 0.5%/hr generated twice
the excess pore pressures generated with 0.25%/hr at the same void ratio. The
strain rate of 1.0%/hr, however, generated approximately 4.5 times higher excess
pore pressures than the 0.25%/hr rate. In the July specimens, a 1.0%/hr strain
rate generated excess pore pressures one order of magnitude larger than with
0.1%/hr. The results indicate that there is a linear relationship between the
excess pore pressures and the applied strain rate.
The linear relationship between strain rate and excess pore pressures is well
illustrated in Figure 5.24 where the excess pore pressures normalized by the
actual strain rate are plotted against axial strain. For the April soil, the normalized
excess pore pressure curves are in good agreement for strain rates of 0.25 %/hr
(CRS016) and 0.5%/hr (CRS007), while the data for the 1.0%/hr test (CRS017)
are slightly higher than those for the slower tests. The linear relationship is also
apparent for the July specimens where the normalized excess pore pressures
with actual strain rates of 0.1%/hr (CRS044) and 1.0%/hr (CRS046) are almost
identical.
Figure 5.25 shows the pore pressure ratio (uh/v) curves during loading. As
discussed above, the pore pressure ratio is used as a criterion for selecting the
most appropriate strain rate. ASTM recommends to use strain rates that produce
pore pressure ratios smaller than 0.3. For the April soil, as shown in the figure,
two slower strain rates, 0.1%/hr and 0.25%/hr, yielded pore pressure ratios
smaller than 0.2, up to a vertical effective stress of 1000 kPa, and satisfied the
ASTM recommendation. The strain rate of 1.0%/hr yielded a pore pressure ratio
higher than 0.3 at a vertical effective stress of 60 kPa, and increased
continuously reaching its highest value of 0.5 at vertical effective stress of 1000

199
kPa. For the July soil, on the other hand, the pore pressure ratio was below 0.25
within the same effective stress range (v <1000 kPa) when the strain rate of
1.0%/hr was employed.

5.3.3. Strain rate effect on hydraulic conductivity and coefficient of


consolidation
The hydraulic conductivity data obtained from the five tests conducted on the
April and July soil discussed so far are plotted against void ratio in Figure 5.26.
As seen for the compression curves, the curves relative to the two specimens fall
in two distinctive bands with the hydraulic conductivity of the July specimens
being 4 5 times higher, at a given void ratio.
The hydraulic conductivity of the April specimens obtained with a strain rate of
1.0%/hr (CRS017) ( = 1.0%/hr) was consistently 25 30% lower than the values
obtained with the other two slower rates. This is related to the larger excess pore
pressures generated with the strain rate of 1.0%/hr than with the two slower rates
at the same void ratio (Figure 5.23 (a)). On the other hand, the hydraulic
conductivities obtained with the two slower strain rates are in good agreement,
with a variation smaller than 7%. In CRS consolidation test, if a soil exhibits strain
rate dependent behavior, the excess pore pressures are not normalized by the
employed strain rate. In this case, a faster strain rate will cause generation of
significantly larger excess pore pressures. This leads to underestimation of the
hydraulic conductivity. The low value of the hydraulic conductivity obtained with a
strain rate of 1.0%/hr indicates that the April LR soil exhibits strain rate
dependency in this range of the strain rate.
The hydraulic conductivity of the July soil specimens show smaller strain rate
dependency than the April specimens. With a one log cycle increase in strain
rate, the variation in the hydraulic conductivity was much smaller, approximately

200
10%, than the April soil. The variation in the hydraulic conductivity of July
specimens may be related to the sample variability rather than to strain rate
effects.
The coefficients of consolidation of April and July specimens are plotted against
vertical effective stress in Figure 5.27 and Figure 5.28, respectively. Note that
moving average values of the coefficient of consolidation were used in these
figures. The strain rate effect on the coefficient of consolidation is hard to
evaluate since both hydraulic conductivity and constrained modulus affect the
coefficient of consolidation. Force (1998) observed that the coefficient of
consolidation of faster tests tends to come to the straight line portion at a higher
vertical effective stress compared to the slower tests.
For the April specimens, the coefficient of consolidation of CRS007 ( = 0.5%/hr)
was the highest followed by CRS016 ( = 0.25%/hr) and CRS017 ( = 1.0%/hr) at
the same vertical effective stress, in the normally consolidated region. This trend
might be caused by the variations in the hydraulic conductivity and constrained
modulus of the specimens. The hydraulic conductivity and constrained modulus
of CRS007 were higher than the other two specimens at the same vertical
effective stress, thus the coefficient of consolidation of CRS007 was the highest.
Similarly, CRS017 has the lowest hydraulic conductivity and constrained
modulus, yielding the lowest coefficient of consolidation.
For the July specimens, the two coefficient of consolidation curves are almost
parallel, with values of CRS046 ( = 1.0%/hr) being higher than CRS044 ( =
0.1%/hr) by 10 20 % in the normally consolidated region. Note that the CRS046
had a 15% higher hydraulic conductivity and approximately 5% lower constrained
modulus than CRS044 at the same vertical effective stress. Because of the
difference of these two factors, CRS046 has higher coefficient of consolidation.

201
All the results presented in this section indicate that reconstituted LR soil
specimens exhibit negligible strain rate dependency in the strain rate range of
0.25 0.5%/hr for the April specimens, and 0.1 1.0 %/hr for the July specimens.
For the April specimens, the strain rate of 1.0%/hr yielded a pore pressure ratio
greater than 0.3 when the vertical effective stress was larger than 60 kPa and did
not satisfy the ASTM criteria. Based on these results, the strain rate of 0.5%/hr
was selected to be employed for CRS consolidation tests in this experimental
program.

5.4. 1-D consolidation behavior of intact LR soil


An intact block sample of LR soil was collected in April 2000. The natural water
content of this sample was approximately 289%. After being transported to the
laboratory, the block sample was carefully cut into three smaller pieces using a
wire saw, and each piece was covered with two layers of plastic wrap and
covered completely with wax to prevent any moisture loss. The covered samples
were stored in a humid room, where the temperature and relative humidity were
kept at 10 C and 90%, respectively.
For preparation of intact LR soil specimens, the plastic wrap and wax were
removed with a knife. The intact sample was visually inspected to select the best
portion of the sample to be used for trimming the specimen. The block sample
was separated into pieces by vegetation root canal and large void. One intact
piece of block sample of approximately 15 cm wide, 20cm long and 13cm tall
was selected for testing, and separated with a razor blade. The rest of the block
sample was immediately covered with two layers of plastic wrap and one layer of
wax, and stored in the humid room. The selected piece was trimmed into a
prismatic shape with a flat bottom surface. The bottom and side surfaces of the
piece were covered with plastic wrap, and the side surfaces were reinforced with
duct tape.

202
The sample was trimmed into the specimen ring from the top following the same
procedure employed for the reconstituted specimens. The sample was trimmed
with a razor blade ahead of the specimen ring, and the specimen ring was
advanced downward in small increments. The top and bottom surfaces of the
specimen were trimmed again with a razor blade.
Three CRS consolidation tests were performed to evaluate the 1-D consolidation
behavior of intact LR soil. The initial conditions (water content, void ratio and unit
weight) of these specimens are summarized in Table 5.4, alongside the
measured values of the LOI and the specific gravity. As indicated by the table,
the water content values measured on the three specimens are all smaller than
the water content measured on the same soil at the time of sampling (289%).
Given that the liquid and plastic limits of the intact sample were found to be equal
to 405.7 and 252.7%, respectively (section 3.2.5), at the initial water content, the
liquidity index of the intact specimen is approximately 0.10.
Among the three specimens, CRS053 had the lowest values of LOI (57.72%) and
initial water content (214.4%). The average LOI of the intact soil is 62.01 3.73%,
and is about 13% to 25% higher than the disturbed April and July specimens.
The specific gravity of each specimen was back calculated using a linear
relationship between LOI and Gs.
The testing schedules followed in the three tests are summarized in Table 5.5. A
constant strain rate of 0.5%/hr was employed during loading for all three tests,
and the strain rate was reduced to 0.05%/hr during the unloading stage in tests
CRS052 and CRS053. CRS061 involved no unloading/reloading stage. Instead
the soil specimen was consolidated continuously to a vertical effective stress of
6255.7 kPa employing a 5000 lbs load cell.

203
5.4.1. Compression behavior of intact LR soil
Three compression curves for intact LR soil are plotted in Figure 5.29.
Representative compression curves for reconstituted LR soil prepared from the
April (CRS007) and July specimens (CRS044) are also plotted for comparison.
Compared to the reconstituted LR specimens, the compression curves of intact
LR soil exist at a much higher void ratio at the same vertical effective stress. For
example, at the vertical effective stress of 100 kPa, the corresponding void ratios
of reconstituted specimens of July, April and the intact sample are approximately
3, 3.6, and in the 4.4 4.9 range, respectively.
Intact LR soil exhibits a well defined S-shaped compression curve, which
indicates the structured nature of the natural material. According to Tan (2003),
the cementation effect of humic acid causes aggregation of inorganic minerals,
and creates and preserves large voids in the soil matrix. The compression
curves of CRS052 and CRS061 are identical in the normally consolidated region.
CRS053 exhibits slightly steeper compression curve in the normally consolidated
region, but merges into the other two compression curves at a vertical effective
stress of 1000 kPa.
Values of the preconsolidation stress measured on the three specimens of intact
LR soil range from 106.38 kPa (CRS061) to 120.97 kPa (CRS052), with an
average value of 114.90 7.60 kPa (Table 5.6). Note that the intact sample was
obtained at a depth of 1.5 m. Depending on the location of the groundwater table,
the vertical effective stress at the sampling depth is estimated to vary between 15
and 30 kPa. This implies that LR soil in in-situ conditions is over consolidated
with an OCR value of 3.5 8.2. While desiccation, mechanical loading/unloading,
or aging all are known to cause over consolidation of soil, it remains unclear
which source is responsible for the preconsolidation stress measured in this case.

204
In the normally consolidated region, the slope of the compression curve of intact
LR soil increases sharply, but becomes flatter at vertical effective stresses higher
than 1000 kPa. The compression index of the intact specimen decreases
continuously with increasing vertical effective stress from an average value of
2.620 0.230 (2 4 p) to 2.369 0.158 (4 8 p), 1.959 0.167(8 16 p),
and 1.577 0.012(16 32 p). Among the three samples, CRS053 has the
lowest compression index (Table 5.6).
The compression curve of the intact LR soil is compared to a fibrous peat (i.e.
Middleton peat, Mesri et al., 1997) and an inorganic clay (i.e. Resedimented
Boston Blue Clay, Force, 1998) in Figure 5.30. Middleton peat is a fibrous peat
from Middleton, Wisconsin, with about 90 95 % organic content and 623
846 % natural water content (Mesri et al, 1997). RBBC has a LOI of
approximately 4.4% and a water content of 44.2% (Force, 1998). The
compression curve of the intact LR soil lies in between the curves of the two
other soils. The figure shows that the intact LR soil exhibits compression
behavior stiffer than inorganic clays but more compressible than fibrous peats.
The compression index of intact LR soil is compared with other clays and organic
soils as a function of natural water content in Figure 5.31. The average natural
water content = 238% and the average compression index at 2 4 p =2.620 are
used to plot the intact LR soil in the figure. The compression index of the intact
LR soil lies in the upper range among the soils compared in the figure, and falls
in the range for peaty organic silts and Canadian muskeg.
The compression index of intact LR soil is compared with the values measured
on the reconstituted specimens in Figure 5.32. The compression index of the
intact LR soil is equal to or up to 1.3 times higher than the values measured on
the specimens prepared with the April and July samples, respectively. Note that
the July sample has a LOI 12% lower than the April sample, and has a

205
compression index in the normally consolidated region consistently smaller than
the April soil. As shown in Figure 5.32, the compression index of the intact soil
decreases very sharply. In the normally consolidation region (e.g. in the vertical
effective stress range 16 32 p), the intact LR soil exhibits the highest
compressibility followed by the April and July soils. Given that at this stress level
the effects of the natural structure have been erased, the higher Cc measured on
the intact soil once again likely reflects the higher organic content of this material.
Due to the differences in index properties, a direct comparison of the
compression behavior of the intact and reconstituted specimens can not be made.
This is instead possible making use of the framework proposed by Burland
(1990), who normalized the compression curves of naturally sedimented clays
with respect to the intrinsic compression curves of reconstituted clays using the
void index (Iv). The intrinsic compression curve is usually obtained with soils
prepared at a liquidity index between 1.0 and 1.5 (Burland, 1990). The void index
is defined as:

Iv =

*
*
e e100
e e100
=
*
*
e100
Cc*
e1000

(5.5)

where e = void ratio of naturally sedimented soil


e*100 = void ratio of reconstituted soil at v = 100 kPa
and e*1000 = void ratio of reconstituted soil at v = 1000 kPa
When e = e*100, the void index becomes zero, and similarly, when e = e*1000, the
void index is equal to -1. Figure 5.33 shows the intrinsic compression curves in
(a) e-logv space and (b) void index (Iv)- logv space schematically. The
normalization can be applied to soils with wide range of liquid limit to produce a
unique line, termed the intrinsic compression line (ICL) (Burland, 1990), and
described by the following equation:
ICL: I v = 2.45 1.285 ( log 'v ) + 0.015 ( log 'v )

(5.6)

206
Burland (1990) observed that there is a correlation between the Atterberg limits
(liquid limits) and the intrinsic constants of compressibility e*100 and Cc*. The
correlation between eL (void ratio at liquid limit) and e*100 and Cc* are given as
follows:
e*100 =0.109+0.679 eL -0.89 eL2 +0.016 eL3

(5.7)

and
Cc*.=0.256 eL -0.04

(5.8)

Burland (1990) mentioned, however, that equations (5.7) and (5.8) are applicable
only for soils with Atterberg limits lying above the A-line in the plasticity chart, and
for values of eL within the range of 0.6 and 4.5 (equivalent to w = 25% - 160%).
For soils below the A-line, the above correlations do not fit well.
The in-situ void ratios of shallow marine deposits are normalized with void index
and plotted with in-situ vertical effective stress in Figure 5.34. As shown in the
figure, various types of shallow marine deposits are plotted in a well defined
continuous band. The regression line fitted to the data is called the
sedimentation compression line (SCL). The SCL represents average
compression curves of normally consolidated clays. In general, the SCL lies at
higher vertical effective stresses, approximately five times higher, than ICL at the
same void ratio. This implies that most naturally sedimented clays can support
five times higher vertical effective stress than reconstituted clays at the same
void ratio.
The normalized compression curves of intact and reconstituted April and July
specimens are plotted in Figure 5.35. The compression curves of intact LR soil
are normalized using intrinsic properties derived from the reconstituted April
specimens. It should be noted that in this research program, the reconstituted
soil specimens were prepared below the liquid limit and the water content of the

207
specimens decreased further during curing. The e*100 and e*1000 of reconstituted
April specimens are estimated as 3.60 and 2.04, respectively from Figure 5.29.
The sedimentation compression line (SCL) is plotted using the coordinates given
by Burland (1990), and provided in Table 5.7 for reference. As shown in Figure
5.35, at the preconsolidation pressure the void indexes of CRS052 and CRS061
(plotted with open circles) lie on the SCL, while for CRS053 the point
representing the void index at p lies just below the SCL. In the normally
consolidated region, all three compression curves for intact LR soil slowly
converge to the ICL. This process is known as destructuration (Leroueil et al,
1979).
The destructuration process of intact LR soil in the normally consolidated region
can be described using the model proposed by Liu and Carter (1999 and 2000)
According to Liu and Carter, the virgin compression behavior of a structured soil
can be expressed as

e = e* + e

(5.9)

where e = void ratio of structured soil in the normally consolidated region,


e* = corresponding void ratio of reconstituted soil at the same vertical
effective stress
and e = difference in void ratios of structured and reconstituted soil at the
same vertical effective stress
Figure 5.36 presents the idealized compression behavior of reconstituted and
structured soils and illustrates e* and e (Liu and Carter, 2002).
Liu and Cater (1999 and 2000) observed that the additional void ratio due to the
structured soil (e) is inversely proportional to the current mean effective stress,
or to the vertical effective stress in 1-D consolidation. Based on this observation,
equation (5.9) can be expressed as:

208

'
e = e + e = e + ei p for v p
'v
*

(5.10)

where ei = difference in void ratio at preconsolidation pressure (p)


and b = compression destructuring index (0 b < )
The compression destructuring index (b) represents the rate of reduction in the
additional void ratio (e) or destructuring process. Figure 5.37 illustrates the
influence of the destructuring index b on the destructuring behavior of a
structured soil during compression. When b = 0, the difference in void ratio at the
preconsolidation pressure (ei) is constant in the normally consolidated region. In
this case the compression curve is parallel to the compression curve of the
reconstituted soil, and thus no destructuring takes place during consolidation. On
the other hand, if b = , e becomes zero immediately after the preconsolidation
pressure, thus the destructuring process is almost instantaneous. After
examining the compression behavior of various types of clay, Liu and Carter
(2000) observed that the compression destructuring index (b) generally falls in a
range between 0 and 30. The value of destructuring index b is smaller than one
for stiff soils, and is equal to or greater than 1 for soft soils.
Figure 5.38 (a) (c) present the destructuring behaviors of stiff and soft
structured soils during compression. The virgin compression line of stiff
Vallericca clay (Figure 5.38 (a)) is parallel to the normal compression line for
reconstituted soil and indicates b=0 type destructuring behavior. The
destructuring behaviors of four stiff clays are shown in Figure 5.38 (b). The
compression curve of Pietrafitta, of which b = 0.6, converges to the normal
compression line for reconstituted soil at a vertical effective stress 10 times of the
preconsolidation pressure. Albeit slowly, soils with b > 0.5 clearly shows
destructuring during compression. The destructuring of an artificially bonded soft
clay (b 1), as shown in Figure 5.38 (c), occurs very rapidly with high
compression index. Once the destructuring is completed, the compression

209
curves follow the normal compression line for reconstituted soil (or intrinsic
compression line). For natural medium sensitivity soft clays b typically lies
between 0.3 and 1
The compression curve of CRS061 and the intrinsic compression line are
shown in Figure 5.39. Two destructuring curves, obtained employing equation
(6.8) are shown in the figure. The solid line represents the destructuring curve
obtained with a compression destructuring index (b) of 0.8, assuming that the
destructuring process starts at the preconsolidation pressure (p = 106.4 kPa).
This curve deviates from the actual compression line in the early portion of the
normally consolidated region. The dashed line represents a destructuring line
obtained with a compression destructuring index (b) of 0.8 and p = 200 kPa.
This curve fits the actual compression curve of CRS061 and represents better
the destructuring of intact LR soil. In both cases, the compression destructuring
index of 0.8 yielded the curves most similar to the actual compression curve,
indicating that intact LR soil has developed a structure similar to that of other soft
soils.
The constrained modulus curves of untreated LR soil are plotted against
vertical effective stress in Figure 5.40. CRS052 has slightly higher constrained
modulus in the over consolidated region, and clearly indicates strain softening
behavior. In the normally consolidated region, the constrained modulus of all
three specimens falls in a very narrow band, showing good repeatability.
Representative curves of constrained modulus versus vertical effective stress for
the reconstituted soil specimens are also plotted in the same figure. A direct
comparison between these curves can not be made since the constrained
modulus depends on the vertical effective stress level. Therefore, the same
constrained modulus data of intact and reconstituted specimens are plotted in
Figure 5.41 this time as a function of vertical effective stress normalized by the
preconsolidation pressure. In the over consolidated region, the reconstituted July

210
specimen has a higher constrained modulus than intact LR soil. This may be
related to the large voids in the intact specimens such as cracks, root canal and
so on in intact soil specimens. In the normally consolidated region, the intact LR
soil exhibits a much stiffer response than the reconstituted specimens with the
constrained modulus 2 3 times larger. This is thought to be a reflection of the
presence of soil structure in the intact LR soil specimens.

5.4.2. Hydraulic conductivity and coefficient of consolidation of intact LR soil


The excess pore pressures measured in the intact specimens are shown in
Figure 5.42 as a function of void ratio. In specimens CRS052 and CRS061, the
excess pore pressure in the early part of the over consolidated region was very
small, less than 0.1 kPa, and was not stable. The excess pore pressures
stabilized and began to consistently increase at void ratios of 5.0 and 4.5,
respectively. This is probably due to the closure of voids, cracks and root canals
of the specimen as mentioned in section 5.4.1. These voids serve as a preferred
channel through which the excess pore pressures dissipate easily. Thus, the
excess pore pressure could not be measured until the soil specimens were
consolidated enough to close all these channels. At void ratios of 3.5 and smaller,
the excess pore pressure curves of the two specimens merged together. In the
CRS053 specimen, a root canal approximately 0.2 mm in diameter going through
the entire length of the specimen was observed in the center of the specimen
prior to setup. Not surprisingly, as a result, no excess pore pressure could be
measured until the specimen was consolidated to a void ratio of 3.75 (axial strain
of 17%). Afterwards, the excess pore pressures increased very quickly and
reached 10 kPa at a void ratio of 3.25. From this point on the excess pore
pressure generation became similar to that observed in the other two specimens.
Curves of the hydraulic conductivity of intact LR soil against void ratio are plotted
in Figure 5.43. Note that only values of k at values of void ratio where the
measured excess pore pressure was higher than 0.1 kPa are shown in this figure.

211
The hydraulic conductivity of intact LR soil has a relatively larger variation, of
about 25 50%, between specimens than reconstituted specimens. The
hydraulic conductivity of the specimen CRS061 was the smallest and the k of the
specimen CRS053 was the highest. The average hydraulic conductivity of intact
LR soil is approximately the same as the reconstituted April specimen. However,
the reconstituted July specimen has about 4 5 times higher hydraulic
conductivity at the same void ratio. Note that the reconstituted July specimen has
the lowest LOI of 37%; 12% smaller than the April specimen (49%), and 25%
smaller than the intact sample (62%). The slope of the linear portion of the
hydraulic conductivity curve (Ck) ranges from 0.821 (CRS061) to 0.894 (CRS052
and CRS053). The Ck/eo of the intact LR soil is in the range of 1/6.69 (CRS061)
to 1/5.35 (CRS053), with an average value of 1/(5.92 0.69). The average Ck/eo
of most peats is 1/4, and 1/2 for soft clay and silt (Mesri et al, 1997). The Ck of
Resedimented Boston Blue Clay (RBBC) is approximately 0.45 (Force, 1998)
The average hydraulic conductivity values of intact and reconstituted LR soil are
compared to those of Middleton peat and RBBC in Figure 5.44. The hydraulic
conductivity of both Middleton peat and RBBC were obtained from CRS
consolidation tests. The figure indicates that there is a clear trend between the
organic content (LOI) and the hydraulic conductivity, although the hydraulic
conductivity of RBBC is available only in void ratio ranges of 0.5 1.2. With
increasing organic content, the hydraulic conductivity decreases at the same void
ratio. The rate of change in hydraulic conductivity, Ck, on the other hand,
increases with increasing organic content.
The coefficient of consolidation of intact LR soil is shown in Figure 5.45. Similar
to hydraulic conductivity, the coefficient of consolidation of intact LR soil were
calculated only if the excess pore pressure was higher than 0.1 kPa. The large
variations in coefficient of consolidation in the over consolidated region, and also
in the normally consolidated region, up to a vertical effective stress of 300 kPa,

212
were due to the structural discontinuities present in the specimens. The
specimen CRS053, which had the highest hydraulic conductivity and lowest LOI,
had a coefficient of consolidation of 9 20% higher than the other two specimens,
for vertical effective stresses larger than 300 kPa.
The coefficient of consolidation of the intact LR soil and the other two
reconstituted specimens is compared in Figure 5.46, as a function of the vertical
effective stress normalized by the preconsolidation pressure. CRS061 was
selected to represent the coefficient of consolidation of intact LR soil, since it had
the best excess pore pressure generation among the three specimens. As shown
in the figure, the coefficient of consolidation of intact LR soil was more than 1.5
orders of magnitude higher than the reconstituted specimens at the
preconsolidation pressure, but continued to decrease significantly in the normally
consolidated region. This is related to the decrease in hydraulic conductivity.
Note that the constrained modulus of the intact soil increased throughout the
normally consolidated region, as shown in Figure 5.41. The coefficient of
consolidation of intact LR soil decreased much faster than the reconstituted
specimens in the normally consolidated region.
The coefficient of consolidation of the intact LR soil is compared to that of the
RBBC in Figure 5.47. The coefficient of consolidation of RBBC is characterized
by a sharp decrease in the over consolidation region followed by a steady
increase with vertical effective stress in the normally consolidation region. The
coefficient of consolidation of most inorganic clays exhibits the same behavior as
RBBC. The coefficient of consolidation of RBBC lies in the range of 2 x 10-3 2 x
10-2 cm2/sec at v < 1400 kPa. On the other hand, the coefficient of
consolidation of the intact LR soil continuously decreases in the normally
consolidated region, followed by a sharp decrease in the over consolidated
region. The coefficient of consolidation of the intact LR soil is consistently lower
than the value for RBBC at vertical effective stress greater than 170 kPa. In

213
addition, the difference of the coefficient of consolidation of the two soils
increases with increasing vertical effective stress.

5.5. Creep behavior of reconstituted LR soil


One of the most serious engineering problems associated with highly organic
soils and peats is their tendency to exhibit significant secondary compression
(creep). In geotechnical engineering, creep is defined as the deformation under
constant effective stress. In general, the creep behavior of soils is investigated by
performing incremental loading (IL) consolidation tests. In these tests, the load is
applied for 24 hours (or more) and the change of the height of the specimen is
measured with time. The load is then increased to a higher stress level using a
load increment ratio of one. At select increments (typically the maximum vertical
stress) the soil is allowed to creep for an additional 24 hrs. From each load
increment data, the time settlement curve is obtained, and the secondary
compression index (C = e/logt) is obtained as the slope of the time settlement
curve after the end-of-primary (EOP). Identification of the end-of-primary typically
relies on Casagrandes procedure.
In this research, a slightly different testing approach was employed. The IL
consolidation tests were performed by applying each load increment soon after
the EOP for the previous increment had been reached. This was possible
because the end of primary was determined based on the base excess pore
pressure measurements. According to Fox et al. (1992), the EOP can be
estimated as the point at which 95% of excess pore pressures have dissipated. It
thus becomes possibly to increase the load on the specimen without having to
wait for a significant amount of secondary to take place (necessary instead when
the identification of the end of primary conditions relies on Casagrandes method).
This was the procedure followed to perform some of the incremental tests
conducted in this research. Reflecting the fact that loading of the specimen

214
occurs shortly after the end of primary (EOP) is reached, these tests are herein
termed EOP IL consolidation tests. The only exception in this loading scheme
was represented by the load increment corresponding to the maximum vertical
stress at which the soil was allowed to creep for an extended period (a month or
more). In addition to the value of the secondary compression index (C) derived
from this creep phase, additional values of C could be derived for most of other
stress increments given that while the test is referred to as an EOP IL test, there
was some delay in the application of the load increment after reaching the EOP.
Albeit short, in most cases this delay was sufficient to allow determination of C.
The decision to perform IL tests was also motivated by the desire to compare the
results of incremental tests to those of CRS tests. With regard to characterizing
the creep behavior, IL test has a second advantage in that for each increment the
C/Cc ratio can be obtained directly from the compression index obtained at any
vertical effective stress level (in both the over consolidated and the normally
consolidated regions) and the secondary compression index obtained at the
corresponding vertical effective stress.
In this research, both conventional 24-hour IL and EOP IL consolidation tests
were performed on reconstituted LR soil. A conventional 24 hour IL test (IL025)
was performed on a specimen prepared with an April sample, and an EOP-IL
(IL037) test was performed on a specimen prepared with a July sample. The
index properties of the two specimens are summarized in Table 5.8. The applied
load and the results of the IL consolidation tests are summarized in Tables 5.9
and 5.10.
In IL025, the vertical effective stress was increased with a load increment ratio of
two except for the second load increment where it was increased by 10 times. In
IL037, the vertical effective stress was increased with the load increment ratio of
one except for the first load increment (LIR = 9). In addition, two load increments

215
were added near the preconsolidation pressure at v = 37.4 kPa and 74.9 kPa to
obtain more data points for better estimation of the preconsolidation pressure.
Note that in IL037, each load increment except the one corresponding to v =
1600 kPa was applied for between 1.2 (v = 800 kPa) and 2.3 (v = 100 kPa)
times the time required for reaching the end of primary (tEOP) to ensure that
dissipation of 95% of excess pore pressures were completed.
This section discusses the results of the IL consolidation tests and the evaluation
of C and the C/Cc ratio of untreated LR soil from the IL test results.

5.5.1. Determination of the End-Of-Primary points


The IL consolidation tests were performed with the same apparatus employed for
CRS consolidation tests. In both IL consolidation tests, the excess pore
pressures were measured with the cell and pore pressure transducers. However,
in IL025, the bottom of the specimen was not sealed completely, and the excess
pore pressures were dissipated through both the top and bottom of the
specimens. Therefore, the EOP points in IL025 were determined employing
Casagrandes procedure.
Figure 5.48 presents the time settlement curve obtained from IL025 at a vertical
effective stress of 69.9 kPa. The time settlement curve is characterized by two
linear portions, in primary and secondary consolidation stages, and a gradual
change of the curve near the EOP point. Two lines were drawn along the linear
portions and the EOP point was estimated as the intersection of the two lines
using Casagrandes procedure.
In the EOP-IL consolidation test (IL037), since the load was applied until the EOP
was reached, Casagrandes procedure could not be applied. Instead, the EOP
was estimated based on the excess pore pressures measurement. Figure 5.49

216
presents the time settlement and the dissipation of excess pore pressure from
IL037 at a vertical effective stress of 50.0 kPa. With an increase of the vertical
effective stress of 12.4 kPa, the maximum excess pore pressures generated
were 11.8 kPa. Given that the maximum excess pore pressure was 93.7 % of the
increase in the vertical effective stress, the measured maximum excess pore
pressure (rather than the total stress increment) was used to determine the
magnitude of 5% of the excess pore pressure and eventually to estimate the
EOP point. As summarized in Table 5.10, the maximum excess pore pressures
developed in each load increment of IL037 were consistently greater than 95% of
the actual load increments, except at the first load increment ( 75%) and at v =
50 kPa ( 93.7%). However, since the excess pore pressures were not the same
as the actual load increments, estimation of EOP based on the excess pore
pressure measurement may have caused some experimental error. This will be
discussed in section 5.5.2.
In IL037 at v > 56.7 kPa, the time required for the generation of maximum
excess pore pressure increases gradually from 0.23 minutes at v = 74.9 kPa to
2.4 minutes at v = 1600 kPa. These results are consistent with the observation
made by Mesri et al. (1997) from consolidation tests performed on Middleton
peat with approximately 90 - 95% of organic content. The delay of the generation
of maximum pore pressure may be related to the significant decrease in the
hydraulic conductivity with increase in effective stress of highly organic soils and
peats.

5.5.2. Comparison of the results from IL and CRS consolidation tests


The EOP compression curves of reconstituted LR soil obtained from IL
consolidation tests are compared with the curves from CRS consolidation tests in
Figure 5.50. The results from EOP-IL and CRS consolidation tests are in good
agreement and show good repeatability in the compressibility parameters. The

217
preconsolidation pressure of the specimen IL037 could not be determined since
only one data point was obtained in the over consolidated region. The
preconsolidation pressure of the specimen IL025 is 56.70 kPa, and is slightly
higher than the value obtained from the CRS tests, mainly due to the specimen
IL025 had a lower initial void ratio as shown in Figure 5.50.. However, the virgin
compression line is in good agreement with the CRS results, which indicates that
the higher preconsolidation pressure of IL025 is related to the lower initial void
ratio.
The hydraulic conductivity and the coefficient of consolidation obtained from the
IL consolidation tests are compared with the results from CRS consolidation tests
in Figure 5.51 and 5.54, respectively. Due to the dissipation of the excess pore
pressure from both the top and the bottom of the specimen in IL025, as
mentioned in section 5.5.1., the coefficient of consolidation of the specimen was
obtained assuming double drainage condition. The coefficient of consolidation
from the EOP-IL consolidation test (IL037) is approximately 10% smaller than the
values from the CRS consolidation tests at v >800.7 kPa. This variation is
related to the method used for determination of the end-of-primary (EOP) point.
To illustrate the experimental error of the results from EOP-IL consolidation test,
the EOP points obtained with both Casagrandes procedure and the excess pore
pressure measurements at v = 1600 kPa = is shown in Figure 5.53. As
indicated with open circles in Figure 5.53, the EOP point obtained with
Casagrandes procedure corresponds to eEOP = 1.503 and tEOP = 1104.4 minutes,
where the EOP from the excess pore pressures data corresponds to eEOP =
1.481 and tEOP = 1529.7 minutes. The variations in eEOP and tEOP between these
two methods would change the values of the coefficient of consolidation and
hydraulic conductivity as follows
Cv Casagrande = 4.13 E-05 (cm2/sec) and k Casagrande = 4.22 E-10 (cm/sec)
Cv excess pore pressure = 3.63 E-05 (cm2/sec) and k excess pore pressure = 3.71 E-10
(cm/sec)

218

At this vertical effective stress, the values of the coefficient of consolidation and
hydraulic conductivity calculated based on Casagrandes procedure are about
12% greater than the values from excess pore pressures measurements. This
may be one of the reasons why the coefficient of consolidation from the EOP-IL
consolidation test was smaller than the values obtained from the CRS tests. Yet,
the measurements of excess pore pressure provide a reliable method for
determining the EOP without allowing significant secondary compression of the
soil.
Except the small variation in the coefficient of consolidation, the results from IL
and CRS consolidation tests are in good agreements. The comparison of the
results from the IL and CRS consolidation tests indicates that the strain rate
effect is negligible for LR soil within the strain rate range up to 0.5%/hr for April
soil and up to 1.0%/hr for July soil.

5.5.3. Evaluation of C and C/Cc of reconstituted LR soil


The secondary compression behavior of reconstituted LR soil was studied from
the time-settlement curves of each of the load increments in IL025 (April soil) and
from the creep test performed at v = 1600 kPa on IL037 (July soil).
The secondary compression index of reconstituted April LR soil was obtained at
each load increment as the slope of the time settlement curve, as shown in
Figure 5.48. The C values are summarized in Table 5.9. The C of reconstituted
April LR soil changes with the vertical effective stress level between 0.129 (v =
1567.7 kPa) and 0.181 (v = 368.6 kPa).
The time settlement and the dissipation of the excess pore pressure with time
obtained from the creep test IL037 are shown in Figure 5.53. The duration of the
creep test was 50065.2 minutes (~34.8 days). The EOP was reached after

219
1529.7 minutes (based on the excess pore pressure measurement). During the
creep test, the specimen was allowed to creep for 1.5 cycles of secondary
compression. The time settlement curve of LR soil exhibits the characteristics of
the time settlement curve derived from Terzaghis consolidation theory: a linear
portion in the primary consolidation stage where excess pore pressures dissipate
and a second linear portion in the secondary compression stage where all the
excess pore pressures have dissipated.
The secondary compression index (C) of reconstituted July LR soil calculated
from t = 1529.7 to t = 50045.5 minutes is 0.123. As indicated by the line of time
settlement curve during secondary compression in Figure 5.53, the secondary
compression index remains constant during the time range investigated.
Although the secondary compression index is used to describe the creep
behavior of soils, it must be emphasized that its magnitude, as that of many other
soil parameters (e.g. Cc, Cv and k), changes depending on the effective stress
level, as well as on the over consolidation ratio. This is well illustrated in Figure
5.54, where the time settlement curves for three different stress increments (v =
50.0 kPa, 200.5 kPa, and 1600 kPa) from IL037 are plotted together. As
indicated in this figure, the secondary compression index of reconstituted LR soil
changes from 0.056 (v = 50.0 kPa) to 0.145 (v = 200.5 kPa) and to 0.123 (v
= 1600.0 kPa).
A parameter used for describing the susceptibility of a soil to creep is the C/Cc
ratio. According to Mesri et al. (1997), the change of C with time is directly
related to the change of Cc with vertical effective stress, and thus as Cc remains
constant, decreases, or increases with vertical effective stress, C also remains
constant, decreases, or increases with time, respectively. It should be noted that
the C/Cc concept can be applied in both the over consolidated and the normally
consolidated regions, so Cc represents the slope of the compression curve in

220

both regions. Using the C/Cc concept, the change of C of reconstituted LR soil
with vertical effective stress, illustrated in Figure 5.54, can be explained in the
following manner: C increases from v = 50.0 kPa to v = 200.5 kPa because
the compression index increased as the soil specimen is consolidated from the
over consolidated region to the normally consolidated region. The decrease of C
from v = 200.5 kPa to v = 1600.0 kPa is related to the fact that the
compression curve of reconstituted LR soil is slightly S-shaped so that the
compression index decreases with vertical effective stress.
The C/Cc ratio should be obtained from the secondary compression index (C)
and the compression index (Cc) associated with C. In IL025, as mentioned
before, the specimen was allowed to creep at each of the load increments.
Figure 5.55 presents the EOP points as well as the 24h points, which represent
the void ratio after 24 hour of loading. Figure 5.56 shows (a) the EOP and 24
hour void ratios and (b) time settlement curve at v = 69.9 kPa. When the vertical
effective stress was increased from 6.5 kPa (point A) to 69.9 kPa, the specimen
consolidated to point B (EOP point) during the primary consolidation stage as the
excess pore pressures dissipated. Since the load was applied for 24 hours, the
specimen continued to consolidate to point C (24 hour point) under the same
vertical effective stress during secondary consolidation. According to Mesri and
Castro (1986), a soil develops a preconsolidation pressure which is called the
quasi-preconsolidation pressure during secondary consolidation and thus it
becomes over consolidated. The magnitude of the OCR or the quasipreconsolidation pressure of soils developed during secondary consolidation can
be calculated with eq. (5.1.) This indicates that, as the specimen creeps from
point B to point C during secondary consolidation, the vertical effective stress of
the specimen will increase from point B to point D. The C/Cc concept (Mesri et
al., 1997) predicts that the reloading path of the specimen from point C will follow
the dark line indicated in Figure 5.56 (a); From C to D, and from D to E along the
virgin compression line. For calculation of C/Cc, the secondary compression

221
index obtained along the points B-C (in Figure 5.56 (b)) should be used with the
compression index obtained along the points B-D (in Figure 5.56 (a)).
Figure 5.57 and 5.60 present the plots of C and the corresponding Cc of the
reconstituted April and July LR soils, respectively. From the curves, the ratio of
C/Cc can be obtained as the slope of the linear best-fitting curve that goes
through the origin. The result shows that the values of C/Cc for reconstituted
April and July LR soils are 0.102 and 0.095, respectively. It should be also noted
that C/Cc for reconstituted July LR soil calculated with a single value of
secondary compression index obtained from the creep test (Figure 5.53) and the
corresponding compression index yields a value of 0.094, and well represents
the C/Cc of the soil.
The values of C/Cc for various geotechnical materials are summarized in Table
2.13. Compared to the average value of C/Cc for peats and muskegs (0.06
0.01), reconstituted LR soil has a much higher value of C/Cc. However, as
indicated in Table 2.14, the C/Cc for peat deposits reported in the literature
ranges from 0.035 to 0.10, and C/Cc for reconstituted LR soil falls on the upper
bound of this range.

5.5.4. Evaluation of C and C/Cc of intact LR soil


Two EOP-IL consolidation tests (IL05 and IL071) were performed on undisturbed
block LR soil specimens to evaluate the primary and secondary compression
behavior of LR soil. Similar to the Block LR soil specimens employed for the
CRS tests (Chapter 5.4), the two specimens used for the EOP-IL consolidation
tests had large voids. The same trimming approach applied for the CRS
specimens was used to reduce disturbance effects.

222
The same testing method applied for EOP-IL consolidation tests performed on
reconstituted specimens was employed for block specimens, i.e. the end of
primary was determined based on the excess pore pressures and the next load
increment was applied soon after reaching EOP instead of waiting for 24 hours.
The testing procedures (applied stress, time for each load increment, time for
EOP, etc.), measured excess pore pressures, and primary and secondary
compression indexes of each load increment of the two EOP-IL tests are
summarized in Tables 5.11 and 5.12. As summarized in Tables 5.11 and 5.12,
the excess pore pressure developed in the over consolidated region (v < 150
kPa) is very small, less than 20% of the increase in the applied stress, probably
due to the voids present within the specimens. Thus, the time for load increment
was increased in IL071 to obtain more data in the secondary consolidation stage
so that Casagrandes method for determination of EOP can be applied.
Figure 5.59 presents the EOP compression curves of block LR soil obtained from
two EOP-IL consolidation tests. The EOP compression curve of IL071 falls on
the compression curves of block LR soil obtained from CRS tests, whereas that
of IL056 lies below the other compression curves. This may be due to sample
variability. Nonetheless, the compression curves of block LR soil lie above the
compression curves of reconstituted LR soil, and shows destructuring behavior in
the normally consolidated region. The preconsolidation pressures of the
specimens for IL056 and IL071 were 132 kPa and 151 kPa, respectively.
Figure 5.60 presents the constrained modulus of the two block LR soil specimens.
As shown in the figure, the results are in good agreement with the results from
CRS consolidation tests and show good repeatability.
Figure 5.61 shows the hydraulic conductivity of the two block LR soil specimens
as a function of void ratio. The hydraulic conductivity of the specimen for IL 071
is in good agreement with the values from CRS tests, except one value at a void

223
ratio of 3.32 (vertical effective stress of 390.9 kPa). However, the hydraulic
conductivity of the specimen for IL056 is about one order of magnitude higher,
especially in the over consolidated region (e > 3.70). This is likely related to the
voids present in the specimens, and the that hydraulic conductivity values
measured are likely not to be representative of the behavior of the soil matrix
until the specimens are consolidated well into the normally consolidated region
and the voids are closed.
Figure 5.62 presents the coefficient of consolidation of the two block LR soil
specimens as a function of vertical effective stress. Even with the high hydraulic
conductivity of the two specimens, the results are in good agreement with the
results from CRS tests. The coefficients of consolidation of block LR soils are
significantly higher than that of reconstituted LR soil at vertical effective stress
smaller than about 500 kPa. This is mainly due to the high void ratio of block LR
soil: dissipation of pore water and thus consolidation process occurs much faster
in block LR soil due to higher void ratio.
In IL056 and IL071, one creep test was performed at the last load increment (at a
vertical effective stress of about 1570 kPa). The time-settlement curves and the
dissipation of excess pore pressures of the two block LR soils are plotted in
Figures 5.65 and 5.66, respectively. In IL056, the specimen reached the EOP in
1299.5 minutes, and the specimen was allowed to creep for 77749 minutes (1.8
log cycles of secondary compression). In IL071, the specimen reached the EOP
in 1586.4 minutes. The total time for the creep test was 20260.6 minutes, and
the specimen was allowed to creep for 1.1 log cycles of secondary compression.
As shown in the Figures 5.65 and 5.66, the slope of the time settlement in the
secondary compression range is linear, and does not change with time. The
secondary compression indexes, estimated from the slope of the time-settlement
curve in secondary compression stage, are 0.135 (IL056) and 0.158 (IL071).

224

The C/Cc of block LR soil was estimated based on the secondary compression
index obtained from the two EOP-IL consolidation tests and the compression
index obtained from the two EOP-IL and CRS consolidation tests. As mentioned
in Section 5.5.3, the secondary compression index obtained from a load
increment has to be coupled with the compression index obtained from the next
load increment (in EOP-IL test) or at the same vertical effective stress level at
which the secondary compression index was obtained (in CRS test)to estimate
the C/Cc. Since the long-term creep test was performed at the last load
increment of IL056 and IL071, the secondary compression indexes of IL056 and
IL071 were coupled with the compression index obtained from CRS061 at the
vertical effective stress of 1570 kPa. Figure 5.65 presents the average C/Cc of
block LR soil. The secondary compression indexes from the vertical effective
stresses of 95.5 kPa (step 5) and 200.3 kPa (step 6) of IL071 were not included
in this figure, since the EOP points at the vertical effective stress of 200.3 kPa
lies about 0.25 void ratio higher than the other compression curves (as shown in
Figure 5.59). Using the compression index of IL071 in 95.5 kPa 200.3 kPa
range will over estimate the C/Cc and will under estimate in 200.3 kPa 390.9
kPa range. As shown in Figure 5.65, the average C/Cc of block LR soil is about
0.095. This value is in good agreement with the C/Cc of reconstituted LR soil
(Section 5.5.3). Even though the secondary compression index of block LR soil
is higher than that of reconstituted LR soil, the high compression index of block
soil resulted in the C/Cc in the same range of reconstituted soil. The result
indicates that the C/Cc of soil or the relationship between the primary and the
secondary compression behaviors of soil may not directly related to the structure
of a soil. Instead, the C/Cc of soil may is related to the type or the composition
(both organic and inorganic components) of a soil. Compared to other soils, the
C/Cc of LR soil lies in the upper range of the average C/Cc of organic soils and
peats (Mesri et al., 1997).

225
Table 5.1: Summary of reconstituted LR soil specimen data
Test #

Bin #

Sampling Date

Curing

Curing time

wi

LOI

(kPa)

(days)

(%)

(%)

surcharge

ei

CRS006

LR-B5

April, 2000

48

14

179.16

45.34

4.033

CRS007

LR-B5

April, 2000

48

14

196.99

49.18

4.256

CRS008

LR-B5

April, 2000

48

14

219.47

36.37

4.257

CRS016

LR-B7

April, 2000

48

14

206.62

54.40

4.317

CRS017

LR-B7

April, 2000

48

14

214.75

55.16

4.450

CRS030

LR-B7

April, 2000

48

14

181.85

54.40

4.196

CRS044

LR-B9

July, 2000

48

14

156.27

36.57

3.800

CRS046

LR-B9

July, 2000

48

14

155.43

37.47

3.635

Table 5.2: Summary of testing process of CRS tests performed on reconstituted


LR soil
Strain
Test #

rate,

Unload rate 1st creep


factor

(%/hr)

stress

Creep time

2nd creep
stress

Creep time Final stress

(kPa)

(hrs)

(kPa)

(hrs)

(kPa)

CRS006

0.5

1.0

1595.3

48

65.4

72

2006.1

CRS007

0.5

1.0

998.9

72

85.7

72

2005.8

CRS008

0.5

1.0

194.4

48

10.9

48

2005.1

CRS016

0.25

1.0

995.0

48

90.2

48

2005.3

CRS017

1.0

1.0

995.9

72

60.3

96

2008.2

CRS030

1.0

1.0

1232.1

96

116.65

96

2007.3

CRS044

0.1

1.0

1001.5

72

100.4

72

1153.4

CRS046

1.0

0.1

1205.1

24

100.1

24

2011.5

226
Table 5.3: Summary of results of the CRS consolidation test performed on
reconstituted LR soil

(%/hr)

(kPa)

Test #

Cc
2-4p

4-8p

Cs

Cr

Ck/eo

8-16p

1-2OCR

2-4OCR

4-2OCR

1-2OCR

Ck/Cc

006

0.5

61.21 1.528 1.641 1.566

0.061

0.110

0.342

0.460

1/4.42 1/1.72

007

0.5

57.00 1.492 1.676 1.603

0.059

0.106

0.297

0.428

1/4.38 1/1.67

008

0.5

67.29 1.685

0.085

0.133

0.276

0.475

1/4.66 1/1.83

016

0.25

59.48 1.730 1.776 1.673

0.086

0.153

0.266

0.450

1/4.33 1/1.76

017

1.0

64.22 1.813 1.822 1.647

0.059

0.086

0.283

0.458

1/4.02 1/1.65

044

0.1

44.26 1.285 1.299 1.282

0.061

0.124

0.163

0.333

1/4.61 1/1.56

046

1.0

49.72 1.300 1.391 1.382

0.128

0.166

0.238

0.381

1/3.94 1/1.49

n/a

n/a

Table 5.4: Summary of intact LR soil specimen data


Gs

wi

LOI

(%)

(%)

CRS052

248.46

64.43

1.898

5.120

1.082

CRS053

214.39

57.72

1.969

4.780

1.095

CRS061

251.51

63.88

1.904

5.490

1.101

Average

238.12 20.61

62.01 3.73

1.924 0.039

5.130 0.355

1.093 0.010

Test #

ei

(g/cc)

Table 5.5: Summary of testing process of CRS tests performed on intact LR soil
Strain
Test #

rate,

Unload rate 1st creep


factor

(%/hr)

stress

Creep time

2nd creep
stress

Creep time Final stress

(kPa)

(hrs)

(kPa)

(hrs)

(kPa)

CRS052

0.5

0.1

998.8

72

141.8

CRS053

0.5

0.1

2016.1

24

202.0

72

2460.9

CRS061

0.5

6255.7

227
Table 5.6: Results of the CRS consolidation tests performed on intact LR soil
Test #

Cc

Cs

Cr

Ck/eo

Ck/Cc

(kPa)

2-4p

4-8p

8-16p

1-2OCR

2-4OCR

4-2OCR

1-2OCR

052

120.97

2.699

2.445

2.060

0.179

0.241

1/5.73

1/2.74

053

117.35

2.361

2.187

1.766

0.088

0.175

0.207

0.389

1/5.35

1/2.41

061

106.38

2.801

2.474

2.050

1/6.69

1/2.37

Averag 114.90 2.620

2.369

1.959

0.134

0.208

1/(5.92 1/(2.51

0.230 0.158 0.167

0.064

0.047

0.69)

7.60

0.20)

Table 5.7: Coordinates of sedimentation compression line (SCL) (Burland, 1990)

v (kPa)

Void index (Iv)

0.4

3.84

3.24

2.42

10

1.92

40

1.22

100

0.77

400

0.13

1000

-0.30

4000

-0.94

10000

-1.36

228
Table 5.8: Summary of index properties of reconstituted LR soil specimen data
for IL consolidation tests
Sampling

Test #
IL025
IL037

Date
April, 2000
July,2000

Test type
24h IL
EOP IL

wi

LOI

Gs

ei

(%)

(%)

194.44

52.54

2.02

4.38

178.52

1.99

3.67

45.97

*: The average LOI of the bin was used for specimen IL037

Table 5.9: Summary of data from 24h-IL consolidation test performed on


reconstituted April LR soil (IL025)
v
(kPa)

LIR

6.5

eEOP

tEOP

1/mv

Cv

(min)

(kPa)

(cm2/sec)

(cm/sec)

4.33

Cc

69.9

10.8

3.89

221.4

770.1

5.47e-4

6.97e-8

0.427

0.131

170.2

2.4

3.32

293.9

945.0

1.63e-4

1.70e-8

1.475

0.178

368.6

2.2

2.72

400.9

1785.1

1.00e-4

5.50e-9

1.788

0.181

768.7

2.1

2.20

440.9

4148.6

6.36e-5

1.50e-9

1.629

0.167

1567.7

2.0

1.73

661.6

9175.6

3.39e-5

3.62e-10

1.519

0.129

229
Table 5.10: Summary of data from EOP-IL consolidation test performed on
reconstituted July LR soil (IL037)
v
(kPa)

eEOP

tEOP

uhmax

uhmax/v

tuhmax

tf

(min)

(kPa)

(%)

(min)

(min)

tf/tEOP

Cc

2.71

3.67

25.4

3.57

41.3

17.1

75.4

0.18

63.2

1.53

0.106

0.037

37.4

3.48

111.4

11.4

95.0

0.22

200.5

1.80

0.514

0.047

50.0

3.40

178.4

11.8

93.7

0.30

343.6

1.93

0.673

0.056

74.9

3.24

338.2

24.4

98.0

0.23

510.4

1.51

0.902

0.093

100

3.10

429.7

24.8

98.8

0.23

1000.0

2.33

1.133

0.098

200.5

2.72

712.7

98.4

97.9

0.43

1001.0

1.40

1.253

0.145

399.6

2.30

909.2

191.8

96.3

0.75

1440.1

1.58

1.391

0.159

800.7

1.87

1204.4

382.8

95.4

1.4

1440.1

1.20

1.421

0.154

1600.0

1.48

1529.7

760.2

95.1

2.4

50065.2

32.7

1.297

0.123

Table 5.11: Summary of data from EOP-IL consolidation test performed on block
LR soil (IL056)
v
(kPa)

eEOP

tEOP

uhmax

uhmax/v

tuhmax

tf

(min)

(kPa)

(%)

(min)

(min)

tf/tEOP

Cc

0.92

4.26

6.15

4.26

0.32

0.3

5.7

0.8

5.0

15.6

12.1

4.25

0.07

0.8

13.4

0.03

6.0

85.7

0.027

0.001

25.0

4.23

0.22

1.5

11.6

0.03

5.0

22.7

0.072

0.004

49.9

4.16

0.37

3.1

12.4

0.03

6.0

16.2

0.220

0.014

99.6

4.04

0.55

6.7

13.5

0.02

9.5

17.3

0.393

146.5

3.92

0.53

6.3

13.4

0.08

6.0

11.3

0.745

0.034

200.0

3.70

6.5

9.5

17.8

0.05

10.1

1.6

1.654

396.4

2.94

416.0

104.3

53.1

0.2

451

1.1

2.552

787.7

2.29

893.6

363.4

92.9

0.95

915

1.0

2.177

1577.0

1.77

1299.5

672.3

85.2

4.0

77749

59.8

1.718

0.135

a : secondary compression index (C) not measured

0.002

230
Table 5.12: Summary of data from EOP-IL consolidation test performed on block
LR soil (IL071)
v

tEOP

uhmax

uhmax/v

tuhmax

tf

(min)

(kPa)

(%)

(min)

(min)

0.23

0.06

1.1

0.2

60.9

264.8

11.8

5.32
5.30

0.27

0.3

4.8

0.2

60.3

223.3

0.077

0.005

24.8

5.25

0.72

0.9

6.9

0.1

60.3

83.8

0.142

0.010

47.8

5.17

0.30

3.2

13.9

0.1

31.5

105.0

0.276

0.021

95.5

4.99

0.55

9.2

19.3

0.2

10.0

18.2

0.593

0.072

200.3

4.48

2.97

49.8

47.5

0.2

45.2

15.2

1.594

0.172

390.9

3.32

440.6

173.3

90.9

0.1

882

2.0

3.999

0.246

781.3

2.53

946.1

370.6

94.9

0.7

1900

2.0

2.634

0.237

1563.2

1.89

1586.4

732.9

93.7

1.0

20260.6

12.8

2.103

0.158

(kPa)
5.6

eEOP

tf/tEOP

Cc

0.005

231
a)
4.5
A

'p

CRS007[RC, 0.5%/hr]

4.0

Void ratio

3.5

3.0

2.5
E

2.0
D
1.5

10

100

'p
F

1000

Vertical effective stress, 'v (kPa)

b)
4.5
A

CRS007[RC, 0.5%/hr]

4.0

Void ratio

3.5

3.0

2.5
E

2.0
D
1.5

C
500

1000

1500

2000

Vertical effective stress, 'v (kPa)

Figure 5.1: Compression curves of reconstituted LR soil (CRS007) a) in log scale


and b) in arithmetic scale: A-B loading, B-C creep, C-D unloading, D-E creep,
and E-F reloading

232

20

Work per unit volume, W (kJ/m3 )

CRS007[RC, 0.5%/hr]
16

12

'p

Post-yield line

Pre-yield line
0

50

100

150

200

Vertical effective stress, 'v (kPa)

Figure 5.2: Estimation of preconsolidation pressure (p) using the strain energy
method

4.5
'p

CRS007[RC, 0.5%/hr]

4.0

Void ratio

3.5

3.0

2.5

2.0

1.5

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 5.3: Estimation of preconsolidation pressure (p) using Casagrandes


graphical construction

233

250
CRS007[RC,0.5%/hr]

Excess pore pressure, uh (kPa)

200

150

100

50

-50

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 5.4: Generation of excess pore pressure during loading

0.9

Strain rate (%/hr)

0.8

Platen advance rate


Actual strain rate

0.7

0.6

0.5

0.4

0.3
2.6

2.4

2.2

2.0

1.8

1.6

1.4

Sample height (cm)

Figure 5.5: Increase of actual strain rate with decrease in specimen height

234

0.20
CRS007[RC,0.5%/hr]

Pore pressure ratio, uh/v

0.18
0.16
0.14
0.12
0.10
0.08
0.06

10

100

1000

Vertical effective stress, 'v (kPa)

2.1

300

2.0

250

1.9

200

1.8

150

1.7

100

1.6

1.5
0.1

Time-settlement
Excess pore pressure

10

Excess pore pressure, uh (kPa)

Void ratio

Figure 5.6: Change of pore pressure ratio during loading

50

100

1000

0
10000

Elapsed time (min)

Figure 5.7: Time-settlement and dissipation of excess pore pressure during creep
after loading (CRS007).

235

Time-settlement
Excess pore pressure

Void ratio

2.1

-50

2.0

-100

1.9

-150

1.8

-200

1.7
0.1

10

100

Excess pore pressure, uh (kPa)

2.2

-250
10000

1000

Elapsed time (min)

Figure 5.8: Time-settlement and dissipation of excess pore pressure during creep
after unloading (CRS007).

Excess pore pressure, uh (kPa)

800

600

400

Reloading
200

Loading
0

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 5.9: Generation of excess pore pressure during reloading

236

1.0
0.9

Loading
Reloading

Strain rate (%/hr)

0.8
0.7
0.6
0.5
0.4
0.3

10

100

1000

Vertical effective stress (kPa)

Figure 5.10: Comparison of actual strain rate during loading and reloading.

0.6

Pore pressure ratio, uh/v

0.5
Loading
Reloading
0.4

0.3

0.2

0.1

0.0

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 5.11: Change in pore pressure ratio during reloading

237

Constrained modulus, D = 1/mv (kPa)

1e+6

Unloading

Actual data
Moving average

1e+5

Reloading
1e+4

Loading

1e+3

1e+2

1e+1

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 5.12: Change in constrained modulus (D) during loading, unloading and
reloading

0.0
CRS007[RC, 0.5%/hr]

0.5

Axial strain, (%)

1.0
1.5
2.0
2.5
3.0
3.5
4.0

10

20

30

40

Vertical effective stress, 'v (kPa)

Figure 5.13: Stress-strain curve in the over consolidated region

238

4.5
CRS007[RC, 0.5%/hr]
4.0

Void ratio

3.5

3.0

2.5

2.0

1.5
1e-10

1e-9

1e-8

1e-7

1e-6

1e-5

Hydraulic conductivity, k (cm/sec)

Figure 5.14: Change in hydraulic conductivity with void ratio

1e-5

Hydraulic conductivity, k (cm/sec)

CRS007[RC,0.5%/hr]
1e-6

Loading

1e-7

1e-8

Reloading
1e-9

Unloading
1e-10

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 5.15: Change in hydraulic conductivity with vertical effective stress

239

Coefficient of consolidation, Cv (cm2/sec)

1e-1
Actual data
Moving average
1e-2

Loading
1e-3

Unloading

Reloading

1e-4

1e-5

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 5.16: Change in coefficient of consolidation with vertical effective stress

4.5
CRS006[RC,0.5%/hr]
CRS007[RC,0.5%/hr]
CRS008[RC,0.5%/hr]

4.0

Void ratio

3.5
3.0
2.5
2.0
1.5
1.0

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 5.17: Comparison of compression curves

240

Excess pore pressure, uh (kPa)

500
CRS006[RC,0.5%/hr]
CRS007[RC,0.5%/hr]
CRS008[RC,0.5%/hr]

400

300

200

100

Void ratio

Figure 5.18: Comparison of excess pore pressure generation

Constrained modulus, D=1/mv (kPa)

1e+5
CRS006[RC,0.5%/hr]
CRS007[RC,0.5%/hr]
CRS008[RC,0.5%/hr]
1e+4

1e+3

1e+2

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 5.19: Comparison of constrained modulus

241

4.5
4.0

Void ratio

3.5
3.0
2.5
2.0

CRS006[RC,0.5%/hr]
CRS007[RC,0.5%/hr]
CRS008[RC,0.5%/hr]

1.5
1.0
1e-10

1e-9

1e-8

1e-7

1e-6

Hydraulic conductivity, k (cm/sec)

Figure 5.20: Comparison of hydraulic conductivity

Coefficient of consolidation, cv (cm2/sec)

1e-1
CRS006[RC,0.5%/hr]
CRS007[RC,0.5%/hr]
CRS008[RC,0.5%/hr]
1e-2

1e-3

1e-4

1e-5

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 5.21: Comparison of coefficient of consolidation

1e-5

242

4.5
CRS016
CRS007

CRS017

4.0
CRS044
CRS046

Void ratio

3.5

3.0

CRS007[April,0.5%/hr]
CRS016[April,0.25%/hr]
CRS017[April,1.0%/hr]
CRS046[July,1.0%/hr]
CRS044[July,0.1%/hr]

2.5

2.0

1.5

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 5.22: Effect of strain rate on compression curve

b)

Excess pore pressure, uh (kPa)

100

Excess pore pressure, uh (kPa)

a)
CRS017[April, 1.0%/hr]

100

CRS007[April, 0.5%/hr]
10

CRS016[April, 0.25%/hr]

10

15

20

Axial strain, (%)

25

30

CRS046[July, 1.0%/hr]

10

CRS044[July, 0.1%/hr]

0.1

10

20

30

40

Axial strain, (%)

Figure 5.23: Effect of strain rate on generation of excess pore pressure for a)
April and b) July specimens

Excess pore pressure/actual strain rate, uh/ (kPa/%/hr)

243

200
CRS007[April,0.5%/hr]
CRS016[April,0.25%/hr]
CRS017[April,1.0%/hr]
CRS046[July,1.0%/hr]
CRS044[July,0.1%/hr]

150

CRS007
CRS046

CRS016
100
CRS017
CRS044
50

10

20

30

40

50

Axial strain, (%)

Figure 5.24: Excess pore pressure normalized by actual strain rate

0.6
CRS007[April,0.5%/hr]
CRS016[April,0.25%/hr]
CRS017[April,1.0%/hr]
CRS046[July,1.0%/hr]
CRS044[July,0.1%/hr]

Pore pressure ratio, uh/v

0.5

CRS017

0.4

0.3
CRS046
0.2

CRS007
CRS016

0.1

CRS044
0.0

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 5.25: Effect of strain rate on pore pressure ratio

244

4.5
CRS016

CRS017
4.0
CRS007
CRS046

Void ratio

3.5

CRS044

3.0

2.5
CRS007[RC,0.5%/hr]
CRS016[RC,0.25%/hr]
CRS017[RC,1.0%/hr]
CRS046[RC,1.0%/hr]
CRS044[RC,0.1%/hr]

2.0

1.5
1e-10

1e-9

1e-8

1e-7

1e-6

1e-5

Hydraulic conductivity, k (cm/sec)

Coefficient of consolidation, Cv (cm2/sec)

Figure 5.26: Effect of strain rate on hydraulic conductivity

1e-2

CRS007[April,0.5%/hr]
CRS016[April,0.25%/hr]
CRS017[April,1.0%/hr]

1e-3

1e-4

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 5.27: Effect of strain rate on coefficient of consolidation of April soil

Coefficient of consolidation, Cv (cm2/sec)

245

1e-2

CRS044[July,0.1%/hr]
CRS046[July,1.0%/hr]

1e-3

1e-4

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 5.28: Effect of strain rate on coefficient of consolidation of July soil

6.0
5.5
5.0

Void ratio

4.5

[RC,April]

4.0
[RC, July]

3.5
3.0
2.5
2.0

CRS052[Block,0.5%/hr]
CRS053[Block,0.5%/hr]
CRS061[Block,0.5%/hr]

1.5
1.0

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 5.29: Compression curves of intact LR soil

10000

246

14
CRS052[Block,0.5%/hr]
CRS053[Block,0.5%/hr]
CRS061[Block,0.5%/hr]

12

Void ratio

10
8

Middleton peat
6
4
2
0

RBBC

10

100

1000

10000

Vertical effective stress, 'v (kPa)

Figure 5.30: Comparison of the compression curve of intact LR soil with a peat
(Middleton peat, Mesri et al, 1997) and an inorganic clay (RBBC, Force, 1998)

247

Compression index, Cc

10

Natrual LR soil

Clay and Silt Deposits


Peats

0.1
10

100

1000

Initial water content, Wo (%)


Figure 5.31: Comparison of the compression index of intact LR soil with other
soils (after Mesri and Rokhsar, 1974)

248

3.0
2.8

Undisturbed sample
RC, April sample
RC, July sample

Compression index, Cc

2.6
2.4
2.2
2.0
1.8
1.6
1.4
1.2
1.0

1-2

2-4

4-8

8-16

16-32

32-

Normalized vertical effective stress, 'v/'p

Figure 5.32: Comparison of the compression index

Iv =

*
*
e e100
e e100
=
*
*
e100
e1000
Cc*

Intrinsic compression
Line (ICL)

Void index, Iv

Void ratio, e

Intrinsic compression
Line (ICL)

e*100
e*1000

log v' (kPa)


(a)

0
-1

log v' (kPa)


(b)

Figure 5.33: The intrinsic compression line in (a) e-logv and (b) Void index (Iv)logv (after Burland, 1990)

249

Void index, Iv

Sedimentation Compression
Line (ICL)
2

Intrinsic Compression
Line (ICL)
-1

-2
0.1

10

100

1000

10000

log v' (kPa)

Figure 5.34: Intrinsic and sedimentation compression lines normalized with void
index (Burland, 1990)

250

1.5
Sedimentation compression line (SCL)
1.0
[RC, July]

0.5

Void index, Iv

[RC, April]
0.0
-0.5
-1.0
CRS052[Block, 0.5%/hr]
CRS053[Block, 0.5%/hr]
CRS061[Block, 0.5%/hr]

-1.5
-2.0

10

Intrinsic compression line (ICL)

100

1000

10000

Vertical effective stress, 'v (kPa)

Figure 5.35: Sedimentation compression curves of intact LR soil

Elastic behavior

Virgin yielding behavior

Void ratio, e

ei
Structured soil:
e = e*+e

e
e
e*
Reconstituted soil: e*

py,i

Mead effective stress, lnp


Figure 5.36: Idealization of the compression behavior of reconstituted and
structured soils (after Liu and Carter, 2002)

251

Void ratio, e

Structured soil

b=0
b=1
b=2
b=1000

b=0.1
b=0.2
b=0.5

Reconstituted soil
Mead effective stress, lnp
Figure 5.37: Influence of the destructuring index b during compression (afterLiu
and Carter, 2000)

252

Vertical effective stress, v, kPa


100
1000
10000
100000

10
(a)

Normal compression line


for reconstituted soil

Void index, Iv

Void ratio, e

1.75

1.65

-0.2

Sedimentation
compression line

b=0

b=0.1
-1.2

10000

-2.2

Normal compression line


for reconstituted soil
Todi b=0.2
Vallericca b=0.2
Pietrafitta b=0.6
Corinth marl b=0.1

b=0.6

b=0.2

(c)

1.1

Void ratio, e

100
Mean effective stress, p kPa

Sedimentation
compression line

-0.7

-1.7
1.55
10

(b)

A calcarenite
b=30

0.95

An artificially
bonded clay

0.8

Simulation with b=1


Simulation with b=3

0.65
10

100
1000
Mean effective stress, p kPa

10000

Fig. No.

Soil

ei

Preconsolidation
pressure (kPa)

Reference

(a)

Stiff Vallericca clay

0.111

1800

Amorosi &
Rampello (1998)

(b)

Four stiff clays:


Todi
Pietrafitta
Vallericca
Corinth marl

0.35Cc*
0.5 Cc*
0.46 Cc*
0.8 Cc*

13000
750
1600
2300

0.2
0.6
0.2
0.1

0.225
0.225
0.155

117
117
2400

3
1
30

Artificially bonded soft soil


(c)
A soft calcarenite soil

Burland et al. (1996)

Maccarini (1987)
Fontana el al. (1998)

Figure 5.38: Destructuring behavior of (a) stiff Vallericca clay and (b) four stiff
clays and (c) an artificially bonded soft clay and a soft calcerenite soil during
compression (after Liu and Carter, 2000)

253

6.0
CRS061[Block, 0.5%/hr]
Destructuring at 'v = 106.4 kPa

5.5
5.0

Destructuring at 'v = 200.0 kPa

4.5

Void ratio

4.0
3.5
Intrinsic Compression Line
(ICL)

3.0
2.5
2.0
1.5
1.0
0.5
10

100

1000

10000

Vertical effective stress, 'v (kPa)

Constrained modulus, D=1/mv (kPa)

Figure 5.39: Destructuring of intact LR soil

CRS052[Block,0.5%/hr]
CRS053[Block,0.5%/hr]
CRS061[Block,0.5%/hr]

1e+5

1e+4

[RC, July]

1e+3

[RC,April]

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 5.40: Constrained modulus of intact LR soil

10000

Constrained modulus, D=1/mv (kPa)

254

1e+5

1e+4

[RC, July]

1e+3

[RC,April]

0.01

0.1

10

100

Normalized vertical effective stress, 'v /'p

Figure 5.41: Constrained modulus against normalized vertical effective stress

6.0
CRS052[Block,0.5%/hr]
CRS053[Block,0.5%/hr]
CRS061[Block,0.5%/hr]

5.5
5.0
4.5

Void ratio

CRS061
4.0

CRS052

3.5

CRS053

3.0
2.5
2.0
1.5
1.0
0.001

0.01

0.1

10

100

1000

10000

Excess pore pressure, uh (kPa)

Figure 5.42: Generation of excess pore pressure in intact LR soil specimens

255

6.0
CRS052[Block,0.5%/hr]
CRS053[Block,0.5%/hr]
CRS061[Block,0.5%/hr]
Average[Block]

5.5
5.0

Void ratio

4.5

[RC,April]

4.0

[RC, July]

3.5
3.0
2.5
2.0
1.5
1.0
1e-11

1e-10

1e-9

1e-8

1e-7

1e-6

1e-5

1e-4

Hydraulic conductivity, k (cm/sec)

Figure 5.43: Hydraulic conductivity of intact LR soil

5.0
4.5
4.0

Middleton Peat

RC,April
RC,July
Block,April

Void ratio, e

3.5
3.0
2.5
2.0
1.5
1.0

RBBC

0.5
0.0
1e-11

1e-10

1e-9

1e-8

1e-7

1e-6

Hydraulic conductivity, k (cm/sec)

Figure 5.44: Comparison of the hydraulic conductivity of LR soil to that of RBBC


an Middleton peat

256

Coefficient of consolidation, cv (cm2/sec)

1e+0

CRS052[Block]

1e-1

CRS061[Block]

1e-2

1e-3

[RC,April]

[RC, July]

CRS053[Block]

1e-4

1e-5

10

100

1000

10000

Vertical effective stress, 'v (kPa)

Figure 5.45: Coefficient of consolidation of intact LR soil

Coefficient of consolidation, cv (cm2/sec)

1e+0

1e-1

1e-2

CRS061[Block]
[RC, July]

1e-3

[RC,April]

1e-4

1e-5

0.1

10

100

Normalized vertical effective stress, 'v /'p

Figure 5.46: Coefficient of consolidation against normalized vertical effective


stress

257

Coefficient of consolidation, cv (cm2/sec)

1e+0

RBBC
CRS061[Block,0.5%/hr]
1e-1

1e-2

1e-3

1e-4

1e-5

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 5.47: Comparison of the coefficients of consolidation of intact LR soil and


an inorganic clay

4.4

4.3

Void ratio

4.2

4.1

Primary
Consolidation

4.0
3.9
3.8

3.7
0.01

Reconstitued LR soil
(IL025)
'v = 69.9 kPa

0.1

eEOP

10

100

Secondary
Consolidation

1000

10000

Elapsed Time (min)

Figure 5.48: Determination of the EOP with Casagrandes procedure (IL025)

258

3.48

14
Time settlement
Excess Pore Pressure

12

3.44

10

3.42

3.40

EOP

3.38

3.36

Reconstituted LR soil
(IL037)
'v = 37.4 -> 50.0 kPa

3.34
3.32
0.01

0.1

EOPExcess pore pressure

10

Excess Pore Pressure (kPa)

Void ratio

3.46

-2
1000

100

Elapsed time (min)

Figure 5.49: Determination of the EOP based on the excess pore pressures
measurement (IL037)

4.5
4.0

Void ratio

3.5

3.0

2.5
2.0

CRS007 [April,0.5%/hr]
IL025 [April,24h-IL]
CRS044 [July,0.1%/hr]
IL037 [July,EOP-IL]

1.5

1.0

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 5.50: Compression curves of reconstituted LR soil obtained from CRS and
IL consolidation tests

259

4.5
CRS007 [April,0.5%/hr]
IL025 [April,24h-IL]
CRS044 [July,0.1%/hr]
IL037 [July,EOP-IL]

4.0

Void ratio

3.5
3.0
2.5
2.0
1.5
1.0
1e-10

1e-9

1e-8

1e-7

1e-6

Hydraulic conductivity, k (cm/sec)

Figure 5.51: Hydraulic conductivity of reconstituted LR soil obtained from CRS


and IL consolidation tests

CRS007 [April,0.5%/hr]
IL025 [April,24h-IL]
CRS044 [July,0.1%/hr]
IL037 [July,EOP-IL]

Coefficient of consolidation, cv (cm /sec)

1e-2

1e-3

1e-4

1e-5

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 5.52: Coefficient of consolidation of reconstituted LR soil obtained from


CRS and IL consolidation tests

260

1.9

800
Time settlement
Excess Pore Pressure

700

1.7

600

1.6

500
EOPCasagrande method

1.5

400

1.4

300

1.3

200

1.2
1.1
1.0
1e-2

Reconstituted LR soil
(IL037)
'v = 1600 kPa
1e-1

1e+0

100

Excess Pore Pressure (kPa)

Void ratio

1.8

EOPExcess pore pressure


0

1e+1

1e+2

1e+3

1e+4

-100
1e+5

Elapsed time (min)

Figure 5.53: Long term creep test for determination of secondary compression
index (C) of reconstituted LR soil (IL037)

261

4.0

Step3: 'v = 37.4


C = 0.056

3.5

Void Ratio

3.0

50.0 kPa

Step6: 'v =100.0


C = 0.145

200.5 kPa

Step9: 'v =800.7


C = 0.123

1600.0 kPa

2.5

2.0

1.5

1.0
1e-2

Reconstituted LR soil
EOP-IL (IL037)

1e-1

1e+0

1e+1

1e+2

1e+3

1e+4

1e+5

1e+6

Elapsed Time (min)

Figure 5.54: Comparison of the secondary compression index of reconstituted LR


soil at different vertical effective stress levels

262

4.5

4.0

Void ratio

3.5

3.0

2.5

2.0
CRS008 [April,0.5%/hr]
eEOP [April, IL025]
eFinal [April, IL025]

1.5

1.0

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 5.55: EOP consolidation curve from IL025

263
(a)
4.4

A
4.2

Void ratio

4.0

B
3.8

3.6

3.4

CRS008 [April,0.5%/hr]
eEOP [April, IL025]

3.2

10

100

Vertical effective stress, 'v (kPa)

(b)
4.4

A
4.3

Void ratio

4.2

4.1

Primary
Consolidation

4.0

3.9

3.8

3.7
0.01

Reconstitued LR soil
(IL025)
'v = 69.9 kPa

0.1

B = eEOP

Secondary
Consolidation
C

10

100

1000

10000

Elapsed Time (min)

Figure 5.56: Estimation of Cc associated with secondary consolidation for


determination of C/Cc: (a) compression curve and (b) time settlement curve at
v = 69.9 kPa (IL025)

264

0.20

Secondary compression index, C

0.18
0.16
0.14

Reconstituted April LR soil


(IL025)
C/Cc = 0.102

0.12
0.10
0.08
0.06
0.04
0.02
0.00
0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

2.0

Compression index, Cc

Figure 5.57: Calculation of C/Cc for reconstituted April LR soil (IL025)

0.18

Secondary compression index, C

0.16
0.14

Reconstituted July LR soil


(IL037)
C/Cc = 0.095

0.12
0.10

C/Cc
at 'v = 1600 kPa

0.08
0.06
0.04
0.02
0.00
0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

Compression index, Cc

Figure 5.58: Calculation of C/Cc for reconstituted July LR soil (IL037)

265

6.0
Preconsolidation pressure ('p)

5.5
5.0

Void ratio

4.5

[RC,April]

4.0
[RC, July]

3.5
3.0
2.5

CRS052[Block,0.5%/hr]
CRS053[Block,0.5%/hr]
CRS061[Block,0.5%/hr]
IL056[Block]
IL071[Block]

2.0
1.5
1.0

10

100

1000

10000

Vertical effective stress, 'v (kPa)

Constrained modulus, D=1/mv (kPa)

Figure 5.59: EOP compression curves of two block LR soil specimens

CRS052[Block,0.5%/hr]
CRS053[Block,0.5%/hr]
CRS061[Block,0.5%/hr]
IL056[Block]
IL071[Block]

1e+5

1e+4

[RC, July]

1e+3

[RC,April]

10

100

1000

10000

Vertical effective stress, 'v (kPa)

Figure 5.60: Constrained modulus of two block LR soil specimens

266

6.0
CRS052[Block,0.5%/hr]
CRS053[Block,0.5%/hr]
CRS061[Block,0.5%/hr]
Average[Block,0.5%/hr]
IL056[Block]
IL071[Block]

5.5
5.0

Void ratio

4.5

[RC,April]

4.0

[RC, July]

3.5
3.0
2.5
2.0
1.5
1.0
1e-11

1e-10

1e-9

1e-8

1e-7

1e-6

1e-5

1e-4

Hydraulic conductivity, k (cm/sec)

Figure 5.61: Hydraulic conductivity of two block LR soil specimens

Coefficient of consolidation, cv (cm2/sec)

1e+0

1e-1

CRS061[Block]

1e-2

1e-3

[RC, July]

CRS052[Block]

[RC,April]

CRS053[Block]
1e-4

1e-5

IL056[Block]
IL071[Block]

10

100

1000

10000

Vertical effective stress, 'v (kPa)

Figure 5.62: Coefficient of consolidation of two block LR soil specimens

267

2.4

700
Time settlement
Excess Pore Pressure

600

2.0

500

EOPCasagrande method

Void ratio

1.8

400

1.6

300

1.4

200

1.2
1.0
0.8
1e-2

100

Block LR soil
(IL056)
'v = 1577 kPa
1e-1

1e+0

Excess Pore Pressure (kPa)

2.2

EOPExcess pore pressure


0

1e+1

1e+2

1e+3

1e+4

-100
1e+5

Elapsed time (min)

Figure 5.63: Long term creep test for determination of secondary compression
index (C) of Block LR soil (IL056)

268

2.6

800
Time settlement
Excess Pore Pressure

700

2.2

600

Void ratio

2.0

500

EOPCasagrande method

1.8

400

1.6

300

1.4

200

Block LR soil
(IL071)
'v = 1563 kPa

1.2
1.0
0.8
1e-2

1e-1

1e+0

EOPExcess pore pressure

100

Excess Pore Pressure (kPa)

2.4

1e+1

1e+2

1e+3

1e+4

-100
1e+5

Elapsed time (min)

Figure 5.64: Long term creep test for determination of secondary compression
index (C) of Block LR soil (IL071)

Secondary compression index, C

0.30

0.25

Best-fit curve
IL056[Block]
IL071[Block]

0.20

0.15

0.10

Block LR soil (Average)


C/Cc = 0.095

0.05

0.00
0.0

0.5

1.0

1.5

2.0

Compression index, Cc

Figure 5.65: C/Cc of Block LR soil

2.5

3.0

269

CHAPTER 6. 1-D CONSOLIDATION BEHAVIOR OF PC TREATED LR SOIL

6.1. Introduction
The objectives of this chapter are (1) to present and discuss the results of CRS
and IL consolidation tests performed on PC treated LR soil, and (2) to present
the results of the effects of curing on the treatment with PC, and (3) to present
the results of the evaluation of the interaction of the organic matter present in LR
soil with cement.
Section 6.2 presents the results of the CRS and IL consolidation tests performed
on the PC treated LR soil. The effects of the treatment on the preconsolidation
pressure, compressibility, constrained modulus, the hydraulic conductivity, the
coefficient of consolidation, and the structure of the soil are discussed. The
results of creep tests performed to evaluate the effect of treatment on secondary
compression are also discussed.
Section 6.3 presents the results of consolidated tests performed to evaluate the
effects of curing on the 1-D consolidation behavior of PC treated LR soil. The
effects of curing time and curing surcharge are also discussed.
Section 6.4 presents a framework for the effects of treatment on the 1-D
consolidation behavior of chemically stabilized soils.
Section 6.5 presents the results of chemical tests performed to evaluate
reactions between the organic matter present LR soil and Portland cement. This
work is based on extraction and fractionation of humic substances from the PC
treated LR soil. The results of the elemental analyses performed to investigate
the change in the chemical composition of each fraction of the humic substances
of the PC treated LR soil are presented. The section also presents the results of
Fourier Transform Infra Red (FT-IR) spectroscopy analyses performed to

270
investigate the change and the reactivity of the organic functional groups with
cement.

6.2. 1-D consolidation behavior of PC treated LR soil


To evaluate the 1-D consolidation behavior of PC treated LR soil, CRS and IL
consolidation tests were performed on LR soil specimens treated with Type I
ordinary Portland cement (PC). The cement employed in this experimental
program was manufactured by Lone Star, Inc. Type I ordinary Portland cement
was selected as the binder since it is the most economic and the most widely
used cement in practice. More importantly, it was observed that Type I ordinary
Portland cement was the most effective binder for improving the unconfined
compressive strength of LR soil (Humphrey, 2001).
The amounts of PC used for treatment in this experimental program ranged from
about 8 to 20, 50, and 100 % by the dry mass of LR soil. Expressing the cement
content by PC percentage can be often misleading since the value may change
significantly for different types of soil depending on the water content and the
specific gravity of the soil or the method of binder introduction, such as wet or dry
methods. In practice, the amount of cement is expressed as the cement factor,
which is the ratio of the mass of cement (kg) used per unit volume of soil (m3).
The cement factor can be expressed using either the volume of treated or
untreated soil. The PC % investigated in this experimental program are
expressed in terms of cement factor in Table 6.1. The cement factor was
calculated assuming that the initial water content (wi) = 289 %, LOI = 50 %,
specific gravity of LR soil (Gs) = 2.05, specific gravity of hydrated cement = 2.15,
and that the PC was introduced as slurry (wet method) with a cement to water
ratio of 1:2. The range of cement factors employed in practice for the wet method
is about 200 400 kg/m3 (Bruce, 2000).

271
At least two samples were prepared for each cement content from the same
batch, and one CRS and one IL consolidation tests were performed to check the
repeatability of the results between these two types of tests and also address
strain rate effects. All the specimens were prepared following the same
procedure described in Chapter 4, with a curing surcharge of 48 kPa and curing
time of 14 days. The initial condition of each specimen is summarized in Table
6.2, including the sample bin number the LR soil was taken from, the amount of
PC (%PC by dry mass), the LOI (%), the initial water content following curing (w)
and initial void ratio (eo) of each specimen. All PC treated samples were prepared
with the disturbed LR soil samples obtained during the July 2000 field sampling.
The testing phases for CRS consolidation tests are summarized in Table 6.3.
Prior to testing in both CRS and IL consolidation tests, the PC treated LR soil
specimens were saturated by applying a back pressure of 560 600 kPa (80
85 psi) for 36 48 hours. All CRS consolidation tests were performed with a
strain rate of 0.5%/hr, the same strain rate used for undisturbed and
reconstituted LR soil as described in Chapter 5, to eliminate strain rate effects
from evaluation of the effects of treatment. The two creep stages, one after
loading and one after unloading, lasted for 72 hours, except in CRS055 where
the creep was not allowed after unloading. For 100% PC treated specimens
(CRS055 and CRS064), a 5000 lbs capacity load cell was incorporated to
consolidate the specimens up to approximately 6400 kPa. All CRS consolidation
tests performed on the PC treated specimens were performed following the same
procedure employed for the untreated specimens: first the specimen is loaded to
the first target stress (e.g. 1003.96 kPa for CRS043), then a creep stage follows
for 72 hours, and the specimen is unloaded to about 10% of the first target stress
(e.g. 106.3 kPa for CRS043), a second creep state follows for 72 hours, and
finally the specimen is loaded to the final stress (e.g. 1598.5 kPa for CRS043).

272
6.2.1. Calculation of the initial void ratio
Calculation of the initial void ratios of PC treated LR soil specimens presented
some challenges since the specific gravity of PC treated LR soil could not be
measured due to the complex nature of the soil-cement interactions involved.
According to Lea (1956), for example, the volume and mass of cement, and the
specific gravity continuously change with time upon hydration. The changes in
volume and mass depend on the degree of hydration, temperature, vapor
pressure and chemical composition of cement, each of which is very hard to
measure accurately. Lorenzo and Bergado (2004) measured the specific gravity
of cement treated Bangkok clay after 7, 14 and 28 days of curing. They observed
that the specific gravity of the treated soil generally decreased with increasing
cement content, but there was no relationship between the specific gravity and
water content or curing time.
In this experimental program, the void ratio of PC treated LR soil was calculated
by considering cement and LR soil as two different phases, and by calculating
the mass and volume of each phase separately. The four phases of PC treated
LR soil are schematically shown in Figure 6.1. The hydrated cement phase
represents all the products of cement hydration and the water combined in
hydrated cement compounds. The void ratio of PC treated LR soil was calculated
based on the following assumptions:
1) cement particles are fully hydrated after 14 days
2) all water evaporates at 110oC
3) the loss on ignition (LOI) of cement at 440 oC is negligible, and
anything that combusts at this temperature is part of the LR soil
4) the specific gravity of LR soil does not change due to reaction with
cement
These assumptions have shortcomings. Regarding the first assumption, it has
been reported that cement continues hydrating up to 5 years, but most of

273
hydration in general is considered to be completed within 28 days (Lea, 1956).
The water combined in the hydrated cement compounds, other than calcium
hydroxide (CaOH), evaporates below 350 C. The water combined in calcium
hydroxide evaporates completely between 350 and 550 C. Thus, the loss on
ignition of PC treated LR soil includes the loss of organic matter and the loss of
water in the hydrated cement.
During cement hydration, calcium and hydroxyl ions are released into solution,
which leads to an increase the pH of the solution during cement hydration. If the
pH of the solution increases above 12, silica (SiO2) and alumina (Al2O3) on the
surface of clay minerals are dissolved and react with calcium and hydroxyl ions.
In this case, the specific gravity of the LR soil may decrease.
The mass and the volume of the four phases shown in Figure 6.1 were
calculated as follows: the total mass (MT) and the total volume (VT) of each
specimen were measured after trimming was completed. The mass of the solids
in the specimen (Ms), including LR soil and cement, was measured after drying
the specimen at 110 C. The mass of water (Mw) was determined by subtracting
Ms from MT.
During sample preparation, two water contents were determined: after adding
water to raise the water content of the sample to 289% (w1), and after adding the
cement as slurry to the soil and mixing it (w2). Based on the first water content
measurement (w1), the actual amount of solid LR soil in the batch was
determined, and thus the percentage of cement by dry mass of LR soil was
measured. Assuming that cement and LR soil were mixed thoroughly so the soilcement mixture was homogeneous, the mass of solid LR soil (Mls) was calculated
as:

M ls =

Ms
1 + Cement content

(6.1)

274
Once the mass of the LR soil solid (Mls) was calculated, the mass of cement was
calculated by subtracting Mls from Ms.
The volume of each phase was calculated using the specific gravity of each
phase. Assuming that the cement was fully hydrated in 14 days of curing period
(first assumption), the specific gravity of hydrated cement was assumed as 2.15
(Lea, 1956). The specific gravity of the LR soil was calculated using the linear
relationship between the loss on ignition and the specific gravity, as described in
Chapter 3. The volume of air (Va) was calculated by subtracting the volume
occupied by water (Vw), LR soil (Vls) and hydrated cement (Vc) from the total
volume (VT).
Figure 6.2 shows the change in the initial void ratio with the amount of PC (% by
dry mass). The initial void ratio of PC treated LR soil decreased gradually with
increasing %PC from 8 to 50%, and then decreased sharply for 100 %PC. The
initial void ratio of LR soil treated with 100% PC was smaller by approximately 1
point than the void ratio of LR soil treated with 50% PC.
To check the validity of the third assumption, i.e. that the loss of ignition of
cement at 440 oC is negligible, the loss on ignition of the LR soil used in the PC
treated specimen was back calculated and compared to the actual value of the
loss on ignition of LR soil measured for that specimen. After the dry mass of the
specimen (Ms) was measured, the dried soil was burned in a furnace at 440 C
and the mass (M2) was measured again to determine the loss on ignition of the
specimen. The ratio of the change in the mass before and after burning the
specimen in a furnace (M = Ms M2) to the dry mass of the specimen (Ms)
represents the loss on ignition of the PC treated soil specimen. The loss on
ignition of each specimen is summarized in Table 6.2. The ratio of the mass
change (M) to the dry mass of the LR soil in the specimen (Mls) represents the
back-calculated loss on ignition of the LR soil used for the specimen. The actual

275
loss on ignition of LR soil was determined using the LR soil dried to determine
the water content, w1, as mentioned before. The back calculated loss on ignition
and the actual loss on ignition are compared in Table 6.4. As shown in this table,
the differences between these two values of the LOI are very small (0.13
2.34 %), and lie within range of experimental error. The small differences in these
two values indicate that (1) the third assumption is valid and (2) the reaction with
cement does not decrease the amount of organic matter present in the soil
specimens. The significance of this second point will be discussed further in
section 6.4.

6.2.2. 1-D Consolidation behavior of 8.0% PC treated LR soil


One CRS (CRS049) and one IL (IL048) consolidation tests were performed on
LR soil treated with 8.00% PC. In the CRS test, unloading and reloading was
applied to measure the swelling and recompression indexes. The results of the
consolidation test are summarized in Table 6.5 including the preconsolidation
pressure (p), the compressibility parameters (compression index (Cc), swelling
index (Cs), and recompression index (Cr)), Ck/eo and Ck/Cc. The testing phases
and the results of the IL consolidation test are summarized in Table 6.6 including
the vertical stress applied at each increment, void ratio at the end of primary
(eEOP), time required for reaching EOP (tEOP), maximum excess pore pressure
measured during the increment (uhmax), the ratio of the maximum excess pore
pressure to the vertical effective stress increase (uhmax/v), time taken for
generation of the maximum excess pore pressure (tuhmax), the total duration of
the load increment, the coefficient of consolidation (Cv), the hydraulic conductivity
(k), and the compression index (Cc).
The incremental loading consolidation test was performed following the same
procedure described in Section 5.5, by increasing the load as soon as the
specimen reached the end of primary consolidation (EOP-IL test). The end of

276
primary consolidation was determined by continuously monitoring the dissipation
of the excess pore pressures. EOP-IL test was selected as it allows one to more
accurately determine the effects of treatment than the conventional IL
consolidation tests. According to Kassim and Clarke (1999), the incremental
loading consolidation test is not suitable for investigating the compression
behavior of chemically stabilized soil because the stiffness of the soil increases
as the soil specimen creeps after reaching the end of primary. Increasing the
load as soon as primary consolidation is completed accelerates the testing time
and limits the increase in stiffness with time.
Figure 6.3 presents the two compression curves obtained from CRS049 and
IL048. The two compression curves are observed to be almost identical,
indicating that strain rate effects are not significant.
The most apparent effect of treatment on the consolidation behavior was the
increase in the preconsolidation pressure. The values of the preconsolidation
pressure obtained from IL048 and CRS049 are 70.21 and 79.43 kPa,
respectively. These values represent an increase of 50% over the average
preconsolidation pressure of reconstituted LR soil prepared with soil collected in
July 2001 (50.22 kPa, see chapter 5).
The compression curve of the 8% PC treated LR soil is shifted to higher effective
stress, indicating that the treated LR soil can sustain higher vertical effective
stress than the reconstituted LR soil at the same void ratio. The results indicate
that when treated with 8.00%PC LR soil can sustain 50% higher vertical effective
stress than the reconstituted soil at the same void ratio.
The compression index did not change significantly as a result of treatment. The
compression curves of 8%PC treated LR soil are almost parallel to the
compression curve of the reconstituted LR soil. The average compression index

277
of LR soil treated with 8% PC decreases with increasing vertical effective stress
from 1.444 (2 4 p) to 1.428 (4 8 p)and 1.280 (8 16 p) (Table 6.5).
Compared to the reconstituted soil, the compression index of 8%PC LR soil
increased by 12% and 6% in the ranges of 2 4 p and 4 8 p, respectively,
but decreased by 4% in the 8 16 p range. The slight increase in the
compression index at < 8 p may related to the destruction of the structure
developed by treatment. Lorenzo and Bergado (2004) observed similar results in
that the compression index of soft Bangkok clay increases with treatment with
cement. They noted that the increase in the compression index with treatment
was due to the abrupt breakup of the cementation bonds.
While the compression index does not show any significant change, the effects of
treatment on the swelling and reloading indexes are apparent. Compared to
reconstituted LR soil, the swelling index decreased by 57% and 55% to 0.041(1
2 OCR) and 0.066 (2 4 OCR), respectively. The reloading index also
decreased with treatment, more drastically than the swelling index, by 72% and
61% to 0.057 (4 2 OCR) and 0.139 (2 1 OCR) (Table 6.5).
Figure 6.4 presents the compression curves of LR soil treated with 8.00% PC
normalized using the void index (Iv) parameter. The normalization with the void
index allows one to compare the effects of treatment on different types of soil or
treatment with different types of binding agents. The void ratios of the
reconstituted July LR soil at the vertical effective stresses of 100kPa (e100 = 3.60)
and 1000 kPa (e1000 = 2.04) were used as the intrinsic parameters, e*100 and
e*1000, for normalization of the compression curve of 8.00% PC treated soil (see
section 5.4.1.). Using the intrinsic parameters, the void index of the PC treated
LR soil is calculated as:

I v PC treated =

*
ePC treated e100
*
100

*
1000

ePC treated 3.60


3.60 2.04

(6.2)

278
Note that the normalized compression curves of the block LR soil lie on the
sedimentation compression line (SCL) at the preconsolidation pressure, followed
by an abrupt increase of the slope of the curve in the normally consolidated
region. The figure shows that the compression curve of the LR soil treated with
8.00 % PC lies slightly above the intrinsic compression line. This indicates that
8.00% PC treated LR soil did not develop significant soil structure and did not
provide as much resistance to compression as the LR soil in in-situ state.
The values of the constrained modulus of LR soil treated with 8.00% PC are
plotted against vertical effective stress in Figure 6.5. The constrained modulus
obtained from the IL consolidation test was obtained as the slope of the
compression line in arithmetic scale, and thus is plotted at the average vertical
effective stress of two load increments. For example, the constrained modulus
obtained in the vertical effective stress range of 786.9 and 1572.7 kPa was
plotted at 1179.8 kPa. The constrained modulus of IL048 and CRS049 shows
large variation in the over consolidated region due to the variation in the initial
void ratio of the two specimens. In the normally consolidated region, however,
the constrained moduli from the two tests are consistent.
Compared to the reconstituted LR soil, the constrained modulus of LR soil
treated with 8.00% PC is the same as that of the reconstituted LR soil in the
normally consolidated region at the same vertical effective stress. This indicates
that the compression curves of the reconstituted LR soil and the LR soil treated
with 8.00% PC are parallel in the normally consolidated region.
The cumulative excess pore pressures generated in CRS049 are plotted in
Figure 6.6. To eliminate the effect of the increase in the strain rate with
decreasing specimen height, as described in section 5.3.2, the excess pore
pressures are normalized by the actual strain rate. As shown in the figure,
consistently smaller excess pore pressures were generated in LR soil treated

279
with 8% PC than in reconstituted LR soil at the same void ratio. The excess pore
pressures generated in each load increment in the EOP-IL consolidation test are
summarized in Table 6.6. In the over consolidation region, the excess pore
pressures generated were approximately 78 84% of the load increment. In the
normally consolidated region, the excess pore pressures were about 92 100%
of the load increment. The decrease in the generated excess pore pressure in
the LR soil treated with 8.00% PC is related to the increase in the hydraulic
conductivity.
Figure 6.7 presents the hydraulic conductivity of LR soil treated with 8.00% PC.
At the same void ratio, the hydraulic conductivity of IL048 is almost the same or
slightly increased than that of reconstituted LR soil. In CRS049, the hydraulic
conductivity increased by 18 26% than that of the reconstituted LR soil at the
same void ratio in the normally consolidated region. The rate of the change in the
hydraulic conductivity with void ratio (Ck) or vertical effective stress (Ck/Cc)
remained the same as the reconstituted LR soil. The average values of Ck/eo and
Ck/Cc of LR soil treated with 8.00%PC are 1/ (4.35 0.33) and 1/ (1.57 0.01),
respectively.
The values of the coefficient of consolidation of LR soil treated with 8.00% PC
are plotted against vertical effective stress in Figure 6.8. The coefficient of
consolidation of LR soil treated with 8% PC increased by 20% (IL048) or by 50
100% (CRS049) with respect to the reconstituted LR soil at the same effective
stress. Since the constrained modulus did not change significantly with treatment,
the increase in the coefficient of consolidation is mainly due to the increase in
hydraulic conductivity. The results indicate that the rate of consolidation of LR
soil will increase by 1.5 2 times when treated with 8.00% PC.
Figure 6.9 presents the settlement and excess pore pressures dissipation curves
obtained from a creep test performed on LR soil treated with 8.00% PC. The

280
vertical effective stress was 1572.7 kPa. The maximum excess pore pressures
developed was 719.6 kPa, approximately equal to 91.6% of the applied load
increment. As described in section 5.5.1, the end of primary (EOP) was
determined based on the excess pore pressure measurement. The time for EOP
was 605.1 minutes. The creep test lasted for 16.4 days, and the specimen was
allowed to go through 1.6 log cycles of secondary consolidation. As shown in the
figure, the slope of the time settlement curve in the secondary compression stage
remained constant, with the secondary compression index being 0.106. Based
on the compression index for the stress increment at the vertical effective stress
of 1573 kPa, the C/Cc ratio for 8.00% PC treated LR soil is 0.085. Note that the
secondary compression index of reconstituted LR soil obtained at the vertical
effective stress of 1600.0 kPa was 0.123, with the C/Cc ratio of 0.095. The
results indicate that the secondary compression index and the C/Cc ratio of LR
soil decreased by 14% and 10%, respectively, when treated with 8.00% PC.

6.2.3. 1-D consolidation behavior of 18.7% PC treated LR soil


The effects of treatment with 18.7% PC were evaluated by performing one CRS
(CRS043) and one IL (IL042) consolidation tests. The results of these tests are
summarized in Table 6.5 and 6.7, respectively.
The compression curves of the LR soil treated with 18.7% PC are shown in
Figure 6.10. The two compression curves are almost identical and show good
consistency of results, indicating that the strain rate effect is negligible. The
preconsolidation pressures for IL042 and CRS043 were 94.40 and 94.83 kPa,
respectively. The compression curves of these two samples shifted further to the
higher effective stress level with respect to the compression curve of
reconstituted LR soil. The results indicate that when treated with 18.7% PC, LR
soil can sustain almost two times higher vertical effective stress than
reconstituted LR soil at the same void ratio.

281
The compression curves of LR soil treated with 18.7% PC are plotted in Figure
6.11 in terms of void index versus vertical effective stress. The normalized
compression curves are parallel to the intrinsic and sedimentation compression
lines, and do not show destructuring (i.e. do not converge to the intrinsic
compression line) within the vertical effective stress range investigated. When
treated with 18.7% PC, the compression line still lies below that of the
undisturbed block LR soil up to the vertical effective stress of about 780 kPa. At a
vertical effective stress > 780 kPa, the 18.7% PC treated LR soil can sustain two
times higher vertical effective stress than the undisturbed block LR soil at the
same void ratio. This is attributed to the structure formed by cementation. It
should be also noted that the treated LR soil exhibits lower compressibility than
the undisturbed soil.
The average compression index of LR soil treated with 18.7% PC was 1.327 (2
4 p), 1.331 (4 8 p) and 1.280 (8 16 p). The results indicate that the
compression index increased by 2.63% between 2 and 4 p, and decreased by
1.04% and 1.76% in the 4 8 p range and the 8 16 p range, respectively,
compared to the values of the reconstituted soil. The swelling index of LR soil
treated with 18.7% PC was 0.028 for OCR = 1 2 OCR and 0.042 for OCR= 2
4, a 70.4% and 71.0% decrease compared to the reconstituted LR soil. The
reloading index also decreased with treatment with 18.7% PC to 0.026 (OCR =24) and 0.084(OCR=1-2).
The values of the constrained modulus of LR soil treated with 18.7% PC are
compared to the data for the reconstituted LR soil in Figure 6.12. In the over
consolidated region, the treated soil exhibits stiffer response with higher
constrained modulus, but once consolidated past the preconsolidation pressure,
the constrained modulus becomes the same as the value for the reconstituted LR
soil.

282
Figure 6.13 presents the cumulative excess pore pressures generated in
CRS043 normalized by the actual strain rate. The excess pore pressures of
reconstituted LR soil were generated in LR soil treated with 18.7% PC were
about 20 % of the excess pore pressures generated in the reconstituted soil at
the same void ratio. Note that the constrained modulus, which is the stiffness of
the soil in 1-D, did not change significantly after treatment. The small magnitude
of excess pore pressures in this specimen indicates that the hydraulic
conductivity of the soil increased with treatment.
In the IL consolidation test, the excess pore pressures generated in each load
increment were also smaller than in the reconstituted LR soil. The excess pore
pressures generated were about 70 79% of the load increment in the over
consolidation region and about 85 90% of the load increment in the normally
consolidated region (Table 6.7).
The values of hydraulic conductivity of LR soil treated with 18.7% PC are plotted
against void ratio in Figure 6.14. The hydraulic conductivity obtained in the IL
consolidation test (IL042) was consistently smaller than in CRS043 by 15 20%
at the same void ratio. The results indicate that the average values of the
hydraulic conductivity of LR soil increased by approximately 4.6 times the value
for the reconstituted LR soil at the same void ratio. However, the slope of the
hydraulic conductivity curve did not change. The average values of Ck/eo and
Ck/Cc of LR soil treated with 18.7% PC were 1/ (4.23 0.31) and 1/ (1.51 0.07),
respectively. These values represent approximately 1.05% and 0.98% decrease,
respectively, with respect to the values for reconstituted soil. These differences
are within the range of experimental error.
Figure 6.15 presents the coefficient of consolidation of LR soil treated with 18.7%
PC. The values obtained from these two tests show rather large variation, with
the values obtained from the CRS consolidation test being consistently higher.

283
The variation of the results between the two tests may be related to the
experimental error associated with the determination of the EOP in the IL test, as
mentioned in section 5.5.2. However, the results of both of these two tests are
considerably above the reconstituted LR soil. The average value of the
coefficient of consolidation of LR soil treated with 18.7% increased by
approximately 8.6 times with respect to the reconstituted LR soil at the same
effective stress.
The effect of treatment with 18.7% PC on the creep behavior of LR soil was
evaluated by performing a creep test at a vertical effective stress of 1577.8 kPa.
The time settlement curve and the dissipation of excess pore pressure are shown
in Figure 6.16. As summarized in Table 6.7, the end of primary was reached after
153.1 minutes. The load increment was applied for 18.6 days and the specimen
was allowed to creep for 2.2 log cycles of secondary compression. During the
time investigated, the slope of the time settlement curve remained constant,
yielding the secondary compression index (C) equal to 0.080. The C/Cc of LR
soil treated with 18.7% was estimated as 0.063. Note that although the C/Cc of
LR soil decreased by 33% compared to the value for reconstituted LR soil, the
ratio still lies in the average values of C/Cc for peats and organic soils.

6.2.4. 1-D Consolidation behavior of 51.4% PC treated LR soil


The effects of treatment with 51.4%PC were evaluated by performing one CRS
(CRS051) and one IL (IL050) consolidation tests. The results of the IL
consolidation test are summarized in Table 6.8. Figure 6.17 presents the
compression curves of LR soil treated with 51.4% PC. The compression curves
obtained from the two consolidation tests are in excellent agreement. The over
consolidated region of LR soil treated with 51.4% PC extends to higher vertical
effective stresses than the reconstituted LR soil. The average preconsolidation
pressure increased substantially to 403.3 1.7 kPa. This value represents

284
approximately an eight fold increase with respect to the preconsolidation
pressure of reconstituted LR soil.
The normalized compression curves of LR soil treated with 51.4% PC expressed
in terms of the void index (Iv) are plotted in Figure 6.18. The figure shows that the
compression curves lie above the sedimentation compression line in the normally
consolidated region. This indicates that the LR soil treated with 51.4% PC can
sustain more than five times higher vertical effective stress than reconstituted soil
at the same void ratio. The figure also shows that the compression line of 51.4%
PC treated LR soil lies above the compression line of the block LR soil in the
normally consolidated region, and exhibits less compressibility than the block LR
soil. The virgin compression line is almost parallel to both the intrinsic and the
sedimentation compression lines, and does not show any destructuring within
vertical effective stresses up to 1000 kPa.
In CRS051, the compression curve of the second loading extended further to a
higher effective stress range passing the line extrapolated from the virgin
compression line. This is shown in Figure 6.19, where a dashed line represents
the line extrapolated from the virgin compression line of the first loading. The
preconsolidation pressure of the compression curve from the second loading was
1361.0 kPa. The result indicates that the specimen was over consolidated with
the OCR of 1.36 at the vertical stress of 1003.6 kPa at which the specimen was
allowed to creep.
As mentioned in section 5.2.1, soils develop preconsolidation pressure as a
result of secondary compression (Mesri and Castro, 1986). The equation for
estimation of the overconsolidation resulting from secondary compression is
given in eq. (6.3) as:
t
'
OCR = vc =
'vi t p

( C

Cc )(1Cr Cc )

(6.3)

285
The OCR of the LR soil treated with 51.4% PC due to secondary compression
estimated using: t (creep time) = 4320 min, tp (time for EOP) = 18.3 min
(estimated with Cv = 0.0043 cm2/sec at 1003 kPa, see Figure 6.24), C/Cc =
0.042 (discussed later), Cr/Cc = 0.029 (see Table 6.5) is

OCR =

'vc

4320 min
=

1003.6kPa 18.3min

( 0.042 )(1 0.029 )

= 1.25

The preconsolidation developed during creep (vc) is estimated as 1254 kPa.


The actual preconsolidation pressure measured from the CRS051 during second
compression is about 8.5% higher than the preconsolidation pressure estimated
with eq. (6.3). This may indicate that (1) the cement was still hydrating and
additional structure was developed while the test was in progress, or (2) that the
increase in the vertical effective stress at which the soil was allowed to creep has
an additional effect on the development of higher preconsolidation pressure.
The time frame of the test CRS051 is indicated in Figure 6.20. In CRS051,
following consolidation of the soil specimen to the vertical effective stress of
1003.6 kPa, the specimen was allowed to dissipate the excess pore pressures
and creep for three days. The specimen was unloaded to a vertical effective
stress of 102.7 kPa with a constant strain rate of 0.05%/hr, followed by a three
day dissipation and creep at 102.7 kPa. The specimen was reloaded again with a
strain rate of 0.5%/hr to a vertical effective stress of 2014.6 kPa. The whole test
was finished within 25.4 days after mixing the soil with PC. It should be noted
that, as mentioned in section 6.2.1, most of hydration of cement in general is
considered to be completed within 28 days (Lea, 1956). The mechanism
responsible for increase in the preconsolidation pressure during creep will be
further discussed later in this section.
Compared to reconstituted LR soil, the compression index of LR soil treated with
51.4% PC increased by 31.4 % to 1.699 0.030 in the 2 4 p range. However,
the LR soil treated with 51.4% PC exhibits much stiffer response upon unloading

286
and reloading. The swelling index significantly decreased by approximately
82.0% with respect to the reconstituted LR soil to 0.017 (1 -2 OCR) and 0.026 (2
4 OCR). The recompression index also decreased by 93.5% and 86.6% to
0.013 (4 2 OCR) and 0.048 (2 1 OCR), respectively.
The constrained modulus of LR soil treated with 51.4% PC is plotted against
vertical effective stress in Figure 6.21. The LR soil treated with 51.4% PC
exhibits a very stiff response in the over consolidated region, with the constrained
modulus being approximately one order of magnitude higher than the
reconstituted LR soil. However, in the normally consolidated region, the LR soil
treated with 51.4% PC has stiffness similar to that of the reconstituted LR soil at
the same vertical effective stress. This suggests that there is some bond
breaking once the soil is compressed beyond the preconsolidation pressure.
The cumulative excess pore pressures generated in CRS051 are normalized with
the actual strain rate and plotted in Figure 6.22. The increase in the constrained
modulus in the over consolidated region caused generation of very small excess
pore pressure. Although the stiffness of the treated LR soil was slightly smaller
than the reconstituted LR soil in the normally consolidated region, significantly
smaller excess pore pressures were generated. In addition, the rate of the
increase in the excess pore pressure (slope of the curve) was also smaller in
CRS051. This is related to the increase in the hydraulic conductivity of LR soil
with treatment. As shown in Figure 6.23, the hydraulic conductivity of LR soil was
4.0 5.7 times larger than the values for reconstituted LR soil at the same void
ratio. Similar to the previous two cases (8.0% and 18.7% PC), the slope of the
hydraulic conductivity curve did not change significantly. The average values of
Ck/eo and Ck/Cc for LR soil treated with 51.4% PC are 1/ (3.81 0.18) and 1/
(1.63 0.08), respectively. The Ck/eo and Ck/Cc are closely related to the
compressibility of the soil, since the change in hydraulic conductivity depends
mainly on the change in the void ratio of the soil. No significant changes in these

287
two values indicate that the change in the void ratio, thus the compressibility, of
the soil did not change significantly with treatment.
The coefficient of consolidation of LR soil treated with 51.4% is plotted against
vertical effective stress in Figure 6.24. The results indicate that in the high OCR
range, the coefficient of consolidation increased slightly with decreasing OCR.
This result is related to a slight increase in constrained modulus and relatively
small change in hydraulic conductivity. In the 1 2 OCR range the coefficient of
consolidation decreased sharply due to decreases in both the constrained
modulus (strain softening behavior) and the hydraulic conductivity. The
continuous but smaller decrease in the coefficient of consolidation in the normally
consolidated region is caused by the fact that the rate of increase in the
constrained modulus is higher than the rate of decrease in hydraulic conductivity.
The coefficient of consolidation of LR soil treated with 51.4% PC was
approximately 1.7 2.0 orders of magnitude larger than that of the reconstituted
LR soil in the normally consolidated region. This result indicates that
consolidation will be completed about 50 100 times faster in soils treated with
51.4% PC.
The effect of treatment with 51.4% PC on the creep behavior of LR soil was
evaluated by performing a creep test at a vertical effective stress of 1574.2 kPa.
The settlement and the dissipation of excess pore pressures are plotted against
elapsed time in Figure 6.25. In the IL consolidation test, the generation of excess
pore pressures was very poor, especially in the over consolidated region. This
may be related to the high constrained modulus or stiffness of the soil. As
summarized in Table 6.8, the excess pore pressures generated in the over
consolidated region were about 29 70% of the load increment except at a
vertical effective stress of 49.3 kPa. Even in the normally consolidated region, the
excess pore pressures were about 55 86% of the load increment. However, as
the vertical effective stress increased, the generation of excess pore pressures

288
increased. At a vertical effective stress of 1574.2 kPa, the excess pore pressures
were 86% of the load increment. It should be noted that even though the
generation of excess pore pressures was poor in this test, the results including
the compression curve and the hydraulic conductivity and coefficient of
consolidation, were consistent with the results obtained from the CRS
consolidation test. The end of primary was reached after 38.7 minutes in this long
term creep test. The load increment was applied for 22.0 days and the specimen
was allowed to creep for 2.9 log cycles of secondary compression. In this time
range investigated, the secondary compression index remained constant at a
value of 0.070. The C/Cc ratio of the LR soil treated with 51.4% PC obtained
with the compression index from CRS051 is 0.042. This value falls in the average
range of C/Cc for inorganic clays and silts (0.05 0.01).
As described previously, the recompression curve of LR soil treated with 51.4%
PC after creep exhibits over-shooting, passing the virgin compression curve
extrapolated from the first compression. The preconsolidation pressure during
second compression was about 8.5% higher than the preconsolidation pressure
estimated employing equation (6.3). To further investigate the mechanism
responsible for the increase in the preconsolidation pressure during creep, two
CRS consolidation tests, CRS068 and CRS069, were performed on LR soil
treated with 49.1% PC cured under a surcharge of 48 kPa for 14 days.
In CRS068, the specimen was allowed to creep at a vertical effective stress of
1004 kPa for 7 days to evaluate the effect of increase in the creep time. In
CRS069, the specimen was allowed to creep at three vertical effective stresses;
257 kPa, 850 kPa, and 2450 kPa to evaluate the effect of the vertical effective
stress. In both tests, the specimens were continuously loaded after creep without
unloading and reloading.

289
The compression curves of CRS068 and CRS069 are shown in Figure 6.26. The
vertical effective stresses at which the specimens were allowed to creep are
shown in the figure. The compression curve of CRS068 is shown in Figure 6.27.
As shown in the figure, the compression line after 7-day creep extended to higher
vertical stress level, passing the virgin compression line indicated with a dashed
line. The virgin compression line is estimated by projecting the virgin
compression line with a compression index 1.61 to a higher vertical effective
stress level. The projected virgin compression line intersects the compression
curve at a vertical effective stress of 1290 kPa. This value is assumed as the
theoretical preconsolidation pressure. The preconsolidation pressure measured
with the compression curve is about 1430 kPa. This value represents about a
14% increase with respect to the theoretical preconsolidation pressure. The
increase in the preconsolidation pressure of the specimen after 3-day creep in
CRS051 was about 8.5%. The results indicate that the preconsolidation pressure
of PC treated LR soil increases with increasing creep time. However, as
discussed previously, the increase in preconsolidation pressure of PC treated LR
soil is higher than the increase expected for natural clays based on equation
(6.3).
The compression curves of CRS069 after 7 day creep at vertical effective
stresses of 257 kPa, 850 kPa, and 2450 kPa are shown in Figure 6.28 through
6.30, respectively. Since the first creep at a vertical effective stress of 257 kPa
was allowed in the over-consolidated region, the virgin compression curve
without creep was estimated using the same curvature of the compression curve
prior to creep. The estimated virgin compression is indicated with a dashed line
in Figure 6.28. In Figure 6.29 and 6.30, the virgin compression curves without
creep are estimated from the compression indices obtained from the virgin
compression curves before the creep. The projected compression curves are
also indicated with dashed lines in Figure 6.29 and 6.30. As shown in Figures
6.28 through 6.30, the compression curve of CRS069 extended to higher vertical

290
effective stress level after creep. The measured preconsolidation pressures are
approximately 15%, 14% and 21% higher than the theoretical preconsolidation
pressures after 7 day creep at 257kPa, 850 kPa, and 2450 kPa, respectively.
The theoretical preconsolidation pressures were estimated as the intersections of
the estimated virgin compression curves (dashed lines) and the actual virgin
compression curves (solid lines). As mentioned before, the increase in the
preconsolidation pressure after 7 day creep (CRS051) at 1000 kPa was about
14%. This value is the same as the increase in the preconsolidation pressure
after 7 day creep at 850 kPa. The results indicate that the preconsolidation
pressure of PC treated LR soil after creep generally increases with increasing
vertical effective stress level.

6.2.5. 1-D Consolidation behavior of 100% PC treated LR soil


The effects of treatment with 100% PC on the 1-D consolidation behavior of LR
soil were evaluated by performing two CRS (CRS055 and CRS064) and one IL
(IL054) consolidation tests. The specimens CRS055 and IL054 were treated with
103.4% PC, while specimen CRS064 was treated with 100.8% PC. The two CRS
consolidation tests were performed with a 5000lbs capacity load cell and the
specimens were consolidated to a maximum vertical effective stress of
approximately 6400 kPa. However, in the IL consolidation test, a 2000 lbs
capacity load cell was employed and the specimen was consolidated only to
1387.0 kPa. The results of the CRS and IL consolidation test are summarized in
Tables 6.5 and 6.9, respectively.
The compression curves of LR soil treated with 103.4% and 100.8% PC are
shown in Figure 6.31. The results show good repeatability. Even with the 2.6%
difference in the cement content and 0.12 difference in the initial void ratio
between CRS055 and CRS064, the compression curves are almost identical in
the normally consolidated region. The preconsolidation pressures of LR soil
treated with 103.4% and 100.8% PC were 1249.8 kPa and 1129.7 kPa,

291
respectively, with an average preconsolidation pressure of 1189.8 84.9 kPa.
This value represents an eight fold increase over the preconsolidation pressure
of reconstituted LR soil. The preconsolidation pressure of the specimen IL054
was not determined since there was only one EOP data point in the normally
consolidated region. However, the compression curve was in good agreement
with CRS055, and it was assumed that the preconsolidation pressure was the
same.
Similar to the LR soil treated with 51.4% PC, the recompression curves of LR soil
treated with 103.4% and 100.8%PC extended further to higher vertical effective
stress than the virgin compression line, as shown in Figures 6.32 and 6.33,
respectively. As shown in Figures 6.32 and 6.33, the virgin compression lines
extrapolated from the compression curves from the first compression, intersect
the recompression curves at 5127.7 kPa for 103.4%PC treated LR soil and at
3811.8 kPa for 100.8% PC treated LR soil. These values represent the
theoretical preconsolidation pressures of these two specimens. As discussed in
section 6.2.4, the OCR value estimated with equation (5.1) for LR soil treated
with 51.4% PC underestimated the actual preconsolidation pressure developed
during creep. Similarly the OCR values for LR soil treated with 100% PC
estimated with equation (6.3) yield:

'vc

4320 min
=
Treated with 103.4%PC: OCR =

4011kPa 14.3min

'vc

( 0.021)(10.007 )

4320 min
Treated with 100.8% PC: OCR =
=

3065kPa 13.4 min

= 1.13

( 0.021)(1 0.006 )

= 1.13

The OCR values of LR soil treated with 103.4% and 100.8% PC due to
secondary compression are estimated using: t (creep time) = 4320 min, tp (time
for EOP) = 14.3 minutes for 103.4%PC and 13.6 minutes for 100.8%PC LR soils

292
(estimated with Cv = 0.0048 cm2/sec and 0.0055 cm2/sec, respectively, at the
vertical effective stresses at which the specimens were allowed to creep, see
Figure 6.38), C/Cc = 0.021 (discussed later), Cr/Cc = 0.007 for 103.4%PC and
0.006 for 100.8%PC LR soils (see Table 6.5). The preconsolidation pressures
during creep estimated with equation (6.3) are 4532 kPa and 3463 kPa for LR
soils treated with 103.4%PC and 100.8% PC, respectively.
The actual preconsolidation pressures measured from the compression curves
are 5705.8 kPa for 103.4% PC treated LR soil and 4403.3 kPa for 100.8% PC
treated LR soil. Similar to the LR soil treated with 51.4% PC, the preconsolidation
pressure of 100%PC treated LR soils estimated with equation (6.3) significantly
underestimates the actual preconsolidation pressures developed. The main
controlling parameter in equation (6.3) is the C/Cc of the soil. The results
suggest that the mechanism for development of preconsolidation pressures of
chemically stabilized soils during creep may be different from the mechanism
applicable to naturally cemented soil, and that equation (6.3) can not be applied
for chemically stabilized soils.
The actual preconsolidation pressures represent about 11 16% increase with
respect to the theoretical preconsolidation pressures. The increase in the
preconsolidation pressures of these two specimens during creep is higher than
that of LR soil treated with 51.4% PC. As described in section 6.2.4, the
measured preconsolidation pressure of LR soil treated with 51.4% PC is about
8.5% higher than the preconsolidation pressure estimated with equation (6.3).
The compression curves normalized using the void index (Iv) are plotted in
Figure 6.34. Similar to the normalized compression curves of LR soil treated with
51.4% PC (Figure 6.18), the compression curves of LR soil treated with 100%PC
also lie near the sedimentation compression curve. All three compression curves

293
are almost parallel to both the intrinsic and sedimentation curves, indicating no
significant change in the compression index with treatment.
The average compression index of LR soil treated with 100%PC was 1.382
0.008 in the vertical effective stress range of 2 4 p. The results indicated that
the compression index increased by 7% with respect to the reconstituted LR soil
when treated with 100% PC. The average swelling index of LR soil treated with
100% PC decreased by approximately 94% with respect to the reconstituted LR
soil, to 0.007 0.001 (1 2 OCR) and 0.005 (2 4 OCR). The reloading index
also decreased significantly by approximately 98%, to 0.005 0.001 (4 2 OCR)
and 0.010 0.001 (2 1 OCR).
The values of the constrained modulus of these three specimens are plotted in
Figure 6.35 against vertical effective stress. The LR soil treated with 100%PC
exhibits much stiffer compression curve in both the over consolidated region, with
the constrained modulus being more than one order of magnitude higher.
Although the comparison can not be made with the reconstituted LR soil in the
normally consolidated region, the constrained modulus of 100% PC treated LR
soil seems to follow the same trend as the reconstituted LR soil at the same
vertical effective stress. The small change in the compression index also
indicates that the constrained modulus of 100% PC treated LR soil is the same
as the reconstituted LR soil at the same vertical effective stress in the normally
consolidated region.
The excess pore pressures developed in LR soil treated with 100% PC
normalized by the actual strain rate, are plotted in Figure 6.36 as a function of
void ratio. The excess pore pressures results show good repeatability in the
normally consolidated region. The pore pressure ratio of LR soil treated with
100% PC was about 0.3% of the pore pressure ratio of reconstituted LR soil in

294
the normally consolidated region at the same void ratio. This result indicates that
the hydraulic conductivity of LR soil has increased with 100% PC treatment.
The hydraulic conductivity of LR soil treated with 100% PC is shown in Figure
6.37. Compared to the reconstituted LR soil, the hydraulic conductivity of LR soil
increased by 16 20 times at the same void ratio.
The coefficient of consolidation of LR soil treated with 100%PC is shown in
Figure 6.38. The large increase in the preconsolidation pressure makes it hard to
compare the coefficient of consolidations of 100% PC treated LR soil and
reconstituted LR soil at the same vertical effective stress. The Figure 6.38
indicates that the coefficient of consolidation increased by approximately 35
times.
The creep behavior of LR soil treated with 100% PC was evaluated from IL054 at
a vertical effective stress of 1387 kPa. The settlement curve and dissipation of
excess pore pressure with time are shown in Figure 6.39. The total excess pore
pressure developed was 332 kPa, which is about 48% of the load increment of
689 kPa. The poor generation of excess pore pressure is related to the high
stiffness of the soil. However, as shown in the compression curve and from
parameters such as hydraulic conductivity and coefficient of consolidation, the
poor generation of excess pore pressure did not significantly affect the results.
The end of primary was reached in 24.45 minutes. The total duration of the creep
test was 15.9 days. The specimen was allowed to creep for three log cycles of
secondary compression. As shown in the figure, the slope of the time settlement
curve remained constant at 0.029 within the time range investigated. Note that
the slope of the compression curve at this vertical effective stress was constant
with increasing vertical effective stress (Figure 6.31). The C/Cc ratio of LR soil
treated with 100% PC calculated with the secondary compression index obtained
from the creep test and the compression index obtained from the CRS

295
consolidation test at the same vertical effective stress was 0.021. The C/Cc ratio
decreased to 22% of the ratio for the reconstituted soil, and lies in the range for
granular soils (0.02 0.01)

6.2.6. Change of pH of LR soil with PC treatment


As mentioned in section 2.6., the pH of PC treated soil increases as the hydrating
cement particles release OH- ions to the soil solution. The formation of a
cementitious product, C3S2H3 gel, depends mainly on the pH of the soil solution
of the PC treated soil. If the pH increases and remains above 12.6, the C3S2H3
gel is stable and hardens with time. On the other hand, if too much of OH- ions
are consumed for the formation of CSH and CAH gels and the pH does not
increase above 12.6, C3S2H3 gel transforms to CSH gel. In such case, hardening
with time does not occur. This implies that the development of preconsolidation
pressure and the increase in the stiffness of the PC treated soil depends largely
on the pH of the PC treated soil.
To evaluate the change of pH of LR soil with treatment, LR soil equivalent to 15g
of dry mass was treated with 8, 20, 50, and 100% PC. The cement was
introduced as slurry with a cement to water ratio of 2:1. The ratio of total solids,
including LR soil and cement, to water was 4:1. The first measurement of the pH
of the treated LR soil was recorded one hour after the cement slurry was
introduced. The treated soil was kept in a tightly sealed glass container to
prevent any loss of moisture. The pH values were measured again 1 and 7 days
after treatment to monitor the change with time. The measured pH values of PC
treated LR soil are shown in Figure 6.40. The untreated LR soil was slightly
acidic with a pH value of 5.98 with a solid to water ratio of 4:1. With treatment
with 8.0% PC, the pH increased rapidly to 11.13 one hour after treatment. No
change in pH was observed after one or seven days. The pH of LR soil treated
with 20% PC increased to 11.98 in one hour, and continued to increased to 12.20

296
after one day. No additional increase was observed after one day. With 50.0%
PC treatment, the pH increased to 12.20 in one hour, and further increased to
12.61 in one day. The pH of LR soil reached the threshold value of 12.6 when
treated with 50% PC and more. With treatment with 100% PC, the pH values
increased to 12.32 in one hour and to 12.77 in one day. In each case, no further
increase in pH was observed after one day. It should be noted that since the
water content of the specimens for pH measurement was more than two times
higher than the water content of the PC treated specimens for consolidation tests,
the actual pH of the PC treated specimens used in consolidation tests may be
higher than the pH values measured with a solid to water ratio of 1:4.

6.2.7. Destructuring behavior of PC treated LR soil


The destructuring behavior of PC treated LR soils was investigated employing
the framework developed by Liu and Carter (2000). As described in Section 5.4,
in this framework the destructuring index (b) represents the rate of reduction with
increasing effective stress of the additional void ratio. This represents, at any
given effective stress, the difference between the void ratio sustained by the
natural soil and the void ratio of the same soil in the reconstituted state. A
destructuring index of 0 indicates that the virgin compression curve of the
structured soil is parallel to that of the reconstituted soil and that no destructuring
occurs during compression. On the other hand, a destructuring index of infinity
indicates that the structure of the soil collapses immediately after passing the
preconsolidation pressure (Liu and Carter, 2000). In other words, a higher
destructuring index represents a faster rate of destructuring during compression.
The destructuring index of LR soil treated with 8.00% PC during virgin
compression is shown in Figure 6.41. The compression curve of LR soil treated
with 8.00% PC is almost parallel to the compression curve of reconstituted LR
soil. The result shows that the destructuring index of 8.00% PC treated LR soil is

297
zero, and thus 8.00% PC treated LR does not exhibit destructuring behavior
during virgin compression.
The destructuring index of LR soil treated with 18.7% PC during virgin
compression is shown in Figure 6.42. Similar to LR soil treated with 8.00%PC,
the destructuring index of LR soil treated with 18.7% PC is also equal to zero.
The destructuring behavior of LR soil treated with 51.4 %PC during the first and
the second compression phases is shown in Figure 6.43. As shown in the figure,
LR soil treated with 51.4% PC exhibits little destructing behavior in the normally
consolidated region. The destructuring index during first compression is 0.07.
This value represents a relatively small rate of destructuring, according to Liu and
Carter (2000).
As discussed in section 6.2.4, LR soil treated with 51.4% PC developed
additional structure during creep. As shown in Figure 6.43, destructuring of the
additional structure developed during creep is more significant than the
destructuring during first compression. The destructuring index during second
compression is 0.4. The higher destructuring index during second compression
indicates that the additional structure developed during creep is not as stable as
the structure developed during curing.
The destructuring behavior of LR soil treated with 103.4% PC (CRS055) and
100.8% PC (CRS064) during the first and second compression phases is shown
in Figures 6.44 and 45, respectively. The destructuring indices during first
compression range from 0.11 (100.8% PC) to 0.13 (103.4% PC). For LR soils
treated with PC, the destructuring index generally increases with increasing %PC.
The increase in destructuring index implies that the destructuring is more abrupt
and the rate of destructuring is faster when treated with higher % PC.
The destructuring indices during second compression range from 0.5 (100.8%
PC) to 0.6 (103.4% PC), as shown in Figures 6.44 and 45. Similar to LR soil

298
treated with 51.4% PC, the destructuring index during second compression is
higher than the destructuring index during first compression. In addition, the
destructuring index during second compression increases with increasing %PC,
at least for LR soil treated with more than 50% PC.
The increase in the destructuring index of LR soil with increasing amount of
cement is in good agreement with the results presented by Liu and Carter (2000).
The compression curves and the destructuring indices of Louiseville clay treated
with 2 10% of lime are shown in Figure 6.46. Louiseville clay is an inorganic
silty clay with a plasticity index of 41% and an organic content of 1.0% (Locat et
al., 1996). The figure shows that the virgin compression curve of lime stabilized
Louiseville clay gradually shifts to a higher effective stress with increasing lime
content. This is in good agreement with the results observed in LR soil treated
with PC. The destructuring indices of lime treated Louisville clay are summarized
in the table shown in Figure 6.46. Similar to LR soil treated with PC, the
destructuring index of lime stabilized Louiseville clay gradually increases with
increasing lime content from 0.3 (2% lime) to 0.43 (10% lime). The destructuring
index of LR soil treated with PC is relatively smaller than the destructuring index
of lime stabilized Louiseville. This implies that the structure developed in LR soil
treated with 8 100% PC is relatively more stable than the structure developed
in Louiseville clay treated with 3 10% lime.

6.2.8. Discussion: 1-D consolidation behavior of PC treated LR soil


The compression curves of reconstituted and PC treated LR soil are shown in
Figure 6.47. With increasing cement content, the compression curves shift
gradually to higher vertical effective stresses. The preconsolidation pressure
increases with increasing cement content, and the over consolidated region
extends further to higher vertical effective stresses. The results indicate that LR
soil can sustain higher vertical effective stress at the same void ratio with
increasing %PC. Similar results were observed on other chemically stabilized

299
soils. For example, the compression curves of lime stabilized Louiseville (Locat
et al, 1996) and cement treated Bangkok clay (Lorenzo and Bergado, 2004)
gradually shifted to higher vertical effective stresses with increasing amounts of
binding agents, as shown in Figures 6.48 and 49, respectively.
The increase in the preconsolidation pressure with treatment is shown in Figure
6.50. The percentage cement content and the corresponding cement factor,
which represents the amount of cement (kg) used per unit volume of treated soil
(1 m3), are shown in the lower and upper x- axes. The overall shape of the best
fitting curve of the preconsolidation pressure is quadratic within the cement
content range investigated. The increase in the preconsolidation pressure is
relatively small up to a cement content of 20%. With treatment with 8 and 20%
PC, the preconsolidation pressure increased by approximately 50% and 88% to
74.8 kPa and 94.6 kPa from the preconsolidation pressure of the reconstituted
LR soil. The preconsolidation pressure of LR soil increases sharply when treated
with more than 50% PC. When treated with 50% and 100% PC, the
preconsolidation pressure increases approximately by 7 and 22.7 times,
respectively.
The increase in the preconsolidation pressure with treatment is in good
agreement with the change in pH of LR soil with treatment, as shown in Figure
6.40. The sharp increase in the preconsolidation pressure with 50% and 100%
PC treatment is related to the increase in pH to the threshold value of 12.6 at
which the cementitious product, C3S2H3 gel, remains stable and hardens with
time.
The compression curves of LR soils treated with more than 50% PC extended to
higher effective stresses during second compression. The preconsolidation
pressures of PC treated LR soil increase more with increasing cement content,

300
creep time, and the vertical effective stress at which the specimen was allowed to
creep.
While the preconsolidation pressure of LR soil increases with treatment, the
compression index does not show significant change with treatment. As shown in
Figure 6.47, the virgin compression curves of PC treated LR soil are almost
parallel to or slightly steeper than the compression curve of reconstituted LR soil.
In the vertical effective stress range of 2 4 p, the compression index of LR soil
increased by 2.6% (with 18.7 % PC) - 31.4% (with 51.4% PC) with respect to the
compression index of the reconstituted LR soil. The increase in the compression
index with treatment is related to abrupt destructuring of the soil structure
developed by treatment with PC. The destructuring of the bonds developed
between the cement and LR soil particles occurs near the preconsolidation
pressure. The increase in the compression index indicates that the post yield
compression is more brittle in the PC treated LR soil than in the reconstituted LR
soil. Once the treated LR soil specimens are compressed well into the normally
consolidated region, i.e. beyond 8 times the preconsolidation pressure, the
compression curves become parallel to the compression curve of the
reconstituted LR soil.
The destructuring behavior of PC treated LR soil can be quantitatively compared
by evaluating the destructuring index employing the frame work developed by Liu
and Carter (2000). As discussed in section 6.2.7, the destructuring index
generally increases with increasing cement content. The destruction of the bonds
developed by cementation during curing occurs more slowly than the
destructuring of the additional structure developed during creep.
Figure 6.51 presents the compression curves of PC treated LR soil normalized
using the void index (Iv). The normalization of the compression curves with void
index allows one to evaluate the degree of structure produced with treatment.

301
The intrinsic compression line represents the virgin compression curve of soil
prepared in slurry state with water content of 1 1.5 times the liquid limit. As
shown in this figure, the virgin compression curve of reconstituted LR soil lies on
or slightly above the intrinsic compression line. The sedimentation compression
line represents the virgin compression curve of naturally sedimented shallow
marine deposits. The sedimentation compression line is approximately parallel to
the intrinsic compression line. According to Burland (2000), naturally sedimented
clays can sustain about five times higher vertical effective stress than
reconstituted soils at the same void ratio. Note that the virgin compression curve
of intact LR soil falls on the sedimentation compression line at the
preconsolidation pressure, followed by a sharp increase in the compression
index. With treatment, the normalized compression curves of LR soil gradually
move from the intrinsic compression line to the sedimentation compression line.
The normalized compression curve of LR soil treated with 8.00% PC lies at about
50% higher vertical effective stress than reconstituted LR soil at the same void
index. The additional vertical effective stress is sustained by the soil structure
developed by cementation. With 20% PC treatment, the LR soil can sustain 88%
higher vertical effective stress than reconstituted LR soil at the same void ratio.
The normalized compression curve of LR soil treated with 51.4% PC shifts
significantly to higher vertical effective stresses and lies slightly above the
sedimentation compression line. With this cement dosage, the LR soil can
sustain 5 6 times higher vertical effective stress. Note that the normalized
compression curve of 51.4% PC treated LR soil is slightly steeper than the
sedimentation compression line in the normally consolidation region. The
preconsolidation pressure of LR soil treated with 100 %PC increased further to
1190.0 kPa. However, the compression curve did not shift to higher vertical
effective stress than the compression curve of 50% PC treated LR soils. As
shown in this figure, the normalized compression curve of 100% PC treated LR
soil falls slightly below the sedimentation compression line. The increase in the

302
preconsolidation pressure of LR soil treated with 100 %PC with respect to the LR
soil treated with 51.4% PC is mainly due to the decrease in the initial void ratio.
Although the compression index does not show significant effects of treatment,
the swelling index and reloading index decreased with treatment. The swelling
index showed 55.5%, 70.7%, 82.0% and 94.8% decrease with treatment with 8.0,
18.7, 51.4 and 100% PC at the same OCR. The reloading index decreased by
66.3%, 81.8%, 90.0% and 97.5% with treatment.
The values of the constrained modulus of PC treated LR soils during virgin
compression are shown in Figure 6.52. The results clearly show that the stiffness
of LR soil increases in the over consolidated region with treatment. In the
normally consolidated region, however, the constrained modulus of PC treated
LR soil is almost the same as that of the reconstituted LR soil at the same
vertical effective stress. The values of the constrained modulus of PC treated LR
soil are plotted against normalized vertical effective stress in Figure 6.53.
Compared to the reconstituted LR soil, the constrained modulus of LR soil
increased by 1.3, 1.9, 7.6 and 22.9 times with treatment at the same OCR or at
the same vertical effective stress level relative to the preconsolidation pressure.
The results indicate an increase in stiffness of LR soil with treatment.
Figure 6.54 shows the increase in hydraulic conductivity of LR soil with treatment.
Although the initial void ratio gradually decreases with increasing cement content,
the hydraulic conductivity of LR soil increases with treatment at the same void
ratio. The effect of treatment with 8.0% PC is modest with a 18 26% increase at
the same void ratio. With treatment with 18.7% PC, the hydraulic conductivity
increased by 4.6 times at the same void ratio with respect to the reconstituted LR
soil.

303
The ratio of the change in the hydraulic conductivity to the change in the void
ratio (Ck/Cc) generally decreases with increasing cement content (Table 6.5). As
discussed in Chapter 5, the value of Ck/Cc generally decreases with decreasing
organic content.
The increase in the hydraulic conductivity of LR soil with treatment is mainly due
to the change of the fabric of the soil. As discussed in Chapter 2, soil particles
treated with lime and cement become more flocculated and aggregated, forming
bigger soil particles. According to Terzaghi et al. (1997), the hydraulic
conductivity of a soil generally decreases with increasing particle size, as the
specific surface or wet surface decreases with increasing particle size.
Several researchers observed through particle size distribution analysis and
scanning electron microscope (SEM) analysis that the inorganic clays treated
with lime (Tremblay et al., 2001) or cement (Al-Rawas, 2002) become flocculated
and aggregated yielding a more open fabric and more granular-like particles. As
a result, the size of the macropores, which serve as the main channels for flow,
increases causing the increase in hydraulic conductivity.
Although the results showed that the hydraulic conductivity of LR soil increased
with increasing PC, at least within the %PC range investigated, the hydraulic
conductivity is not expected to increase continuously when treated with
significant amounts of PC. For example, one of the main applications of PC
treatment in the field is to decrease the hydraulic conductivity of soil to form an
impermeable barrier. As such, it is expected that the hydraulic conductivity of LR
soil would decrease with significant increases in the PC dosage.
The coefficient of consolidation of PC treated LR soil is plotted against vertical
effective stress in Figure 6.55. In general, the coefficient of consolidation curve
shifts upwards and to the right indicating a gradual increase in the coefficient of

304
consolidation with treatment at the same vertical effective stress. A more direct
comparison can be made by plotting the coefficient of consolidation against the
vertical effective stress normalized by the preconsolidation pressure, as shown in
Figure 6.56. Similar to the previous observation relative to the preconsolidation
pressure and hydraulic conductivity, the effects of treatment with 8.00% PC on
the coefficient of consolidation are small. Compared to the reconstituted LR soil,
the coefficient of consolidation increased by approximately 26%, 6.6 times, 25
times, and 32 times when treated with 8.00%, 18.7%, 51.4%, and 100% PC,
respectively. The increase in the coefficient of consolidation with PC treatment is
due to increases in both the constrained modulus and the hydraulic conductivity.
Similar results were reported by other researchers for various types of soils
treated with lime and cement (Kassim et al. (1999), Cortellazzo et al. (1999) and
Broms (1999)). The increase in the coefficient of consolidation implies that the
consolidation process will occur faster when the soil is treated with
increasing %PC.
The effect of treatment on the secondary compression behavior of LR soil was
evaluated by performing long term creep tests. The specimens were
consolidated into the normally consolidated region following the EOP
compression line. The results showed that the slope of the time settlement curve
remained constant within the time range investigated. Since the secondary
compression index depends on the stress level, the ratio of C/Cc better
represents the creep behavior of each soil investigated. The secondary
compression index (C) obtained for each PC treated LR soil specimen was
normalized with the compression index obtained from CRS consolidation tests at
the same vertical effective stress, as discussed in Chapter 5. As shown in Figure
6.57, the C/Cc ratio gradually decreased with increasing cement content.
Compared to the reconstituted LR soil, the C/Cc ratio of LR soil decreased by
approximately 10%, 33%, 56%, and 78% when treated with 8.0%, 18.2%, 51.4%
and 103.4% PC, respectively. The decrease in the C/Cc ratio is mainly due to

305
the decrease in the secondary compression index. The secondary compression
index of LR soil decreases with PC treatment as a result of the replacement of
LR soil with less creep-susceptible cement hydration products and of the
increase in the soil particle size with treatment. As shown in Table 2.13, the
typical values of C/Cc for various types of geotechnical materials generally
decrease with increasing particle size. The increase in compression index due to
the structure developed with treatment, as shown in Figure 6.47, also contributes
to the decrease in the C/Cc ratio of LR soil.

6.3. Effects of curing


This section describes the effects of curing, curing time and surcharge applied
during the curing stage.
6.3.1. Effects of curing time
Curing time may contribute to an increase of the effects of treatment in two ways.
The first one is the increase in the duration of secondary compression. As the
specimen is allowed to creep longer, it ultimately leads to an increase in the
preconsolidation pressure. This effect may become more significant if the amount
of cement is small as the soil can creep more, as indicated by the higher C/Cc of
soils treated with lower %PC than of soils treated with higher % PC. If the cement
content is too high, the creep potential of the soil decreases as shown in Figure
6.57 and thus the increase in the preconsolidation pressure due to secondary
compression becomes smaller.
The second is the increase in time for cement reaction. For example, Hebib and
Farrell (2003) observed from the results of unconfined compression strength
tests performed on cement treated Ballydermot peat that the unconfined
compressive strength of 90 days cured samples was about 70% of the strength
of soils cured for 1 year. According to Lea (1956), the compressive strength of
cement paste prepared with a cement to water ratio of 2:1 after 28 days curing

306
was approximately 36 MPa and was about 75% of the 90 day strength. These
results suggest that the strength of PC treated soils should increase with curing
time. For the strength of PC treated soils to increase with time, the pH of the soil
solution has to be greater than the threshold value of 12.6, under which the
cementitious product C3S2H3 gel remains stable and hardens with time.
To assess the effects of curing time on the effects of treatment with PC,
additional consolidation tests were performed with LR soil treated with 50% and
100% PC cured for 14 days and 28 days. These cement contents were selected
because the results of pH measurements indicated that the pH of LR soil
reached the threshold value of 12.6 when treated with more than 50% PC. The
PC treated LR soil samples were prepared by curing under the same surcharge
of 48 kPa. For each % PC, two specimens were prepared: one for CRS and one
for IL consolidation tests. The initial conditions of these soil specimens are
summarized in Table 6.10.
The compression curves of LR soil treated with 50% PC and cured for 14 and 28
days are shown in Figure 6.58. The specimen for CRS051 was cured under a
curing surcharge of 48 kPa for 14 days, while the other two specimens for IL059
and CRS060 were cured under the same curing surcharge for 28 days. The two
CRS consolidation tests (CRS051 and CRS060) presented in this figure were
performed following the same procedure, with the same strain rate, and the same
duration for pore pressure dissipation and creep after loading and unloading. As
shown in the figure, the virgin compression lines match perfectly, and do not
show any effect of curing time. The preconsolidation pressures for CRS060 and
CRS059 were 407.44 kPa and 410.19 kPa, respectively. The average
preconsolidation pressure of LR soil treated with 51.4% and cured for 14 days
under a curing surcharge of 48 kPa was 403.25 kPa. The results indicate no
significant increase in the preconsolidation pressure with curing time.

307
The specimen CRS060 developed a preconsolidation pressure of 1414kPa
during creep. This value represents about a 34% increase with respect to the
vertical effective stress at which the specimen was allowed to creep. This value
is in a good agreement with the increase in the preconsolidation pressure of 14
day cured specimens (CRS051) during creep, as discussed in section 6.2.4.
Destructuring during compression and recompression of CRS060 specimen is
shown in Figure 6.59. The destructuring indices (b) for the first and second
compression stages are 0.08 and 0.4, respectively. The b values for CRS051 are
0.07 and 0.4 during compression and recompression. The results confirm that
the compressibility and the rate of destructuring of the structure developed after
14 and 28 days are almost the same. The results indicate that the effect of curing
time on the compressibility of LR soil, at least within the time range investigated,
is negligible.
The constrained modulus of LR soil treated with 50% PC and cured for 14 and 28
days is shown in Figure 6.60. The results are almost identical and show no
significant effect of curing time on the constrained modulus or the stiffness of the
specimens.
Figure 6.61 presents the hydraulic conductivity of the three specimens as a
function of void ratio. The hydraulic conductivity of the 28 days cured specimens
(IL059 and CRS060) is approximately 10% higher than that of the 14 days cured
specimen (CRS051). This may be within the experimental error, and indicates
that the effect of curing time on the hydraulic conductivity is small.
Figure 6.62 presents the coefficient of consolidation of LR soil specimens treated
with 50% PC cured for 14 and 28 days. The constrained modulus and the
hydraulic conductivity data indicate that the effect of curing time is small on these

308
two parameters. No significant effect of curing time on the coefficient of
consolidation is observed.
The effect of curing time on the creep behavior of 50% PC treated LR soil was
evaluated based on the results of a creep test performed in test IL059 at a
vertical effective stress of 1573 kPa. The time settlement and dissipation of
excess pore pressure curves are shown in Figure 6.63 as a function of time. The
end of primary consolidation was reached in 44.6 minutes. The total duration of
the creep test was 34 days, and the specimen was allowed to creep for three log
cycles of secondary compression. The secondary compression index of the
specimen was 0.073. As shown in Figure 6.63, the slope of the time settlement
curve during secondary compression did not change but remained constant
within the time range investigated. This is in good agreement with the constant
slope of the compression curve of CRS060 in this vertical effective stress range.
Note that the secondary compression index obtained from 50% PC treated LR
soil specimen cured for 14 days (IL050) was 0.067 at the same vertical effective
stress. The difference between the coefficient of secondary compression of 50%
PC treated LR soil cured for 14 and 28 days is very small, about 8%. The C/Cc
ratio for the 28 days cured specimen was 0.042 and is in good agreement with
the value of 0.042 of the 14 days cured LR soil (IL050). The results indicate
negligible effect of curing time on the secondary consolidation behavior of LR soil
treated with 50% PC.
Figure 6.64 present the compression curves of LR soils treated with 100.8% PC
and cured for 14 days (CRS064) and for 28 days (CRS065 and IL066). The
compression curves of 28 days cured specimens extended to slightly higher
vertical effective stress levels than the 14 days cured specimen at the same void
ratio. The preconsolidation pressure of the specimen CRS065 was 1211.5 kPa.
The preconsolidation pressure of the specimen IL066 was not determined since
the specimen was not consolidated further into the normally consolidated region.

309
The preconsolidation pressure of the 14 days cured specimen treated with the
same amount of PC was 1129.7 kPa. The result indicates approximately a 7.0%
(or 81.8 kPa) increase in the preconsolidation pressure after 28 days curing. The
increase in the preconsolidation pressure may be within the range of
experimental error.
Figure 6.65 shows the destructuring of 28 days cured LR soil specimen during
first and second compression stages. The b value of the 28 days cured specimen
during compression was 0.07. The results indicate that the structure of the 28
days cured specimen is destructured at a slightly slower rate than the structure of
the 14 days cured specimen. Note that the b value of the 14 days cured
specimen during compression was 0.11 (Figure 6.45). The destructuring index
during the second compression stage did not show any effect of increase with
curing time. The b value during second compression was 0.05 and was similar to
the value (0.04) for the 14 days cured specimen. It should be noted that the
duration of the two dissipation stages and the time taken for unloading and
reloading was the same for the 14 and 28 days cured specimens.
The constrained modulus of the 14 and 28 days cured specimens are shown in
Figure 6.66. As shown in the figure, the constrained modulus of the 28 days
cured specimens are slightly higher than that of the 14 days cured specimen in
the over consolidation region. This indicates that the structure developed in the
28 days cured specimen is slightly stiffer than the structure in the 14 days cured
specimen.
Figure 6.67 presents the hydraulic conductivity of the 14 and 28 days cured LR
soil specimens treated with 100.8% PC. No significant change in the hydraulic
conductivity after 28 days curing was observed, although the hydraulic
conductivity of the 28 days cured specimen was slightly smaller than that of the

310
14 days cured specimen. The difference in the hydraulic conductivity is
approximately less than 10%. This is within the range for experimental error.
Figure 6.68 presents the coefficient of consolidation of the 14 and 28 days cured
specimens. The coefficient of consolidation of the 28 days cured specimen is
slightly higher than the 14 days cured specimen, by less than 10%. Similar to the
hydraulic conductivity, no significant effect of curing time was observed on the
coefficient of consolidation.
The effect of curing time on the creep behavior of 28 days cured LR soil treated
with 100.8% PC was evaluated at a vertical effective stress of 1579 kPa. The
settlement and dissipation of excess pore pressure curves are shown in Figure
6.69. The end of primary was reached in 17.9 minutes. The specimen was
allowed to creep for 46629 minutes or 3.4 log cycles of secondary compression.
The slope of the time settlement curve in the secondary consolidation stage is
constant within the time range investigated. This result is in good agreement with
other long term tests performed on PC treated LR soils. The secondary
compression index of the 28 days cured specimen was 0.0386. The value of
C/Cc estimated using the compression index obtained from the CRS
consolidation test is 0.027. This value is close to that obtained for the 14 days
cured specimen (0.021). The results indicate that the effect of curing time is
negligible on the secondary consolidation behavior of PC treated soils.
As the results indicate, the effect of curing time on the 1-D consolidation behavior
of LR soil treated with 50% and 100% PC is not significant. However, LR soil
treated with 100% PC shows a slight increase in stiffness, as indicated by a
slightly higher constrained modulus, and a slight shift of the virgin compression
curve to a higher effective stress with increasing curing time.

311
6.3.2. Effects of curing surcharge
The initial loading plays a significant role in the mechanical behavior, especially
the strength, of chemically stabilized soil. Humphrey (2001) observed from
unconfined compression tests performed on PC and lime treated LR soil that the
unconfined compression strength and the stiffness of the soil increase
substantially when the soil was cured under higher surcharge. Similar results
were reported by Hebib and Farrell (2003) in that the compressive strength of
preloaded cement treated peat samples were higher than the strength of
specimens that was not preloaded. Another advantage of applying surcharge
during curing is that the application of surcharge improves the quality of the
chemically treated soil by squeezing out the air bubbles that may have been
introduced during the mixing process in the field (Hebib et al, 2002).
The time of application of the curing surcharge is also important. According to
hnberg et al. (2001), the unconfined compressive strength of cement treated

peat prepared by applying a surcharge 24 hours after mixing was about 25% of
the strength of samples that were preloaded immediately after mixing.
The effect of the curing surcharge was evaluated by performing consolidation
tests on LR soil specimens treated with 50% PC. The specimens were cured
under a surcharge of 48 kPa, 96 kPa and 192 kPa for 14 days. These surcharges
represent approximately 2.5, 5.0 and 10.0 m thick fills assuming that the unit
weight of the fill is 19 kN/m3. The initial conditions of the samples for evaluation
of curing surcharge are summarized in Table 6.11.
The compression curves of 50% PC treated LR soil cured under 48 kPa, 96 kPa
and 192 kPa are shown in Figure 6.70. The specimen cured under a higher
surcharge has a lower initial void ratio than the specimen cured under a lower
surcharge. However, the results shows that all the compression curves follow the
same virgin compression curve in the normally consolidated region. The average

312
preconsolidation pressures increases with increasing curing surcharge from
403.4 1.7 kPa (surcharge = 48 kPa) to 471.4 5.0 kPa (surcharge = 96 kPa),
and 672.7 24.4 kPa (surcharge = 192 kPa). The results indicate that the
preconsolidation pressure of PC treated LR soil increases with increasing
surcharge but the virgin compression curve remains the same regardless of the
curing surcharge. The increase in the preconsolidation pressure is mainly due to
the decrease in the initial void ratio.
The preconsolidation pressures of LR soil treated with 50% PC are plotted as a
function of the surcharge in Figure 6.71. The solid line plotted in the figure
represents a 1:1 increase in preconsolidation pressure with curing surcharge.
The results indicate that the increase in the preconsolidation pressure is higher
than the increase in the curing surcharge and that LR soils cured under higher
curing surcharge developed preconsolidation pressures higher than the increase
in the applied curing surcharge. This implies that the preconsolidation pressure of
PC treated soils can be enhanced by curing the soil under higher surcharge.
Figure 6.72 shows the constrained modulus of 50% PC treated LR soils cured
under varying surcharges. The figure clearly shows that the constrained modulus
increases with increasing curing surcharge at the same vertical effective stress in
the over consolidated region. In the normally consolidated region, on the other
hand, the constrained modulus is almost identical at the same vertical effective
stress regardless of the curing surcharge. The increase in the constrained
modulus in the over consolidated region is related to the increase in the
preconsolidation pressure. The results indicate that an increase in the curing
surcharge is also effective to improve the compressibility of PC treated soils in
the over consolidated region.
Destructuring of LR soils cured under a surcharge of 96 kPa and 192 kPa is
shown in Figures 6.73 and 74, respectively. The destructuring index of LR soil

313
during compression decreased gradually with increasing curing surcharge. The
destructuring index of LR soil cured under 48, 96 and 192 kPa was 0.07 (Figure
6.43), 0.04 and 0.0, respectively. This indicates that the destructuring rate of the
structure of PC treated LR soil can be decreased by curing soils under higher
surcharge. The destructuring index of these soils during the second compression
was almost the same: 0.4, 0.4 and 0.3, respectively, and showed no effects of
curing surcharge.
Figure 6.75 presents the hydraulic conductivity of 50% PC treated LR soil cured
under varying surcharges. The hydraulic conductivity of specimen CRS063 is not
included due to large scatter. The results indicate that the hydraulic conductivity
increases gradually with increasing curing surcharge. The increase in hydraulic
conductivity is more apparent in the over consolidated region. The decrease in
the initial void ratio, as shown in Figure 6.70, indicates that the soil particles in
the specimen cured under higher curing surcharge are more tightly compacted.
The gradual increase in the hydraulic conductivity with increasing curing
surcharge, however, indicates that the LR soil cured under higher surcharge
exhibit higher degree of flocculation and aggregation, and formed larger soil
particles. In general, soils with larger particle size exhibit higher hydraulic
conductivity than soils with smaller particles, and the hydraulic conductivity
generally increases with increasing particle size (Terzaghi et al, 1997). Note that
the hydraulic conductivity of LR soil cured under 192 kPa dropped sharply once
consolidated into the normally consolidated region. This may indicate closure of
macropores that serve as a main channel for pore water. However, the effect of
curing surcharge on hydraulic conductivity needs further investigation employing
microscopic analysis, such as SEM to investigate the change in the pore size.
Figure 6.76 shows the coefficient of consolidation of 50% PC treated LR soil
cured under varying surcharges for 14 days. The results indicate that the
coefficient of consolidation increases with increasing curing surcharge. This is

314
related to the increase in constrained modulus and hydraulic conductivity with
increasing curing surcharge, as previously discussed.

6.3.3. Discussion: Effects of curing on the 1-D consolidation behavior of PC


treated LR soil
The importance of curing of chemically stabilized soil is observed by many
researchers. Humphrey (2001) observed that stiffness and unconfined
compressive strength of LR soils treated with PC and lime increased with
increasing curing time and curing surcharge. Hebib and Farrell (2003) observed
that the unconfined strength of PC stabilized peat increased when the soil was
preloaded during curing. They also observed that the unconfined compressive
strength increased with increasing curing time, and the increase in strength was
higher for preloaded soils. Ahnberg et al. (2001) observed that the unconfined
compressive strength of cement treated peat increased when the surcharge was
applied immediately after treatment.
In this experimental program, the effect of curing time was evaluated based on
the test results obtained from LR soils treated with 50% and 100% PC cured for
14 and 28 days. The cement content was selected since the pH of the soil
solution was higher than 12.6 when treated with 50% and more PC. At this pH
level, the C3S2H3 gel remains stable and hardens with time. The results indicate
that the effect of curing time on 50% PC treated soil is negligible. The
preconsolidation pressure and the constrained modulus remained the same after
being cured for 28 days. The hydraulic conductivity slightly increased by 10% at
the same void ratio. This may be related to further flocculation and aggregation
due to CSH and CAH gels. No significant increase in the coefficient of
consolidation was observed for LR soil cured for 28 days.
Slight increases in the preconsolidation pressure and stiffness with increasing
curing time were observed on LR soil treated with 100% PC. The

315
preconsolidation pressure of 100% PC treated LR soil increased by 7% after 28
days of curing. The increase may be related to the continuous hardening of
C3S2H3 gel during the time frame investigated. The constrained modulus also
increased after 28 days for LR soil treated with 100% PC. However, no
significant changes were observed in the hydraulic conductivity and coefficient of
consolidation.
The effect of curing surcharge was investigated by performing consolidation tests
on 50% PC treated LR soil cured under surcharges of 48, 96 and 192 kPa. The
results showed that the preconsolidation pressure gradually increased with
increasing curing surcharge from 403.4 1.7 kPa to 471.4 5.0 kPa, and 672.7
24.4 kPa. The increase in the preconsolidation pressure was higher than the
increase in the curing surcharge. The constrained modulus of 50% PC treated
LR soil also increased with increasing curing surcharge. However, the virgin
compression curves were identical. The increase in preconsolidation pressure in
mainly due to the combination of a decrease in the initial void ratio and an
increase in the constrained modulus in the over consolidated region. It was
observed that the hydraulic conductivity in the over consolidated region
increased with increasing surcharge. This may be related to further flocculation
and aggregation of soil particles due to the CSH and CAH gels. The hydraulic
conductivity in the normally consolidated region showed no significant increase
with curing surcharge. The coefficient of consolidation also increased with
increasing surcharge. The increase is mainly due to the increase in the
constrained modulus.

6.4. Framework: Effects of treatment on the 1-D consolidation behavior of


soils
The effects of chemical treatment on the 1-D consolidation behavior of organic
soils are evaluated by reviewing the results of the consolidation tests performed

316
in this experimental program and work done by other researchers on various
organic soils treated with either cement or lime.
The effect of chemical treatment on the virgin compression behavior is
schematically shown in Figure 6.77. The virgin compression line of chemically
stabilized soil shifts gradually to higher vertical effective stress level (see also
Figure 6.47, 6.48 and 6.49). As shown in Figure 6.77, chemically stabilized soils
can sustain higher vertical effective stress than its natural or in-situ state at the
same void ratio. The shift of the virgin compression to higher stress depends
mainly on the amount of binding agent used for treatment, organic content, and
the pH of the pore water of stabilized soil. As discussed in Chapter 2, the
cementitious product C3S2H3 gel which increases the strength of the soil, is
stable when the pH of the pore water of cement stabilized soil is above 12.6. For
lime stabilization, the optimum lime content is determined as the lime content at
which the pH of the pore water reaches 12.4. In general, organic soils with
higher organic content have lower pH. Therefore, more cement or lime dosage is
likely to be required for organic soils with higher organic content to bring the pH
to 12.6 and to improve their consolidation behavior. In addition, the structure of
the cement treated soil is stable and does not show significant destructuring in
the virgin compression range. As discussed in section 6.1, the destructuring
values of cement treated LR soil are in the range of 0 0.13, indicating no (b=0
for 8 and 18.7% PC specimens) or very small destructuring (b=0.13 for 100% PC
specimens)
The effect of chemical treatment on the preconsolidation pressure is
schematically shown in Figure 6.78. The preconsolidation pressure of cement
treated soil generally increases with treatment, as the virgin compression line
shift to higher vertical effective stresses. The increase in the preconsolidation
pressure depends mainly on the amount of binding agent, organic content, and
the pH of the pore water. As shown in Figure 6.50, the preconsolidation pressure

317
of LR soil increased with increasing PC content. The preconsolidation pressure
of LR soil increased substantially when treated with more than 50% PC. The pH
of the LR soil increased above 12.6 when treated with more than 50% PC.
The preconsolidation pressure of a soil treated with a binding agent increases
also with increasing surcharge applied during curing, and with decreasing water
content. This is due to the fact that the initial void ratio of cement treated soil
decreases with increasing curing surcharge and decreasing water content. As
discussed in section 6.3.2, the preconsolidation pressure of LR soil treated with
50% PC increases with increasing curing surcharge. In addition, the increase in
the preconsolidation pressure was higher than the increase in the curing
surcharge. According to Lorenzo and Bergado (2004), the preconsolidation
pressure of cement stabilized soil increases as the after curing water content of
the soil decreases (see Figure 6.49). As the curing surcharge increases or water
content decreases, the initial void ratio of cement treated soil decreases. As
shown in Figure 6.78, the recompression curve intersects the virgin compression
line at a higher vertical effective stress as the initial void ratio decreases.
However, as noted above, the position of the virgin compression line remains the
same regardless of the curing surcharge or water content.
The effects of treatment on the compressibility show two opposite trends
depending on the vertical effective stress level. The compressibility in the over
consolidated region (i.e. recompression index and swelling index) generally
decreases with treatment (see section 6.2.8) as the stiffness of the soil
increases. The compressibility in the normally consolidated region, on the other
hand, remains the same or slightly increases with treatment. The increase in the
compression index with treatment is due to abrupt and brittle destructuring of the
chemical bonds developed between soil particles and binding agents. The
compression index and the destructuring index, b, generally increases with

318
increasing amount of binding agent. The same results were observed by
Tremblay et al. (2001), Cortellazzo and Cola (1999), and Liu and Carter (2000).
The results of compression tests performed on PC treated LR soil showed that
the constrained modulus remained the same or slightly increased with increasing
cement content at the same vertical effective stress in the normally consolidated
region. This is in good agreement with the results observed on the compression
index and the destructuring index of PC treated LR soils. In the over consolidated
region, the results showed that the constrained modulus increased gradually with
treatment as the stiffness of the soils also increase with treatment. Based on the
results, the effects of treatment on the constrained modulus can be schematically
expressed as Figure 6.79.
When treated with cement, the fabric of the soil changes. The pozzolanic
reaction of cement causes flocculation and aggregation of soil particles. As a
result, the particle size generally increases. The distribution of pore-size, both
macro and micro-pores, also changes, generally yielding a more open fabric with
treatment. Tremblay et al. (2001) and Al-Rawas (2002) observed increases in
particle size and pore size with treatment with lime and cement. Increases in the
particle size and pore size cause increase in hydraulic conductivity. The results
of consolidation tests performed in this experimental program showed that the
hydraulic conductivity of LR soil increased with treatment (Figure 6.54). Although
the results showed a continuous increase of the hydraulic conductivity of LR soil,
at least within the cement content range investigated, it is expected that the
hydraulic conductivity will decrease if the soil is treated with significant amounts
of cement. As shown in Figure 6.80, Locat et al. (1996) observed that the
hydraulic conductivity of Louiseville clay generally increased by 2 4 times when
treated with 1 2 % lime, but the hydraulic conductivity decreased with further
increase in the lime content.

319
The coefficient of consolidation of a soil is a function of the hydraulic conductivity
and the constrained modulus. As discussed previously, the constrained modulus
of the cement treated soil generally increases in the over consolidated region but
remained the same in the normally consolidation region. The hydraulic
conductivity generally increases with treatment at least up to a certain amount of
binding agent. Considering the increase in the hydraulic conductivity and
constrained modulus with treatment, it is expected that the coefficient of
consolidation will increase with treatment. From the perspective of mechanical
behavior of soil, an increase in the hydraulic conductivity indicates that the
excess pore pressure will dissipated quickly, and thus that primary consolidation
will be completed in a relatively shorter period of time. As shown in Figure 6.55,
the coefficient of consolidation of LR soil increased continuously with treatment.
This is mainly due to the continuous increase in the hydraulic conductivity with
treatment. Based on the results, the effect of treatment on the coefficient of
consolidation can be schematically expressed as in Figure 6.79.
The effect of treatment on the secondary consolidation behavior was discussed
in section 6.2.8. Figure 6.57 summarizes the effects of treatment on the
secondary compression behavior of PC treated LR soil. In its natural state, the
C/Cc of LR soil was in the range of 0.095 0.102 (see Chapter 5) in the upper
range of typical values for organic soils and peats. With increasing cement
content, the secondary compression gradually decreases. When treated with
100% PC, the C/Cc of LR soil decreased to typical values for granular soils. The
decrease in the C/Cc of a soil with treatment is due to (1) replacement of organic
soil particles, which that have high secondary compression index, with less creep
susceptible cement hydration products, (2) increase in particle size, and less
importantly (3) slight increase in the compression index with treatment.

320
6.5. Effects of treatment on the organic matter of LR soil
The results of pH measurements on PC treated LR soil show that the pH of the
soil started increasing substantially even with 8.0% PC, and reached the value of
12.6 when treated with 50% PC and more. The results of CRS and IL
consolidation tests indicate that the improvements of the consolidation behavior
are closely related to the change in pH of the soil. While the increase in pH up to
12.6 provides an environment in which hardening of the C3S2H3 gel can occur, it
also significantly changes the environment for the organic matter. Under highly
alkali condition, humic acid and fulvic acid may dissolve into pore water. This
chemical alteration of the organic matter can also cause improvement of the
mechanical properties of the soil by reducing its organic matter. However, back
calculation of the loss on ignition of PC treated LR soil showed that the value of
loss on ignition did not change significantly with treatment (Table 6.4).
To gain insight on the interaction between the organic matter and the PC,
chemical analyses were performed on the LR soil treated with PC. The objectives
of the chemical analyses were: (1) to evaluate the effect of treatment with PC on
the organic content of LR soil, and (2) to evaluate the change in chemistry of the
organic matter of LR soil with treatment. This part of the experimentation involved
fractionation of the humic substances of PC treated LR soil into three main
groups: humin and inorganic components of LR soil, humic acid (HA), and fulvic
acid (FA). The elemental compositions of the each of the three fractions of the
PC treated LR soil were then assessed to evaluate the changes due to treatment
with PC. Fourier-transform Infra Red (FT-IR) spectroscopy analyses were also
performed on each of the three fractions to evaluate the change in the functional
groups of LR soil.
Extraction and fractionation of the humic substances was performed on the LR
soil treated with 20, 50, and 100% PC. The samples were prepared following the
same procedure employed for preparation of the PC treated LR soil samples for

321
CRS and IL consolidation tests, as summarized in Chapter 3. Immediately after
mixing with cement, the samples were stored in a glass jar with an air-tight seal
to prevent any moisture loss during curing. The curing surcharge was not applied
to the samples used for FT-IR analysis. The samples were cured in the glass jar
for seven days. The extraction procedure employed for the PC treated LR soil
was slightly modified from the procedure described in Chapter 3. The pretreatment with 0.1M HCl-HF solution was not applied to prevent dissolution of
silicates (SiO2) that are part of the cementitious products. Instead, the samples
were pre-treated with 0.1M HCl solution to remove carbonates so that organic
matter could be easily separated from the mineral component of the soil. 0.1M
HCl solution was added to each soil sample with a ratio of solution to soil,
including both the LR soil and the PC, of 5:1. The samples were continuously
stirred in a mechanical shaker for 48 hours. After 48 hours of shaking, the
samples were centrifuged at 2000 rpm for 15 minutes to separate the solution
from the soil. The color of the separated solution was yellow to light brown. The
supernatant solution was carefully decanted. The same extraction procedure
employed for fractionation of untreated LR soil described in Chapter 3 was
employed afterward. Except for the process of decanting the supernatant after
the HCl treatment, the fractionation was kept in a closed system during the whole
procedure to minimize any loss of organic matter.
The masses of each fraction extracted from the PC treated LR soils are
summarized in Table 6.12. The average mass of each of the three fractions of
untreated LR soil that was pre-treated with HCl-HF solution are also shown for
comparison (See Chapter 3 for the details of the untreated LR soil). The non
extracted fraction of the untreated LR soil represents the humin and inorganic
part of the LR soil. The non extracted fraction of the PC treated LR soil
represents the humin, inorganic part of the LR soil, cementitious products, and
the humic and fulvic acid that were not extracted and were retained by cement.

322
The results show that the amount of humic and fulvic acid fractions extracted
from the soil decreased substantially with increasing PC content. The decrease
in the amount of the extracted humic acid was especially high. Only 3% of the
humic acid (0.106 g) of the untreated LR soil was extracted from the 20% PC soil,
and only 0.1% (0.003g) of the humic acid was extracted from the 50% PC soil.
For the100% PC soil, the amount of the extracted humic acid was too small and
the mass of the fraction was within the range accurately measured with the scale
(10-4 g). The amount of extracted fulvic acid also decreased with increasing PC
content, but the decrease was less substantial than that observed for humic acid.
From the LR soil treated with 20% PC, only 10% of the fulvic acid was extracted.
The amounts of extracted fulvic acid further decreased to 3.2% (0.042g) and
0.5% (about 0.013g) with increasing 50%PC and 100%PC, respectively. If the
humic and fulvic acids of LR soil were not bound to the cement but were
dissolved into the pore water in the highly basic environment, they should have
been extracted during the extraction procedure. Therefore, the results indicate
that the humic and fulvic acids of LR soil reacted well with cement and were
strongly bound to cement.
Elemental analyses were performed employing the LECO CHN-2000 analyzer
to evaluate the change in the composition of the humic and fulvic acids of the PC
treated LR soil. As shown in Table 6.12, the amounts of humic acid and fulvic
acid extracted from the LR soil treated with 50 and 100% PC were too small, so
that elemental analysis could not be performed on these fractions.
The results of the elemental analyses show that the chemical compositions of the
humic and fulvic acids extracted from the LR soil treated with 20% PC changed
slightly from the chemical composition of the total humic and fulvic acids of the
untreated LR soil. For example, the organic carbon contents of the humic and
fulvic acids extracted from the LR soil treated with 20% PC were approximately
4.5% and 3.9% smaller than the total humic and fulvic acids in the untreated LR

323
soil, respectively. On the other hand, the hydrogen and nitrogen contents of the
humic and fulvic acids from the LR soil treated with 20% PC were higher. This
may indicate that some functional groups have different reactivity with the
cement during hydration.
The percentages of organic C, H, and N of the non-extracted fraction gradually
decreased with increasing PC content. This trend can be misleading because the
total mass of solid, including both LR soil and PC, of the non-extracted fractions
increased with increasing PC%. Therefore, the masses of the organic C, H and N
had to be estimated by multiplying the percentage of each element by the mass
of each fraction. The masses of the organic C of each fraction are summarized in
Table 6.13. The organic content was estimated by multiplying the organic C
content by 1.72. The total organic content of the untreated LR soil and the LR soil
treated with 100% PC were in the range of 40 41% and were higher than the
other five samples. Except these two samples, the total organic content of the
samples was in the 34.9 37.4 % range, and showed good repeatability. The
difference may be related to the variation of the soil samples. The organic
contents determined from these tests were in good agreement with the loss on
ignition values. The average value of the loss on ignition of the LR soils from the
same bin (LR-B9) was 40.0 4.0 (Chapter 3, Table 3.2). The results indicate that
organic matter was not lost during the extraction and fractionation procedures,
which confirms that the whole procedure was kept in a closed system. The
results also show that approximately only 2.6% and 9.0% of organic carbon was
extracted as humic and fulvic acids when treated with 20% of cement compared
to the untreated LR soil.
The masses of H and N in each of the three fractions are summarized in Table
6.14. Similar to the organic carbon, the H and N extracted as humic and fulvic
acids decreased substantially when treated with 20% cement: about 3.5% of H

324
(0.006g) and 4.5% of N (0.006g) were extracted as humic acids, and about 23%
of H (0.0025g) and 40% of N (0.002g) were extracted as fulvic acid.
Fourier Transform Infra Red (FT-IR) spectroscopy analyses were performed on
each of the three fractions separated from the LR soil treated with PC. The
objectives of the FT-IR analysis were (1) to investigate the changes in chemical
composition of the humic and fulvic acids extracted from the PC treated LR soil;
and (2) to evaluate the reactivity of the functional groups present in LR soil. The
KBr pellets for the FT-IR analysis were prepared with approximately 2mg of each
fraction. The FT-IR analysis could not be performed on the humic acid fraction of
the LR soil treated with 100% PC. Two analyses were performed on each
fraction for consistency of results. Although the frequencies of various types of
vibration modes, symmetric or asymmetric stretching and bending vibration
modes, of organic functional groups are well known (see Chapter 3), the
similarity of the chemical composition of some organic functional groups makes it
hard to accurately understand the sources responsible for changes in FT-IR
spectra. Thus, the shifts or changes in the intensity of bands of the FR-IR spectra
were evaluated qualitatively.
Figure 6.82 shows the FT-IR spectra of the non-extracted fraction of the PC
treated LR soil. The most significant change in this fraction is the gradual
increase in the intensity of the broad band in the 3000 3600 cm-1 region. This
region represents the symmetric and asymmetric stretching of O-H. According to
Mollah et al. (1995), the increase in the intensity of this band indicates the
formation of the Ca(OH)2 and the C-S-H gel. Note that since the samples were
completely dried by freeze-drying, the band is not due the O-H in water
molecules.
The sharp increases in the intensity of the bands in the 1415 cm-1 and 872 cm-1
regions are due to the formation of carbonate (CO3) by a process called

325
carbonation. The carbonate is a by-product of the cement hydration and does not
indicate the hydration of cement. The calcium ions released during hydration of
cement react with CO2 in the pore water or in the atmosphere to form calcium
carbonate (CaCO3).
The band in the 1576 1772 cm-1 region gradually increases with respect to the
base line in the 2000 2500 cm-1 region. Among the changes in the intensity of
the band observed in the non-extracted PC treated LR soil fractions, only this
band represents the organic functional groups. The organic functional groups
assigned to this band include: (1) C=O stretch of COOH (1720 cm-1), (2) C=O
stretching of amide groups (1630 1660 cm-1), (3) aromatic C=C stretch and/or
asymmetric COO- stretch (1610 cm-1) and (4) COO- symmetric stretching, N-H
deformation, +C=N stretching (1590 1517 cm-1). Note that most of the organic
functional groups listed here are mainly carbon-containing organic functional
groups, with a few containing H and N. The increase in the C, H, N contents in
the non-extracted PC treated LR soil may be due to the retention of these
functional groups by cementation.
Figure 6.83 presents the FT-IR spectra of the humic acid fraction extracted from
the untreated and the PC treated LR soil. The relatively high intensity of the band
at 3300 3400cm-1 region in the humic acid of untreated LR soil indicates the
presence of phenolic OH (OH stretch of phenolic OH). The increase in the
intensity of the broad band in the 3300 3600 cm-1 region may also indicate that
part of the Ca(OH)2 and the C-S-H gel were dissolved and extracted under highly
acidic environment. Note that the pH of the dissolved humic and fulvic acids were
reduced to 1.0 to precipitate humic acid. Under very acidic environment, the
cementitious gel can be dissolved into pore water. The intensities of the sharp
bands in 2850 2970 cm-1 did not change. These bands represent symmetric
CH stretch of CH2- (2840 cm-1) and asymmetric CH stretch of CH2- (2930 cm1

). The small peak in the 1718 cm-1 region present in the untreated LR soil is

326
clouded by the increase in the intensities of the bands in 1400 and 1546 - 1600
cm-1. These peaks indicate presence of carboxyl group (1600cm-1) and phenolic
OH (1400cm-1) in the humic fraction of LR soil treated with PC. The -C=O stretch
of COOH occurs at 1720 cm-1, and the symmetric and antisymmetric stretching
of COO- occurs at 1517 1590 cm-1 and 1390 1400 cm-1, respectively. The
increase in the intensity of the peak at 1400cm-1, which is related to phenolic OH,
with treatment indicates that the phenolic OH of humic acid does not react well
with cement. As mentioned in section 2.6, phenol is known to retard cement
hydration, and does not react well with cement. Pollard et al (1991) also
observed that phenol was readily leached from cement treated waste materials.
The increase in the intensity of the peaks in the 1600 - 1546cm-1 region, which
are related to carboxyl group, indicates that the content of carboxyl acid in the
extracted humic fraction increases with treatment. The results imply that
carboxyl acid has a relatively lower reactivity with cement than other organic
acids present in the humic acid of LR soil.
Figure 6.84 presents the FT-IR spectra of the fulvic acid fraction extracted from
the untreated and the PC treated LR soil. The spectra pattern in the 2500 4000
cm-1 region of the fulvic acid of the untreated LR soil is similar to that of the humic
acid of the untreated LR soil. The strong band in 3390 cm-1 and the peak at 1400
cm-1 region in the untreated LR soil indicates the presence of phenolic OH.
Although the intensity of the peak at 1400 cm-1 in humic acid fraction increases
with treatment, the same peak in fulvic acid decreased slightly with treatment.
This may is due to experimental error, as the amount of the extracted fulvic acid
of the PC treated specimens is small. The intensities of the peaks at 1630, 1560,
and 1250 cm-1 regions, which indicate carboxyl acid, remain the same or slightly
increase with treatment. The same trend was observed in the extracted humic
acid fraction of the PC treated LR soil. The results indicate that the content of
the carboxyl acid in the extracted fulvic acid fraction of the PC treated LR soil
slightly increases with treatment. However, the change in the intensities of the

327
peaks related to carboxyl acid and phenolic OH of the extracted fulvic acid are
not as high as the change in the intensities of the same peaks of the extracted
humic acid. The results of the FT-IR analysis on the fulvic acid and the small
amount of the fulvic acid in LR soil indicate that the role of the fulvic acid in the
chemical reaction with cement is not as significant as the humic acid.
Based on the results of the LECO and FT-IR analyses, the following conclusions
on the role played by organic matter and the interactions of organic matter and
cement in LR soil can be derived: (1) organic matter does not significantly
interfere with cement hydration, (2) the humic and fulvic acids have high
reactivity with cement and are retained well by cementitious products, (3) humic
acid has a higher reactivity with cement than fulvic acid, and (4) phenolic OH and
carboxyl acids, that are known to act as water reducers/retarders, have relatively
lower reactivity with cement than other organic acids present in organic soils.
This observation is consistent with the slight increase in hydrogen content in the
extracted humic and fulvic acids fractions of cement treated LR soil obtained
from LECO tests. However, the significant decreases in the masses of the
extracted humic and fulvic acids of LR soil with cement treatment may suggest
that phenolic OH and carboxyl acid do not significantly interfere with cement
hydration.
As discussed in section 2.6, organic soils with high organic content (generally
greater than 75%) show less improvement in the compression behavior when
chemically treated. The results of this research suggest that this may be due to
the fact that a soil with higher organic content generally has low pH, and thus
more cement is required to bring the pH of the pore solution to 12.6. However,
more importantly, the content of H in the organic matter of a soil can be a good
indicator of the anticipated degree of improvement as the H-containing organic
groups generally have low reactivity with cement. The C-containing organic
functional groups, especially carboxyl group, have high reactivity with cement.

328
As discussed in Chapter 3, organic soils with higher degree of decomposition
have higher C/N content, which indicates that carbon content increases with
degree of decomposition. Thus, organic soils with higher degree of
decomposition may show higher improvement than organic soils with lower
degree of decomposition.

329
Table 6.1: Comparison of %PC vs. Cement factor
%PC (by dry mass)

Cement factor (kg/m3)


per 1m3 of untreated LR soil per 1m3 of treated LR soil

23.68

23.15

20

59.21

56.01

50

148.02

129.52

100

296.05

230.26

Table 6.2: Summary of the PC treated LR soil specimens


%PC

LOI

(by dry mass)

(%)

(%)

LR-B9

18.71

34.44

166.85

3.737

CRS043

LR-B9

18.71

33.81

169.01

3.650

IL048

LR-B10

8.08

39.29

180.17

3.945

CRS049

LR-B10

8.08

39.87

174.44

3.746

IL050

LR-B10

51.42

28.69

167.98

3.572

CRS051

LR-B10

51.42

29.42

167.58

3.572

IL054

LR-B10

103.43

22.03

125.97

2.580

CRS055

LR-B10

103.43

21.15

125.20

2.572

CRS064

LR-B11

100.80

22.59

129.00

2.694

Test No.

Bin#

IL042

eo

330
Table 6.3: Test process of CRS consolidation tests on PC treated LR soil
Test #

Strain

Unloading

1st creep

Creep

2nd creep

Creep

Final

rate,

rate factor

stress

time

stress

time

stress

(kPa)

(hrs)

(kPa)

(hrs)

(kPa)

(%/hr)
CRS043

0.5

1.0

1003.9

72

106.3

72

1598.5

CRS049

0.5

1.0

1010.3

72

110.6

72

2021.6

CRS051

0.5

1.0

1003.6

72

102.7

72

2014.6

CRS055

0. 5

1.0

4008.1

72

406.1

6322.4

CRS064

0.5

1.0

3066.8

72

307.2

72

6369.0

Table 6.4: Comparison of the back calculated LOI and the actual LOI
Test No.

%PC
(by dry mass)

LOI Back-calculated (%)

LOI Actual (%)

Difference (%)

IL042

18.71

40.67

n/a

n/a

CRS043

18.71

40.89

n/a

n/a

IL048

8.08

42.36

41.68

0.68

CRS049

8.08

43.12

41.68

1.44

IL050

51.42

44.18

44.73

0.55

CRS051

51.42

43.87

44.73

0.86

IL054

103.43

44.71

44.84

0.13

CRS055

103.43

43.34

44.84

1.50

CRS064

100.80

45.20

48.04

2.84

331
Table 6.5: Summary of the results of consolidation tests performed on PC treated
LR soil specimens
Test
#

%PC

p
(kPa)

Cc
2-4p

Cs

4-8p

Cr

8-

1-

2-

4-

1-

16p

2OCR

4OCR

2OCR

2OCR

Ck/eo

Ck/Cc

1/4.45

1/1.56

1/4.01

1/1.46

1/4.58

1/1.58

1/4.11

1/1.56

1/3.68

1/1.69

042

18.71

94.40

1.326

1.323

1.294

043

18.71

94.83

1.327

1.339

1.323

048

8.08

70.21

1.398

1.409

1.250

049

8.08

79.43

1.489

1.446

1.310

050

51.42

404.44

1.720

051

51.42

402.06

1.677

0.017

0.026

0.013

0.048

1/3.94

1/1.57

054

103.43

n/a

055

103.43

1249.8

1.376

0.007

0.005

0.005

0.010

1/3.41

1/1.83

064

100.80

1129.7

1.388

0.006

0.005

0.004

0.009

1/3.71

1/1.91

0.028
0.041

0.042
0.066

0.026
0.057

0.084
0.139

Table 6.6: Summary of IL consolidation test performed on 8.00% PC treated LR


soil (IL048)
v
(kPa)

eEOP

tEOP

uhmax

uhmax/v

tuhmax

tf

Cv

(min)

(kPa)

(%)

(min)

(min)

(cm2/sec)

(cm/sec)

Cc

1.55

3.930

6.32

3.926

6.8

3.7

77.6

0.18

9.0

3.08E-02

0.007

12.1

3.913

22.0

4.6

79.6

0.23

27.0

4.12E-03

0.045

23.9

3.865

30.3

9.9

83.9

0.10

35.1

4.58E-03

3.65E-07

0.160

48.6

3.656

254.6

19.7

79.8

0.08

260.4

7.53E-04

1.25E-07

0.672

98.3

3.301

355.7

49.6

99.8

0.23

380.8

3.30E-04

4.70E-08

1.170

196.1

2.891

383.0

94.6

96.7

0.57

500.6

2.51E-04

2.09E-08

1.368

393.5

2.468

437.7

188.4

95.4

0.40

480.6

1.86E-04

7.91E-09

1.400

786.9

2.044

542.1

377.6

96.0

1.0

600.6

1.21E-04

2.58E-09

1.407

1572.7

1.668

605.1

719.6

91.6

1.3

23625.7

8.43E-05

8.00E-10

1.251

332
Table 6.7: Summary of IL consolidation test performed on 18.7% PC treated LR
soil (IL 042)
v
(kPa)

eEOP

tEOP

uhmax

uhmax/v

tuhmax

tf

Cv

(min)

(kPa)

(%)

(min)

(min)

(cm2/sec)

(cm/sec)

Cc

0.21

3.723

5.87

3.683

14.2

5.67

100.0

0.28

20.1

1.02E-02

0.028

11.9

3.668

6.4

4.19

69.5

0.20

17.0

1.80E-02

9.24E-07

0.049

24.1

3.640

10.9

8.98

73.6

0.15

17.0

1.48E-02

6.86E-07

0.089

48.9

3.577

22.2

19.6

79.0

0.27

41.1

7.33E-03

3.87E-07

0.205

98.3

3.325

78.0

42.2

85.4

0.17

90.2

2.02E-03

2.13E-07

0.831

196.4

2.987

79.4

88.4

90.1

0.25

100.2

1.60E-03

1.14E-07

1.125

394.1

2.586

102.2

176.4

89.2

0.08

110.2

1.07E-03

4.50E-08

1.325

785.2

2.190

109.1

352.4

90.1

0.21

140.4

7.45E-04

1.56E-08

1.323

1557.8

1.805

153.1

694.8

89.9

0.38

26808

4.47E-04

4.49E-09

1.271

Table 6.8: Summary of IL consolidation test performed on 51.4% PC treated LR


soil (IL 050)
v
(kPa)

eEOP

tEOP

uhmax

uhmax/v

tuhmax

tf

Cv

(min)

(kPa)

(%)

(min)

(min)

(cm2/sec)

(cm/sec)

Cc

1.28

3.558

5.62

3.558

0.20

1.62

37.3

0.08

5.0

3.33E-01

12.4

3.562

0.52

2.00

29.5

0.10

6.0

4.94E-02

24.6

3.559

1.43

8.01

65.7

0.08

5.0

1.11E-01

6.43E-07

0.011

49.3

3.551

1.47

25.4

100.0

0.08

5.0

9.50E-02

6.51E-07

0.026

97.7

3.542

0.48

34.6

71.5

0.07

4.0

1.47E-01

5.71E-07

0.029

199.1

3.525

0.50

38.1

37.6

0.05

6.0

1.88E-01

6.68E-07

0.054

395.8

3.350

17.52

109.0

55.4

0.22

19.0

1.43E-02

2.75E-08

0.589

788.9

2.928

28.68

322.8

82.1

0.12

30.0

5.41E-03

1.24E-08

1.407

1574.2

2.422

38.72

675.4

86.0

0.12

31701

2.80E-03

3.87E-08

1.686

333
Table 6.9: Summary of IL consolidation test performed on 103.4% PC treated LR
soil (IL 054)
v
(kPa)

eEOP

tEOP

uhmax

uhmax/v

tuhmax

tf

Cv

(min)

(kPa)

(%)

(min)

(min)

(cm2/sec)

(cm/sec)

Cc

3.5

2.569

5.6

4.0

10.6

5.0

22.2

6.0

43.8

4.0

86.5

2.565

0.20

1.8

4.2

0.07

9.0

2.14E-01

1.05E-06

174.2

2.548

0.27

32.1

36.6

0.05

6.0

4.26E-01

2.19E-06

0.054

348.6

2.537

0.35

33.5

19.2

0.05

5.0

4.23E-01

7.42E-07

0.037

698.1

2.513

2.43

65.0

18.6

0.07

6.0

1.40E-01

2.58E-07

0.078

1387.1

2.354

24.5

332.7

48.3

0.5

22942

1.22E-02

7.71E-08

0.535

Table 6.10: LR soil specimens employed for evaluation of curing time effects
Test No.

Bin#

Test

Curing Time

type

(days)

%PC

LOI*

Gs*

wi (%)

eo

(%)

CRS059

LR-B10

IL

28

52.59

44.82

2.104

168.0

3.645

CRS060

LR-B10

CRS

28

52.59

45.03

2.102

171.1

3.652

CRS065

LR-B11

CRS

28

100.8

42.68

2.127

130.1

2.732

IL066

LR-B11

IL

28

100.8

42.34

2.130

133.0

2.760

334
Table 6.11: LR soil specimens employed for evaluation of the effects of curing
surcharge
Test No.

Bin#

Test

Curing

%PC

type

Surcharge

LOI*

Gs*

wi

(%)

eo

(%)

(kPa)
CRS057

LR-B10

IL

96

52.59

43.08

2.122

158.6

3.398

CRS058

LR-B10

CRS

96

52.59

44.23

2.110

157.7

3.354

CRS062

LR-B11

CRS

192

48.35

44.65

2.106

137.1

3.014

CRS063

LR-B11

IL

192

48.35

44.64

2.106

143.6

3.112

Table 6.12: Mass and organic C, H, O contents of each fraction of the humic
substances extracted from PC treated LR soil (LR-B9)
Extracted fractions

Non-extracted fraction
%PC

Humic acid (HA)

Fulvic acid (FA)

HA/FA

Mass

Mass

Mass

(g)

(%)

(%)

(%)

(g)

(%)

(%)

(%)

(g)

(%)

(%)

(%)

11.435

14.0

3.1

1.0

3.660

49.9

4.7

3.6

0.420

41.2

2.5

1.1

8.71

20

17.853

16.7

2.7

1.1

0.106

45.4

5.6

5.2

0.041

34.8

5.6

4.7

2.58

20

17.850

17.4

2.8

1.1

0.107

45.4

5.5

5.2

0.043

39.9

6.3

5.5

2.51

50

22.490

13.9

2.4

0.8

0.002

0.008

0.22

50

22.478

14.5

2.5

0.9

0.004

0.019

0.20

100

29.998

10.7

2.1

0.5

0.000

0.002

0.0

100

29.998

11.8

2.2

0.7

0.000

0.002

0.0

335
Table 6.13: Mass of the organic carbon in each fraction of LR soil
Total

%PC

Non-extracted

Extracted fractions

fraction

Humic acid (HA)

Fulvic acid (FA)

C (g)

OC (%)

C (g)

C (g)

C (g)

3.600

41.28

1.601

1.826

0.173

20

3.044

34.90

2.981

0.048

0.014

20

3.172

36.37

3.106

0.049

0.017

50

3.126

35.85

3.126

50

3.259

37.37

3.259

100

3.210

36.81

3.210

100

3.540

40.59

3.540

Table 6.14: Mass of H and N in each fraction of LR soil


Non-extracted

Total

%PC

fraction

Extracted fractions
Humic acid (HA)

Fulvic acid (FA)

H (g)

N (g)

H (g)

N (g)

H (g)

N (g)

H (g)

N (g)

0.537

0.251

0.354

0.114

0.172

0.132

0.011

0.005

20

0.490

0.204

0.482

0.196

0.006

0.006

0.002

0.002

20

0.508

0.204

0.500

0.196

0.006

0.006

0.003

0.002

50

0.540

0.180

0.540

0.180

50

0.562

0.202

0.562

0.202

100

0.630

0.150

0.630

0.150

100

0.660

0.210

0.660

0.210

336

Va

AIR

Vw

Water

Vv
Mw

VT
Vls

LR soil

Mls

Vs

MT
Ms

Vc

Hydrated Cement

Mc

Figure 6.1: Four phases of PC treated LR soil (not drawn to scale)

4.5

Initial void ratio, eo

4.0

3.5

3.0

2.5

2.0

20

40

60

80

100

PC content (%PC)

Figure 6.2: The initial void ratio of PC treated LR soil

337

4.0
IL048
CRS049

3.5

Void ratio

3.0

2.5

2.0
CRS046[RC]
CRS049[8.00%PC]
IL048[8.00%PC]

1.5
1

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 6.3: Compression curves of LR soil treated with 8.00%PC

1.5

1.0

'p Block

Void Index, Iv

0.5
SCL
0.0

-0.5
CRS046[RC]
CRS049[8.00%PC]
IL048[8.00%PC]
CRS061[Block]

-1.0

-1.5

10

ICL
100

1000

Vertical effective stress, 'v (kPa)

Figure 6.4: Normalized compression curves of LR soil treated with 8.00% PC

Constrained modulus, D = 1/mv (kPa)

338

CRS046[RC]
CRS049[8.00%PC]
IL048[8.00%PC]

10000

1000

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 6.5: Constrained modulus of LR soil treated with 8.00% PC

Excess pore pressures/Actual strain rate, uh/a

200
CRS046[RC]
CRS049[8.00%PC]
150

100

50

0
4.0

3.5

3.0

2.5

2.0

1.5

Void ratio

Figure 6.6: Cumulative excess pore pressure generated in CRS test on LR soil
treated with 8.00% PC

339

4.0
CRS046[RC]
CRS049[8.00%PC]
IL048[8.00%PC]

3.5

Void ratio

3.0

2.5

2.0

1.5
1e-9

1e-8

1e-7

1e-6

Hydarulic conductivity, k (cm/sec)

Figure 6.7: Hydraulic conductivity of LR soil treated with 8.00% PC

CRS046[RC]
CRS049[8.00%PC]
IL048[8.00%PC]

Coefficient of consolidation, Cv (cm /sec)

1e-1

1e-2

1e-3

1e-4
1

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 6.8: Coefficient of consolidation of LR soil treated with 8.00% PC

340

2.1

800

Void ratio

700

1.9

600

1.8

500

1.7

400

1.6

300

1.5

200

8% PC treated LR soil
(IL048)
'v = 1572 kPa

1.4
1.3
1.2
1e-2

1e-1

1e+0

EOPExcess pore pressure

100

Excess Pore Pressure (kPa)

Time settlement
Excess Pore Pressure

2.0

1e+1

1e+2

1e+3

1e+4

-100
1e+5

Elapsed time (min)

Figure 6.9: Time settlement and excess pore pressures dissipation curves from
long term creep test performed on LR soil treated with 8.00% PC

4.0

'p

Void ratio

3.5

3.0

2.5

2.0

1.5

CRS046[RC]
CRS043[18.7%PC]
IL042[18.7%PC]
1

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 6.10: Compression curves of LR soil treated with 18.7% PC

341

1.5

1.0

Void Index, Iv

0.5
SCL

0.0

-0.5
CRS046[RC]
CRS043[18.7%PC]
IL042[18.7%PC]
CRS061[Block]

-1.0

-1.5

10

ICL

100

1000

Vertical effective stress, 'v (kPa)

Constrained modulus, D = 1/mv (kPa)

Figure 6.11: Normalized compression curves of LR soil treated with 18.7% PC

CRS046[RC]
CRS043[18.7%PC]
IL042[18.7%PC]
10000

1000

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 6.12: Constrained modulus of LR soil treated with 18.7% PC

342

Excess pore pressures/Actual strain rate, uh/a

200
CRS046[RC]
CRS043[18.7%PC]
150

100

50

0
4.0

3.5

3.0

2.5

2.0

1.5

Void ratio

Figure 6.13: Excess pore pressure generated in CRS test on LR soil treated with
18.7% PC

4.0
CRS046[RC]
CRS043[18.7%PC]
IL042[18.7%PC]

Void ratio

3.5

3.0

2.5

2.0

1.5

1e-9

1e-8

1e-7

1e-6

Hydarulic conductivity, k (cm/sec)

Figure 6.14: Hydraulic conductivity of LR soil treated with 18.7% PC

343

CRS046[RC]
CRS043[18.7%PC]
IL042[18.7%PC]

Coefficient of consolidation, Cv (cm /sec)

1e-1

1e-2

1e-3

1e-4
1

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 6.15: Coefficient of consolidation of LR soil treated with 18.7% PC

2.2

800

Void ratio

700

2.0

600

1.9

500

1.8

400

1.7

300

1.6

200

1.5
1.4
1.3
1e-2

18.7% PC treated LR soil


(IL042)
'v = 1578 kPa
1e-1

1e+0

100

Excess Pore Pressure (kPa)

Time settlement
Excess Pore Pressure

2.1

EOPExcess pore pressure

1e+1

1e+2

1e+3

1e+4

-100
1e+5

Elapsed time (min)

Figure 6.16: Time settlement and excess pore pressures dissipation curves from
a creep test performed on LR soil treated with 18.7% PC

344

3.7
3.5

'p

3.3

Void ratio

3.1
2.9
2.7
2.5
CRS046[RC]
CRS051[51.4%PC]
IL050[51.4%PC]

2.3
2.1

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 6.17: Compression curves of LR soil treated with 51.4% PC

1.5
SCL
1.0

Void Index, Iv

0.5

'p 51.4% PC

0.0

-0.5
CRS046[RC]
CRS051[51.4%PC]
IL050[51.4%PC]
CRS061[Block]

-1.0

-1.5

10

ICL

100

1000

Vertical effective stress, 'v (kPa)

Figure 6.18: Normalized compression curves of LR soil treated with 51.4% PC

345

3.7
3.5

'p = 402.66 kPa

3.3

Void ratio

3.1
2.9

'p Estimated = 1254 kPa


'p Measured = 1361.5 kPa

2.7
2.5
2.3
2.1

CRS051[51.4%PC]
Preconsolidation pressure

1000

Vertical effective stress, 'v (kPa)

Figure 6.19: Development of preconsolidation pressure during creep in LR soil


treated with 51.4% PC

346

3.7
3.5

+15.6 days

3.3

Void ratio

3.1
2.9
+17.3 days

2.7

+24.5 days
+21.5 days

2.5
CRS046[RC]
CRS051[51.4%PC]

2.3
2.1

+20.3 days

+25.4 days

10

100

1000

Vertical effective stress, 'v (kPa)

Constrained modulus, D = 1/mv (kPa)

Figure 6.20: Time elapsed since treatment with PC at each stage of CRS
consolidation test

CRS046[RC]
CRS051[51.4%PC]
IL050[51.4%PC]
10000

1000

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 6.21: Constrained modulus of LR soil treated with 51.4% PC

347

Excess pore pressures/Actual strain rate, uh/a

50
CRS046[RC]
CRS051[51.4%PC]

40

30

20

10

0
3.8

3.6

3.4

3.2

3.0

2.8

2.6

Void ratio

Figure 6.22: Excess pore pressures generated in CRS test on LR soil treated
with 51.4% PC

4.0
CRS046[RC]
CRS051[51.4%PC]
IL050[51.4%PC]

Void ratio

3.5

3.0

2.5

2.0

1.5

1e-9

1e-8

1e-7

1e-6

Hydarulic conductivity, k (cm/sec)

Figure 6.23: Hydraulic conductivity of LR soil treated with 51.4% PC

348

Coefficient of consolidation, Cv (cm2/sec)

1e+1
CRS046[RC]
CRS051[51.4%PC]
IL050[51.4%PC]

1e+0

1e-1

1e-2

1e-3

1e-4
1

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 6.24: Coefficient of consolidation of LR soil treated with 51.4% PC

3.0

800

Void ratio

700

2.8

600

2.7

500

2.6

400

2.5

300

2.4

200

2.3
2.2
2.1
1e-2

Treatment with 51.4% PC


(IL050)
'v = 1574 kPa
1e-1

1e+0

1e+1

100
EOPExcess pore pressure

1e+2

1e+3

Excess Pore Pressure (kPa)

Time settlement
Excess Pore Pressure

2.9

1e+4

-100
1e+5

Elapsed time (min)

Figure 6.25: Time settlement and excess pore pressure dissipation curves from a
creep test performed on LR soil treated with 51.4% PC

349

3.9

'v = 257 kPa

3.5

'v = 850 kPa

Void ratio

3.1

'v = 1000 kPa


2.7

'v = 2450 kPa

2.3
CRS068[49.1%PC]
CRS069[49.1%PC]

1.9

1.5

100

1000

Vertical effective stress, 'v (kPa)

Figure 6.26: Consolidation curves of LR soil treated with 49.1% PC

3.1

2.9

'p Estimated = 1290 kPa


'p Measured = 1432.0 kPa

Void ratio

2.7

2.5

2.3

2.1

1.9

CRS068[49.1%PC]

1000

3000

Vertical effective stress, 'v (kPa)

Figure 6.27: Increase in the preconsolidation pressure after 7 day creep at 1000
kPa

350

3.7

'p Estimated = 330 kPa


'p Measured = 382 kPa

Void ratio

3.5

3.3

3.1
CRS068[49.1%PC,48/14]
CRS069[49.1%PC,48/14]

2.9

100

1000

Vertical effective stress, 'v (kPa)

Figure 6.28: Increase in the preconsolidation pressure after 7 day creep at 257
kPa

351

3.3

3.1

'p Estimated = 1050 kPa


'p Measured = 1200 kPa

Void ratio

2.9

2.7

2.5

2.3

2.1

CRS069[49.1%PC]

1000

3000

Vertical effective stress, 'v (kPa)

Figure 6.29: Increase in the preconsolidation pressure after 7 day creep at


850kPa

352

2.3

'p Estimated = 3200 kPa


'p Measured = 3870 kPa

Void ratio

2.1

1.9

1.7
CRS069[49.1%PC,48/14]

1.5
2000

8000

Vertical effective stress, 'v (kPa)

Figure 6.30: Increase in the preconsolidation pressure after 7 day creep at 2450
kPa

2.9

'p CRS064

Void ratio

2.5

'p CRS055

2.1

CRS046[RC]
CRS055[103.4%PC]
IL054[103.4%PC]
CRS064[100.8%PC]

1.7

1.3

10

100

1000

10000

Vertical effective stress, 'v (kPa)

Figure 6.31: Compression curves of LR soil treated with 100% PC

353

2.6

'p = 1249.8 kPa

Void ratio

2.2

1.8

'p Estimated = 5127.7 kPa


'p Measured = 5705.8 kPa
CRS055[103.4%PC]
Preconsolidation pressure

1.4

1000

10000

Vertical effective stress, 'v (kPa)

Figure 6.32: Development of preconsolidation pressure during creep in LR soil


treated with 103.4 % PC

354

2.6

'p = 1129.7 kPa

Void ratio

2.2

'p Estimated = 3811.8 kPa


'p Measured = 4403.3 kPa

1.8

CRS064[100.8%PC]
Preconsolidation pressure
1.4

1000

10000

Vertical effective stress, 'v (kPa)

Figure 6.33: Development of preconsolidation pressure during creep in LR soil


treated with 100.8 % PC

1.5

1.0

Void Index, Iv

0.5

SCL

0.0

'p CRS064
'p CRS055

-0.5
CRS046[RC]
CRS055[103.4%PC]
IL054[103.4%PC]
CRS064[100.8%PC]
CRS061[Block]

-1.0

-1.5

10

ICL
100

1000

10000

Vertical effective stress, 'v (kPa)

Figure 6.34: Normalized compression curves of LR soil treated with 100% PC

355

Constrained modulus, D = 1/mv (kPa)

1e+5
CRS046[RC]
CRS055[103.4%PC]
IL054[103.4%PC]
CRS064[100.8%PC]

1e+4

1e+3

10

100

1000

10000

Vertical effective stress, 'v (kPa)

Figure 6.35: Constrained modulus of LR soil treated with 100% PC

Excess pore pressures/Actual strain rate, uh/a

25

20

CRS055[103.4%PC]
CRS064[100.8%PC]

15

10

0
3.0

2.8

2.6

2.4

2.2

2.0

1.8

1.6

Void ratio

Figure 6.36: Cumulative excess pore pressure generated in LR soil treated with
100% PC

356

4.0

3.5

Void ratio

3.0

2.5

2.0
CRS046[RC]
CRS055[103.4%PC]
IL054[103.4%PC]
CRS064[100.8%PC]

1.5

1.0

1e-9

1e-8

1e-7

1e-6

1e-5

1e-4

Hydarulic conductivity, k (cm/sec)

Figure 6.37: Hydraulic conductivity of LR soil treated with 100% PC

Coefficient of consolidation, Cv (cm2/sec)

1e+1
CRS046[RC]
CRS055[103.4%PC]
IL054[103.4%PC]
CRS064[100.8%PC]

1e+0

1e-1

1e-2

1e-3

1e-4
1

10

100

1000

10000

Vertical effective stress, 'v (kPa)

Figure 6.38: Coefficient of consolidation of LR soil treated with 100% PC

357

2.6

400

2.5

300

2.4

200

2.3

100

Treatment with 103.4% PC


(IL054)
'v = 1387 kPa

2.2

2.1
1e-2

1e-1

1e+0

1e+1

EOPExcess pore pressure

Excess Pore Pressure (kPa)

Void ratio

Time settlement
Excess Pore Pressure

1e+2

1e+3

1e+4

-100
1e+5

Elapsed time (min)

Figure 6.39: Time settlement and excess pore pressures dissipation curves from
a creep test performed on LR soil treated with 1003.4% PC

14
12.6
12

pH

10

6
1 hr after mixing
24 hr after mixing
4

20

40

60

80

Cement content (%PC by dry mass of soil)

Figure 6.40: Change in pH of LR soil with treatment

100

358

4.0
ICL

CRS046[RC]
CRS049[8.00%PC]

Void ratio

3.5

3.0
Destructuring Index (b) = 0
2.5

2.0

1.5
10

100

1000

Vertical effective stress, 'v (kPa)

Figure 6.41: Quantification of post-yield destructuring of LR soil treated with


8.00% PC

359

4.0
ICL

CRS046[RC]
CRS043[18.7%PC]

Void ratio

3.5

3.0

Destructuring Index (b) = 0

2.5

2.0

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 6.42: Quantification of post-yield destructuring of LR soil treated with


18.7% PC

360

3.7
ICL
3.5
3.3

Void ratio

3.1

b = 0.07

2.9
2.7
2.5
2.3
2.1
10

b = 0.4
CRS046[RC]
CRS051[51.4%PC]

100

1000

Vertical effective stress, 'v (kPa)

Figure 6.43: Quantification of post-yield destructuring of LR soil treated with


51.4% PC:

361

2.7
ICL
2.5

Void ratio

2.3
2.1

b = 0.13

1.9
1.7
1.5
1.3
100

b = 0.6

CRS046[RC]
CRS055[103.4%PC]

1000

10000

Vertical effective stress, 'v (kPa)

Figure 6.44: Quantification of post-yield destructuring of LR soil treated with


103.4% PC

362

2.7
2.5

ICL

Void ratio

2.3

b = 0.11

2.1
1.9
1.7

b = 0.5
1.5
1.3
100

CRS046[RC]
CRS064[100.8%PC]

1000

10000

Vertical effective stress, 'v (kPa)

Figure 6.45: Quantification of post-yield destructuring of LR soil treated with


100.8% PC

363

2% lime

16
Void ratio, e

5% lime
10% lime

12
8
4
0% lime

0.0001 0.001 0.01 0.1


1
10
100
Vertical effective stress, v kPa

10000

Soil

ei

Preconsolidation
pressure (kPa)

Reference

Soft Louiseville clay


Lime content:
10%
5%
2%

8.5
9
9.5

24.5
4.2
0.37

0.43
0.34
3

Locat et al. (1996)

Figure 6.46: Compression curves and destructuring indices of lime treated


Louiseville Clay (after Liu and Carter, 2000)

364

'p RC

3.7

'p 18.7%PC

3.3

Void ratio

'p 8.00%PC

'p 51.4%PC

2.9

2.5

'p 103.4%PC

2.1

CRS046[RC]
CRS049[8.00%PC]
CRS043[18.7%PC]
CRS051[51.4%PC]
CRS055[103.4%PC]

1.7

1.3

10

100

1000

10000

Vertical effective stress, 'v (kPa)

Figure 6.47: Effect of treatment on compression behavior of LR soil

365

10

Vertical effective stress, v kPa


10
100

8
Void ratio, e

7
6
5
4
3

Curing stress 8 kPa

1000

% lime
0
2
5
10

2
1
0
Figure 6.48: Compression curves of lime stabilized Louiseville clay (after Locat et
al, 1996)

366

5.5

3.5
3.0
2.5
2.0

3.0
2.5
2.0
1.5

1.0
0.5

1.0
0.5

5.5

(b) 10% cement content

3.5

1.5

10
100
1000
10000
Vertical effective stress, v kPa

10
100
1000
10000
Vertical effective stress, v kPa

(c) 15% cement content

4.5
4.0
Void ratio, e

4.5
4.0
Void ratio, e

Void ratio, e

4.5
4.0

5.5

(a) 5% cement content

Symbol

Remolding
water content (%)
200
160
130
100
Undisturbed
Intrinsic
Generated post
yield compression line

3.5
3.0
2.5
2.0
1.5
1.0
0.5
10
100
1000
10000
Vertical effective stress, v kPa

Figure 6.49: Compression curves of cement treated soft Bangkok clay: (a) 5%,
(b) 10%, and (c) 15% cement contents (after Lorenzo and Bergado, 2004)

367

Cement factor (kg/m3)


1400

56

106

152

193

230

265

20

40

60

80

100

120

Preconsolidation stress, 'p (kPa)

1200
1000
800
600
400
200
0

Cement content (%PC)

Figure 6.50: Increase in the preconsolidation pressure with treatment

368

1.5
CRS046[RC]
CRS049[8.00%PC]
CRS043[18.7%PC]
CRS051[51.4%PC]
CRS055[103.4%PC]
CRS061[Block]

Void index (Iv)

1.0

0.5

0.0

-0.5

SCL
-1.0

ICL
-1.5
10

100

1000

10000

Vertical effective stress, 'v (kPa)

Figure 6.51: Development of structure with treatment

Constrained modulus, D = 1/mv (kPa)

1e+5
CRS046[RC]
CRS049[8.00%PC]
CRS043[18.7%PC]
CRS051[51.4%PC]
CRS055[103.4%PC]

'p 103.4%PC

1e+4

'p 51.4%PC

'p 18.7%PC
1e+3

'p 8.00%PC
'p RC
1

10

100

1000

10000

Vertical effective stress, 'v (kPa)

Figure 6.52: Increase in the constrained modulus with treatment

369

Constrained modulus, D = 1/mv (kPa)

1e+5

1e+4

1e+3

CRS046[RC]
CRS049[8.00%PC]
CRS043[18.7%PC]
CRS051[51.4%PC]
CRS055[103.4%PC]

1e+2
0.01

0.1

10

Normalized vertical effective stress, 'v/'p

Figure 6.53: Constrained modulus vs. normalized vertical effective stress

4.0

3.5

Void ratio

3.0

2.5

2.0

CRS046[RC]
CRS049[8.00%PC]
CRS043[18.7%PC]
CRS051[51.4%PC]
CRS055[103.4%PC]

1.5

1.0
1e-10

1e-9

1e-8

1e-7

1e-6

1e-5

Hydarulic conductivity, k (cm/sec)

Figure 6.54: Increase in hydraulic conductivity with treatment

370

Coefficient of consolidation, Cv (cm2/sec)

1e+0
CRS046[RC]
CRS049[8.00%PC]
CRS043[18.7%PC]
CRS051[51.4%PC]
CRS055[103.4%PC]

1e-1

1e-2

1e-3

1e-4
1

10

100

1000

10000

Vertical effective stress, 'v (kPa)

Figure 6.55: Increase in coefficient of consolidation with treatment

Coefficient of consolidation, Cv (cm2/sec)

1e+0

1e-1

1e-2

1e-3

1e-4

1e-5
0.01

CRS046[RC]
CRS049[8.00%PC]
CRS043[18.7%PC]
CRS051[51.4%PC]
CRS055[103.4%PC]

0.1

10

Normalized vertical effective stress, 'v/'p

Figure 6.56: Coefficient of consolidation vs. normalized vertical effective stress

371

Cement Factor (kg/m3)


0.10

56

106

230

193

152

265

0.09
0.08

C/Cc

0.07
Peat and muskeg

0.06
0.05

Inorganic clays and silts

0.04
0.03
0.02
0.01

Granular soils

20

40

60

80

100

120

%PC (by dry mass)

Figure 6.57: Effect of treatment on the C/Cc of LR soil

3.7
3.5
3.3

Void ratio

3.1
2.9
2.7
2.5
2.3
2.1
1.9
10

CRS051[51.4%PC,48kPa/14days]
CRS060[52.6%PC,48kPa/28days]
IL059[52.6%PC,48kPa/28days]
100

1000

Vertical effective stress, 'v (kPa)

Figure 6.58: Effect of curing time on the compression curve of LR soil treated
with 50% PC

372
3.7
ICL
3.5
3.3

Void ratio

3.1

b = 0.08

2.9
2.7
2.5

b = 0.4

2.3
CRS060[52.6%PC,48kPa/28days]
2.1
1.9
10

100

1000

Vertical effective stress, 'v (kPa)

Figure 6.59: Effect of curing time on the development of structure in LR soil


treated with 50% PC

Constrained modulus, D = 1/mv (kPa)

1e+5

1e+4

CRS051[51.4%PC,48kPa/14days]
CRS060[52.6%PC,48kPa/28days]
IL059[52.6%PC,48kPa/28days]
1e+3

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 6.60: Effect of curing time on the constrained modulus of LR soil treated
with 50% PC

373

3.7
3.5
3.3

Void ratio

3.1
2.9
2.7
2.5
CRS051[51.4%PC,48kPa/14days]
CRS060[52.6%PC,48kPa/28days]
IL059[52.6%PC,48kPa/28days]

2.3
2.1
1.9
1e-8

1e-7

1e-6

Hydarulic conductivity, k (cm/sec)

Figure 6.61: Effect of curing time on hydraulic conductivity of LR soil treated with
50% PC

Coefficient of consolidation, Cv (cm /sec)

0.1

0.01
CRS051[51.4%PC,48kPa/14days]
CRS060[52.6%PC,48kPa/28days]
IL059[52.6%PC,48kPa/28days]
0.001
10

100

1000

Vertical effective stress, 'v (kPa)

Figure 6.62: Effect of curing time on the coefficient of consolidation of LR soil


treated with 50% PC

374

3.0

800

Void ratio

700

2.8

600

2.7

500

2.6

400

2.5

300

2.4

200

Treatment with 52.6% PC


(CRS059)
'v = 1573 kPa

2.3
2.2
2.1
1e-2

1e-1

1e+0

1e+1

100
EOPExcess pore pressure

1e+2

1e+3

Excess Pore Pressure (kPa)

Time settlement
Excess Pore Pressure

2.9

1e+4

-100
1e+5

Elapsed time (min)

Figure 6.63: Effect of curing time on creep behavior of LR soil treated with 50%
PC

2.9
2.7
2.5

Void ratio

2.3
2.1
1.9
1.7
CRS064[100.8%PC, 48kPa/14days]
CRS065[100.8%PC, 48kPa/28days]
IL066[100.8%PC, 48kPa/28days]

1.5
1.3

10

100

1000

10000

Vertical effective stress, 'v (kPa)

Figure 6.64: Effect of curing time the on compression curve of LR soil treated
with 100.8% PC

375
2.9

2.7

Void ratio

2.5

2.3

b = 0.07

ICL

2.1

1.9

b = 0.5
CRS065[100.8%PC, 48kPa/28days]

1.7

1.5
10

100

1000

10000

Vertical effective stress, 'v (kPa)

Constrained modulus, D = 1/mv (kPa)

Figure 6.65: Effect of curing time on the development of structure in LR soil


treated with 100.8% PC

1e+5

CRS064[100.8%PC, 48kPa/14days]
CRS065[100.8%PC, 48kPa/28days]
IL066[100.8%PC, 48kPa/28days]

1e+4

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 6.66: Effect of curing time on the constrained modulus of LR soil treated
with 100.8%PC

376

2.9
2.7
2.5

Void ratio

2.3
2.1
1.9
1.7
CRS064[100.8%PC, 48kPa/14days]
CRS065[100.8%PC, 48kPa/28days]

1.5
1.3

1e-8

1e-7

1e-6

1e-5

1e-4

Hydarulic conductivity, k (cm/sec)

Figure 6.67: Effect of curing time on the hydraulic conductivity of LR soil treated
with 100.8% PC

Coefficient of consolidation, Cv (cm /sec)

0.1

0.01
CRS064[100.8%PC, 48kPa/14days]
CRS065[100.8%PC, 48kPa/28days]
IL066[100.8%PC, 48kPa/28days]
0.001
10

100

1000

Vertical effective stress, 'v (kPa)

Figure 6.68: Effect of curing time of the coefficient of consolidation of LR soil


treated with 100.8% PC

377

2.7

700

Void ratio

600

2.5

500

2.4

400

2.3

300

2.2

200

2.1

Treatment with 100.8% PC


(IL066)
'v = 1579 kPa

2.0
1.9
1e-2

1e-1

1e+0

1e+1

100

Excess Pore Pressure (kPa)

Time settlement
Excess Pore Pressure

2.6

EOPExcess pore pressure


0

1e+2

1e+3

1e+4

-100
1e+5

Elapsed time (min)

Figure 6.69: Effect of curing time on creep behavior of LR soil treated with
100.8% PC

3.7
3.5
3.3

Void ratio

3.1
2.9
2.7
2.5
CRS051[51.4%PC, 48kPa/14days]
CRS058[52.6%PC, 96kPa/14days]
IL057[52.6%PC, 96kPa/14days]
CRS062[48.4%PC, 192kPa/14days]
IL063[48.4%PC, 192kPa/14days]

2.3
2.1
1.9

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 6.70: Effect of curing surcharge on compression behavior of LR soil


treated with 50% PC

378

750

Preconsolidation pressure, 'p (kPa)

700
650
600
550
500
450
400
350

48

96

144

192

Curing surcharge (kPa)

Figure 6.71: Increase in preconsolidation pressure with increasing curing


surcharge

Constrained modulus, D = 1/mv (kPa)

1e+5

1e+4

CRS051[51.4%PC, 48kPa/14days]
CRS058[52.6%PC, 96kPa/14days]
IL057[52.6%PC, 96kPa/14days]
CRS062[48.4%PC, 192kPa/14days]
IL063[48.4%PC, 192kPa/14days]
1e+3
10

100

1000

Vertical effective stress, 'v (kPa)

Figure 6.72: Effect of curing surcharge on constrained modulus of LR soil treated


with 50% PC

379

3.5
3.3
3.1

b = 0.04

Void ratio

2.9
ICL
2.7
2.5

b = 0.4

2.3
2.1
1.9
10

CRS058[52.6%PC, 96kPa/14days]

100

1000

Vertical effective stress, 'v (kPa)

Figure 6.73: Destructuring of 52.6%PC treated LR soil cured under 96 kPa

3.1

2.9

b = 0.0
ICL

Void ratio

2.7

2.5

b = 0.3

2.3

2.1

1.9
10

CRS062[48.4%PC, 192kPa/14days]

100

1000

Vertical effective stress, 'v (kPa)

Figure 6.74: Destructuring of 48.4%PC treated LR soil cured under 192 kPa

380

3.7
3.5
3.3

Void ratio

3.1
2.9
2.7
2.5
CRS051[51.4%PC, 48kPa/14days]
CRS058[52.6%PC, 96kPa/14days]
IL057[52.6%PC, 96kPa/14days]
IL063[48.4%PC, 192kPa/14days]

2.3
2.1
1.9
1e-8

1e-7

1e-6

Hydarulic conductivity, k (cm/sec)

Figure 6.75: Effect of curing surcharge on the hydraulic conductivity

Coefficient of consolidation, Cv (cm /sec)

0.1

0.01
CRS051[51.4%PC, 48kPa/14days]
CRS058[52.6%PC, 96kPa/14days]
IL057[52.6%PC, 96kPa/14days]
IL063[48.4%PC, 192kPa/14days]
0.001

10

100

1000

Vertical effective stress, 'v (kPa)

Figure 6.76: Effect of curing surcharge on the coefficient of consolidation

381

Void ratio

Untreated
Treated

Improvement
with treatment

Effect on Virgin Compression Line =


f (amount of binding agent, organic content,
pH of pore water )

Vertical effective stress, 'v (log)

Figure 6.77: Schematic diagram of the effect of cement treatment on the position
of the virgin compression curve

Void ratio

Untreated

Treated

Initial void ratio


Preconsolidation puressure

Recompression curve

Improvement
with treatment

Effect on Preconsolidation pressure =


f (amount of binding agent, organic content,
curing surcharge, curing water content,
pH of pore water)

Vertical effective stress, 'v (log)

Figure 6.78: Schematic diagram of the effect of cement treatment on the


preconsolidation pressure

382

Constrained modulus, D = 1/mv (log)

Preconsolidation pressure

Over Consolidated Region


Improvement
with treatment

Treated

Normally Consolidated Region

Treated

Untreated

Effect on Constrained Modulus =


f (amount of binding agent)

Vertical effective stress, 'v (log)

Figure 6.79: Schematic diagram of the effect of cement treatment on the


constrained modulus

383

3.5

Void ratio, e

3.0
2.5
2.0
0%lime
1%
2%
3.5%
5%
10%

1.5
1.0
0.5 -11
10

10-10
10-9
Hydraulic conductivity, k (m/sec)

10-8

Figure 6.80: Hydraulic conductivity of lime stabilized Louiseville clay (Locat et al,
1996)

Treated

Preconsolidation pressure

Coefficient of consolidation, Cv (cm /sec)

384

Untreated

Over Consolidated Region


Improvement
with treatment
Normally Consolidated Region
Effect on Coefficient of Consolidation =
f (amount of binding agent)

Vertical effective stress, 'v (log)

Figure 6.81: Schematic diagram of the effect of cement treatment on the


coefficient of consolidation

385

Figure 6.82: Changes in FT-IR spectra of non-extracted fraction of the PC treated


LR soil

386

Figure 6.83: Changes in FT-IR spectra of humic acid extracted from the PC
treated LR soil

387

Figure 6.84: Changes in FT-IR spectra of fulvic acid extracted from the PC
treated LR soil

388

CHAPTER 7. CONCLUSIONS AND RECOMMENDATIONS

7.1. Introduction

This research performed for this thesis was built around four major work streams
aimed at fulfilling the objectives outlined in Chapter 1: sampling and
characterization of the soil; consolidation tests on natural LR soil; consolidation
tests on cement treated LR soil; chemical studies of the interaction between
cement and organic soil. This chapter summarizes the results obtained and the
conclusions drawn from this work, and is organized according to these four work
streams. Additionally, recommendations for future research are presented.
7.2. Results and Conclusions
7.2.1. Sampling and characterization of LR soil
The soil employed in this experimental program was obtained in the vicinity of
Lindberg Road, in West Lafayette, Indiana. It is referred to throughout this thesis
as LR soil. Two types of soil samples, block samples and disturbed samples,
were collected in April and July of 2001. Significant variability was observed in
the index properties. The average index properties of the soil samples can be
summarized as follows:
- natural water content = 155 % 266.6 %;
- loss on ignition = 40.0 % 60.1 %;
- liquid limit = 260.6 % 405.7 %;
- liquid limit after oven-drying = 82.2 % 102.1 %;
- plastic limit = 113.5 % 252.7 %.
The specific gravity of the soil varies between approximately 1.9 and 2.2 and
shows a linear relationship with the measured loss on ignition (Gs = -0.0104 x
LOI (%) + 2.570) (Figure 3.3). On the plasticity chart, LR soil plots slightly below

389
the A-line and falls in a region of the chart typical of highly plastic organic clays
(OH) and various types of peats (Figure 3.8). Lindberg Road (LR) soil consists of
approximately 4.9 % sand (>0.075 mm), 53.2% silt (<0.075 and >0.002 mm) and
41.5% clay (<0.002mm) (Figure 3.12)) as defined by the USCS. Based on the
particle size distribution and the liquid limit ratio (ratio of the value measured after
oven drying to that measured on the natural soil), LR soil can be classified as an
organic silt with sand (OH), according to the USCS (ASTM D2488).
The fiber content of LR soil was approximately 2.29%. This would categorize LR
soil as sapric, an organic soil with fiber content less than 33%, according to
ASTM D1997.
LR soil was fractionated into two main components, inorganic and organic
components, to evaluate the mineralogy of the inorganic component and the
chemical composition of the organic matter. The particle size distribution of the
inorganic component of the soil showed that the soil contained about 62% silt
size and 38% clay size fractions, thereby indicating (compare to fractions listed
above for the natural soil) that the presence of organic matter leads to
aggregation and agglomeration of the soil particles, particularly belonging to the
smaller size fractions (Figure 3.12).
X-ray diffraction (XRD) analyses were performed on various size fractions of the
mineral component of LR soil. The results of these analyses indicate the
presence of highly active clay minerals such as smectite and vermiculite in the
clay size fraction (Table 3.4). These clay minerals are likely to be responsible for
adsorbing the organic matter and aggregating to form larger particles.
The organic component of LR soil was extracted and fractionated into three main
fractions: humin and bound inorganic components (74%), humic acid (24%) and
fulvic acid (2%). The elemental compositions of the three fractions were

390
evaluated by performing LECO tests, and were found to be similar to those of
various other organic soils. However, the carbon content of both the humic and
fulvic acids of LR soil was observed to be lower, for example, than that typical of
some histosols, evidence of the high degree of humification of the LR soil (the
C/N ratio of humic acid decreases with humification). This is consistent with the
very low fiber content (about 2.29%) of LR soil.
Fourier-transform Infrared Spectroscopy (FT-IR) analyses were performed on the
three LR soil fractions to determine the types of organic acids present in the soil.
Among the organic acids identified in the three fractions of LR soil from the FT-IR
spectrum, are phenolic OH, carboxyl acid, and polysaccharide, compounds that
are known to act to retard and even inhibit cement hydration (Table 3.10).
7.2.2. Review of the 1-D Consolidation Behavior of Intact and Reconstituted
LR soil
Consolidation tests were performed on LR soil specimens prepared from both the
block samples and the disturbed samples collected in April and July of 2001.
The block soil samples were used to investigate the 1-D consolidation behavior
of the intact soil. Reconstituted soil samples were manufactured in the laboratory
following a sample preparation procedure (consisting of a mixing, a compaction
and a curing stage) developed to simulate the kneading action applied in the
deep mixing method (see details in Section 4.3). The results of the consolidation
tests performed on reconstituted soil samples were used as a reference against
which to evaluate both the behavior of the intact structured soil, and the effects of
cement treatment on the 1D consolidation behavior of the soil. For both sample
types the one-dimensional consolidation behavior was investigated by performing
both constant rate of strain (CRS) and incremental loading (IL) tests.
The effects of strain rate on the preconsolidation stress, the compressibility, the
hydraulic conductivity and the coefficient of consolidation were evaluated by
performing CRS consolidation tests on reconstituted soil specimens employing

391
strain rates ranging from 0.1%/hr to 1.0%/hr. The results indicated that the
excess pore pressures developed at the bottom of specimen increased linearly
with increasing strain rate. However, the excess pore pressures normalized by
the actual strain rate remained essentially constant regardless of the employed
strain rate (Figure 5.24). The results also showed that, with the possible
exception of the results of one test conducted at a strain rate of 1%/h, within the
range of rates evaluated, the 1-D consolidation behavior of LR soil was largely
unaffected by rate (Figure 5.23, Figure 5.26, Figure 5.27). Based on the results,
a strain rate of 0.5%/hr was employed for the CRS tests on both block and
reconstituted soil specimens.
The CRS tests on the intact LR soil indicate a S-shaped compression curve
(Figure 5.29), evidence of the metastable structure of the natural material, and a
reflection of the high quality of the sampling operations. Also supporting the
quality of the sampling operations are the measurements of the reconsolidation
strains, which are very small.
Based on the CRS tests the preconsolidation stress of the intact soil is found to
range between 106 kPa and 121 kPa, corresponding to an in-situ
overconsolidation ratio around 20, probably resulting from the combination of
aging and desiccation.
The compressibility of the natural soil is consistent with relationships proposed in
the literature to natural water content (Figure 5.31). Due to the S-shape of the
compression curve, the compression index of the soil changes significantly over
the stress range investigated (Figure 5.32). In the normally consolidated region,
Cc first increases sharply, but then, beyond approximately 2'p, it decreases
continuously with increasing vertical effective stress from an average value of
2.620 (in the 24 'p range) to 2.369 (48 'p), 1.959 (816 'p), and 1.577 (16
32 'p).

392
The hydraulic conductivity of the intact soil was calculated from the excess pore
pressure measured at the bottom of the CRS specimens during loading. The
block samples contained numerous cracks, which made it difficult to measure the
excess pore pressure accurately at the beginning of the test and thus the
estimates of the in-situ k are not considered reliable. Once the soil was
consolidated into the virgin compression region the data was, instead,
considered representative of the hydraulic conductivity of the soil matrix (Figure
5.43). From these data a value of Ck (Ck = e/logk), ranging between 0.821
and 0.894, was calculated. This value falls in between values reported for clays
and peats.
The coefficient of consolidation (Cv) curve of the block LR soil is characterized by
a sharp and significant decrease in the over consolidated region, followed by a
continuous but less significant decrease in the normally consolidated region
(Figure 5.45). The continuous decrease of Cv in the normally consolidated region
is due to the fact that the reduction in the hydraulic conductivity is greater than
the increase in the constrained modulus (D = 1/mv) (Figure 5.41).
The results of the consolidation tests performed on the intact LR soil specimens
were compared to data for two other soils used as reference geomaterials:
Middleton peat from Middleton, Wisconsin, with 9095 % organic content, whose
behavior is considered representative of that of natural, highly fibrous peats; and
Boston blue clay (BBC), an inorganic illitic low plasticity marine clay. It was
found that both the compression and the hydraulic conductivity data for LR soil fell
in between the results for these two reference soils (Figure 5.30, Figure 5.44).

The secondary compression (creep) coefficient of LR soil was measured from the
creep stages of the IL consolidation tests (Figure 5.54), and the results
interpreted in the framework of the C/ Cc ratio proposed by Mesri and Castro
(1987). For this purpose the secondary compression index (C = e/logt)

393
obtained from each load increment was normalized by the compression index
obtained from CRS consolidation tests in correspondence to the same vertical
effective stress. The results showed that for LR soils C/ Cc is about 0.095
(Figure 5.57, Figure 5.58). This values lies at the very high end of the range
reported for peats. For these soils Mesri et al. (1997) report data as high as
0.10, but more typically in the 0.05-0.07 range.
The compression behavior of the intact LR soil was normalized based on the
consolidation test results performed on reconstituted LR soil specimens,
employing the framework proposed by Burland (1990). The compression curves
of reconstituted LR soil were used to define the intrinsic compression line (ICL).
The void index (Iv) of the block LR soil at the preconsolidation pressure falls very
close to the sedimentation compression line (SCL) (Figure 5.35).
According to Burland (1990) the SCL represents the in-situ state of various
naturally occurring normally consolidated clays, and lies at a vertical effective
stress about five times higher than the ICL line at the same void ratio. In
correspondence to the preconsolidation stress the data for intact LR soil fall right
around the SCL. This suggests that the SCL may represent a useful reference line
even for soils with high organic content such as that examined in this research.

The compression curves approach the ICL (Figure 5.35) at higher stresses in the
normally consolidated region indicating destructuring of the natural structure. The
rate of destructuring with increasing vertical effective stress in the normally
consolidated region was quantified using the destructuring index (b) introduced
by Liu and Carter (2000). For LR soil the b index is approximately 0.8. This is
consistent with results for natural soft clays provided by Carter (2000) which
show b between 0.3 and 1.
The tests conducted on reconstituted soil specimens prepared from samples
collected at different times and having slightly different organic content (LOI = 45-

394
55% versus 36-37%) also provided an opportunity to isolate the impact of
variations in organic content (given the proximity between the sampling locations
it was assumed that there were no major differences in the nature of the mineral
substrate) on the position of the compression curve (Figure 5.22) as well as on
the hydraulic conductivity at a given void ratio (Figure 5.26). Specifically, it was
found that the increase in LOI was associated with the ability to sustain a greater
effective stress at the same void ratio (by about a factor of 2 in the NC region)
and a reduction in k (by a factor of 3-5 in the NC region).

7.2.3. Conclusions on the 1-D Consolidation Behavior of LR soil


The work performed on intact and reconstituted LR soil leads to conclusions that
pertain: a) to sampling methods and sample quality for highly organic soils; b) to
the behavior of a (sapric) transitional geomaterial; c) to the local variability of
organic soils and its impact on engineering properties; d) to the structure of
highly organic soils and the relationship between properties of intact and
reconstituted soil; e) to the high susceptibility to creep of sapric soils.
a) sampling methods and sample quality
The research performed on intact natural LR soil relied on testing block samples
obtained using a sampler specifically designed for this research project. The
results from the engineering tests performed on this soil attest to the high quality
of this sample. Specifically, it is found that the volumetric strains associated with
reconsolidation to the in situ vertical effective stress are significantly smaller than
the threshold value used to differentiate between excellent and good samples.
Additionally, the S-shape of the compression curves observed in the
compression tests is evidence that no significant destructuring occurred during
sampling.
It can be concluded that it is possible to obtain high quality samples of highly
organic soils using the equipment designed for this project.

395
b) behavior of a (sapric) transitional geomaterial
The material investigated in this research has 40-60% organic matter, formed
primarily by highly processes humic substances, and an inorganic fraction
formed primarily by clay. While there exists much data in the literature on
inorganic clays and much is known about the behavior of peats (especially
fibrous), very little data are available for soils with intermediate values of the
organic content.
Thus the results obtained for the soil tested can be useful to draw some
conclusions on the behavior of a sapric soil that can be neither classified as a
clay, nor as a peat. It is found that many of the properties investigated
(compressibility, hydraulic conductivity, position in stress space) for this soil fall in
between those typical of inorganic clays and those of peats, i.e. the behavior is
controlled by organic content (or water content). As for peats, there is evidence
from the hydraulic conductivity data that the concept of dual porosity (macro
versus micro porosity) applies also for these soils.
In one important aspect the behavior of LR soil falls out of the range defined by
inorganic clays and peats, that is the secondary compression, which is discussed
below (point e)
c) local variability of organic soils and impact on engineering properties
Samples used for testing in this research program were obtained at two different
times in close proximity to each other. Despite this, significant sample to sample
variability in the index properties was observed. In particular the organic content,
as measured by the LOI test, was observed to go from 36-37% in one sample to
45-55% in the other. Comparison of the data obtained from these two samples
leads to conclude that even relatively small changes in organic content can have
significant effect on the vertical effective stress that can be sustained at a given
void ratio and the values of hydraulic conductivity at a given void ratio. For LR
soil both increase with LOI by factors of 2 and 3-5, respectively.

396
d) structure of highly organic soil and relationship between properties of intact
and reconstituted soil
In this research the investigation of the structure of the intact natural soil relied on
two complementary approaches: observations using cryo-SEM which provide a
direct assessment of the nature and organization of the microstructure of the
soil; and the comparison of compression data for intact and reconstituted soil,
which provides information on mechanical aspects of the structure, i.e. the
ability to sustain a high effective stress at a given void ratio and the stable or
metastable nature of the structure upon loading to higher stresses.
From the cryo-SEM images it is possible to conclude that in LR soil the organic
matter and the inorganic fraction are deeply interpenetrated with organic matter
coating the mineral substrate. This is an important observation as information
and micrographs of the microstructure of non fibrous organic materials are not
available in the geotechnical literature. Note that understanding the
microstructure of organic soils is critical to explaining the high tendency to creep
(see point e below) as well as the increase in strength parameters reported in the
literature for even small increases in organic matter.
Comparison of the behavior of reconstituted and intact soil leads to conclude that
the degree of structuring measured for LR is consistent with that observed for
sedimentary inorganic clays, based on the work by Burland (1990). It is also
concluded that the structure is metastable, and that the post yield structure
degradation quantified through Liu and Carters (2000) destructuring index is
consistent with that typical of many soft inorganic clays.
Comparison of the data for intact and reconstituted soil also leads to conclude
that certain parameters, specifically Ck and C/Cc, are independent of structure.
e) high susceptibility to creep of a sapric organic soil
While many of the engineering properties of LR soil largely fall in between
the values reported for inorganic clays and peats, one important exception is
represented by the C/Cc ratio, which, at 0.095, significantly exceeds typical

397
values reported for peats. This suggests that the common assumption that creep
is most significant in peats and muskegs may be inaccurate.
7.2.4. Review of the effects of treatment on the 1-D consolidation behavior of
LR soil
The main goal of this research was to evaluate the effects of cement treatment
on the 1-D consolidation behavior of a highly organic soil. Humphrey (2001)
observed that Type I ordinary Portland cement was the most effective binder to
improve the compressive strength of LR soil. As such, the same binder was
employed in this research.
Cement treated specimens were prepared using the disturbed LR soil samples
collected from April, 2001, by adding cement during the mixing phase of the
same sample preparation procedure used for the untreated reconstituted soil
samples. The cement dosage was varied between 8% (by dry mass of the soil)
and 100%. These dosages correspond to values of the cement factor equal to
approximately 23 and 230 kg of cement per m3 of treated soil, which are in the
range of values used in field applications of deep mixing (see Section 2.6). For
each %PC, at least 2 CRS and 1 IL consolidation tests were performed. In
addition, the effects of curing time and surcharge were also evaluated.
7.2.4.1. Effects of % of Portland cement
The test results showed that the compression curves shift gradually to higher
vertical effective stresses with increasing cement content (Figure 6.47). The
preconsolidation pressure increases with increasing cement content, and the
over consolidated region extends to higher vertical effective stresses. At any
void ratio the cement treated soil can sustain higher vertical effective stress
compared to the untreated soil.
For a given curing surcharge (see below for discussion of effect of curing
surcharge), the preconsolidation stress increases with %PC (Figure 6.50). For
example, for the reference curing stress (~48 kPa) 'p goes from 74.8 kPa (8%

398
PC), to 94.6 kPa (18.7% PC), to 403.3 kPa (51.4% PC), to 1189.8 kPa (100.8
103.4% PC). This last value represents a 2431% increase with respect to the
average value of 'p determined for the reconstituted soil (57 kPa). These data
show that 'p increased slightly up to a cement content of 18.7%, but that the
increase was much sharper when the soil was treated with 50% or more PC. The
increase in the preconsolidation pressure with treatment is closely related to the
change in pH of the soil. The pH of the soil reached above 12.6 when treated
with more than 50% PC (Figure 6.40). This is the pH above which cementitious
product remains stable and can continue hardening with time.
Similar to the compression curves of block LR soil specimens, the compression
curves of PC treated LR soils were normalized employing the framework
proposed by Burland (1990), using the intrinsic soil parameters e*100 and e*1000
obtained from the reconstituted LR soil specimens. With treatment, the
normalized compression curves of LR soil gradually move from the intrinsic
compression line towards the sedimentation compression line (Figure 6.51). In
correspondence to 51.4% PC the normalized compression curve lies slightly
above the sedimentation compression line, i.e. with this cement dosage, the LR
soil can sustain 5 6 times higher vertical effective stress compared to the
reconstituted material. For 100.8 103.4 %PC the compression curve for
this %PC also lies just to the right of the SCL (the greater value of 'p is a result
of the lower void ratio of the 100% PC treated soil).
The destructuring index (b) of PC treated LR soil increased slightly with
increasing %PC, from zero (8% and 18.7% PC, Figure 6.41 and Figure 6.42) to
0.07 (51.4% PC, Figure 6.43) and to 0.11 (100.8% PC, Figure 6.44) and 0.13
(103.4% PC, Figure 6.45). While these values of b are fairly small, indicating the
stable nature of the structure generated by treatment with cement, there is a
trend of increasing brittleness with increasing cement content. This observation is
consistent with data for stabilized Louiseville clay reported by Locat et al. (1996).

399
While the preconsolidation pressure of LR soil increases with treatment, the
compression index does not show significant change with treatment. The virgin
compression curves of PC treated LR soils are almost parallel to or slightly
steeper than the compression curve of reconstituted LR soil. In the vertical
effective stress range of 2 4 p, the compression index of LR soil increased by
2.6% (with 18.7 % PC) to 31.4% (with 51.4% PC) with respect to the
compression index of the reconstituted LR soil.
Although the compression index does not show significant changes with
treatment, the swelling index and reloading index decreased with treatment. The
swelling index showed 55.5%, 70.7%, 82.0% and 94.8% decrease with treatment
with 8.0%, 18.7%, 51.4% and 100.8 103.4% PC at the same OCR. The
reloading index decreased by 66.3%, 81.8%, 90.0% and 97.5% with treatment.
The hydraulic conductivity of LR soil was found to increase with treatment at the
same void ratio (Figure 6.54). The effect of treatment with 8.0% PC is modest
with a 18 26% increase in k at the same void ratio. With treatment with 18.7%
PC and 51.4% PC, the hydraulic conductivity increased by 4.6 to5.7 times at the
same void ratio with respect to the reconstituted LR soil. For 100.8 103.4% PC
there a 16-20 fold increase in k was observed. The increase in the hydraulic
conductivity of LR soil with treatment is mainly due to the change of the fabric of
the soil, which, as directly observed by Tremblay et al. (2001) and Al-Rawas
(2002), becomes flocculated and aggregated yielding a more open fabric and
more granular-like particles. As a result, the size of the macropores, which serve
as the main channels for flow, increases causing the increase in hydraulic
conductivity. Although the hydraulic conductivity of LR soil increased significantly
with cement treatment, the change of hydraulic conductivity with void ratio (Ck)
remained constant.

400
As the constrained modulus (Figure 6.53) and the hydraulic conductivity (Figure
6.54) of LR soil increase with cement treatment, the coefficient of consolidation of
LR soil also increases with cement treatment (Figure 6.55). Similar to the
previous observation relative to the preconsolidation pressure and hydraulic
conductivity, the effects of treatment with 8.00% PC on the coefficient of
consolidation are small. Compared to the reconstituted LR soil, the coefficient of
consolidation increased by approximately 7 times, 25 times, and 32 times when
the soil was treated with 18.7%, 51.4%, and 100.8 103.4% PC, respectively.
The increase in the coefficient of consolidation with PC treatment indicates that
consolidation takes place much faster when the soils are treated with cement.
For practical purpose, this implies that construction time can be significantly
reduced by treating soils with cement.
The effect of treatment on the secondary compression behavior of LR soil was
evaluated by performing long term creep tests. The specimens were
consolidated into the normally consolidated region following the EOP
compression line. The results showed that the slope of the time settlement curve
remained constant within the time range investigated (e.g. Figure 6.25 and
Figure 6.39. Again, as in the case of the untreated soil, the secondary
compression data was interpreted in terms of the C/Cc ratio. It was found that
the C/Cc ratio gradually decreased with increasing cement content (Figure 6.57).
Compared to the reconstituted LR soil (C/Cc =0.095), the C/Cc ratio of the LR
soil decreased to approximately 0.085 (8.0% PC), 0.063 (18.7% PC), 0.043
(51.4% PC), and 0.022 (100.8 - 103.4% PC). This last value is in the range
typically observed for granular soils and clearly illustrates the change in the
nature of the soil with cement treatment. Given that the compression index of LR
soil remained constant or slightly increased with cement treatment, the decrease
in the C/Cc ratio is mainly due to the decrease in the secondary compression
index. The decrease of C/Cc ratio with cement treatment is due to (1) the
replacement of the soil with less creep susceptible cement, (2) immobilization of

401
soil particles, especially organic matter, due to the bonding developed by
cementitious product, and (3) increase in the soil particle size. Organic soils and
peats have the highest C/Cc ratio among various types of soils (Mesri et al,
1997). The results obtained in this research clearly showed that significant
reduction in secondary compression can be obtained by cement treatment.
7.2.4.2. Effects of Curing Time
To evaluate the change in the 1-D consolidation behavior of cement treated LR
soil with time, additional CRS and IL consolidation tests were performed on soils
specimens treated with about 52.6% and 100.8% PC. Soil specimens were
cured for 14 and 28 days under a surcharge of 48 kPa. The results generally
indicate that the effects of curing time on 1-D consolidation behavior including
preconsolidation pressure (Figure 6.58), compression index, hydraulic
conductivity (Figure 6.61), and coefficient of consolidation (Figure 6.62), were
almost negligible, at least within the time range evaluated. The destructuring
index b also remained the same with time (Figure 6.59), indicating no significant
modification to the structure.
7.2.4.3. Effects of Surcharge Stress
Hebib and Farrell (2003) observed that the compressive strength of cement
treated peat samples increased when the samples were preloaded during the
preparation procedure. Humphrey (2001) also observed a similar effect on the
unconfined compressive strength and stiffness of cement and lime treated LR
soil cured under different surcharge stresses. To evaluate the effects of
surcharge on the 1-D consolidation behavior of cement treated LR soil, CRS and
IL consolidation tests were performed on soil specimens treated with about 48.4
52.6% PC. The soil samples were cured for 14 days by applying surcharges of
48, 96, and 192 kPa.
The test results showed that the average preconsolidation pressure increased
with increasing curing surcharge from 403.4 1.7 kPa (surcharge = 48 kPa) to

402
471.4 5.0 kPa (surcharge = 96 kPa), and to 672.7 24.4 kPa (surcharge = 192
kPa) (Figure 6.70). The results also indicate that the increase in the
preconsolidation pressure is higher than the increase in the surcharge (Figure
6.71). Test results showed that the initial void ratio of the soil specimen
decreased with increasing surcharge. This implies that the soil and cement
particles are more closely compacted when a higher surcharge is applied, and
thus an imporved bonding between soils and cement particles is developed. The
results also imply that the preconsolidation pressure of PC treated soils can be
enhanced by curing the soil under higher surcharge. Although the
preconsolidation pressure of PC treated LR soil increased with increasing
surcharge, the virgin compression curve remained the same regardless of the
magnitude of the applied surcharge (Figure 6.70). Therefore, the increase in the
preconsolidation pressure with surcharge is mainly due to the decrease in the
initial void ratio of the soils rather than the change in the virgin compression line.
The destructuring index of LR soil during compression decreased gradually with
increasing curing surcharge. The destructuring index of LR soil cured under 48,
96 and 192 kPa was 0.07 (Figure 6.43), 0.04 (Figure 6.73) and 0.0 (Figure 6.74),
respectively. The results also support the hypothesis that the stiffness of the
structure developed by the bonding of soil and cement particles increases with
increasing surcharge.
Despite the scatter in the hydraulic conductivity data in the early stage of
consolidation tests, the test data generally indicate that the hydraulic conductivity
of LR soil specimens treated with about 48.4 52.6% PC slightly increases with
increasing surcharge (Figure 6.75). The gradual increase in the hydraulic
conductivity with increasing curing surcharge also suggests that the LR soil cured
under higher surcharge exhibits higher degree of aggregation. The coefficient of
consolidation also increased with increasing curing surcharge (Figure 6.76),

403
mainly due to the increase in hydraulic conductivity and constrained modulus
with surcharge.

7.2.5. Review of organic soil / cement interactions


In addition to the investigation of the effects of cement treatment on the 1-D
consolidation behavior of LR soil, this study investigated the interaction between
organic matter and cement. This phase of the work relied on the following
experiments:
a) Tests on the cement treated soil (20, 50 and 100% PC) for the extraction of
the humic and fulvic acids.
These tests, which followed the same procedures used to fractionate the organic
matter of the natural LR soil (see Chapter 3), were aimed at verifying to what
extent, after treatment with cement, humic acid and fulvic acid could still be
extracted from the soil cement matrix. The tests showed that following treatment
with cement only a small percentage of the humic acid and fulvic acid originally
present in the soil could be extracted (Table 6.12). For example, in the case of
20% PC, only 3% of the humic acid was extracted. This percentage fell to only
0.1% in the case of 50% PC. Following treatment with 100%, the amount of the
extracted humic acid was too small to be quantified. The amount of extracted
fulvic acid also decreased with cement treatment, albeit less significantly.
These results indicate that the humic acid and fulvic acid of LR soil either react
with the cement or are retained by cementitious products within the soil-cement
matrix.
b) Quantification of carbon content of three fractions through LECO test
These tests were performed to verify that, following the establishment, as a result
of the cement hydration process, of alkaline conditions, the fulvic and humic
acids had not dissolved into the soil solution. This occurrence could otherwise
also explain the inability to extract them from the soil cement matrix. The

404
measurements of the carbon content through the LECO test on the extracted
humic and fulvic acids and on the non-extracted fraction demonstrated that the
overall carbon content remained unchanged following treatment (Table 6.13),
confirming that the humic and the fulvic acid of the LR soil had not left the soilcement matrix and, were instead retained within it.
c) FT-IR tests on extracted humic and fulvic acids, and on the non-extracted
humin and inorganic component of the soil.
The aim of these experiments was to evaluate chemical changes occurring with
cement treatment through comparison of the FT-IR spectra to the corresponding
spectra obtained for the untreated LR soil.
The analysis of the FT-IR spectra showed the following important differences
between the two sets of spectra:
- the appearance in the spectrum of the non-extracted fraction of the PC treated
LR soil of a broad band in the 3000 3600 cm-1 region, indicating the
formation of cementitious products (Figure 6.82);
- the increase in the spectrum of the non-extracted fraction of the intensities of
the peaks (1576 1772 cm-1 and 1032 1085 cm-1 ) corresponding to
specific organic functional groups which thus appear to have been
incorporated and retained in the soil cement matrix (Figure 6.82);
- the increase in the spectra of the humic acid fraction of the PC treated LR soil
of the intensities of the bands corresponding to phenolic OH and carboxyl
acid (Figure 6.83). This means that on a relatively basis the humic acid
extracted from the cement treated soil is richer in these acids than the original
untreated soil. This implies that phenolic OH and carboxyl acid (COOH) have
relatively lower reactivity with cement than other organic acids present in the
humic acid of LR soil, and can be leached from the soil-cement matrix.
The significant decrease in the mass of the extracted humic acid with cement
treatment attests, however, to the fact that phenolic OH and carboxyl acid do

405
not significantly interfere with cement hydration. The results outlined above
become increasingly more significant with an increase with cement treatment.
7.2.6. Conclusions on the effects of cement treatment on the 1-D
consolidation behavior, the structure and the chemistry of a highly
organic soil

This section presents the conclusions drawn from the extensive work conducted
on cement treated LR soil. The discussion is organized to address the following
main topics:
a) Engineering properties of cement treated highly organic soil
b) Structure and chemistry of cement treated highly organic soil
c) Conditions for effectiveness of treatment
d) Applicability and advantages of deep mixing method for the
improvement of highly organic soils
a) Engineering properties of cement treated LR soil
The work performed for this research has led to the development of a unique
dataset on the engineering properties (relevant to 1-D consolidation) of a cement
treated highly organic soil. The data paint a novel and complete picture of the
behavior in 1D consolidation of the geomaterials created through treatment of the
soil with cement, mapping the change in important engineering properties
(preconsolidation stress, compressibility, hydraulic conductivity, coefficient of
consolidation, creep coefficient, C/Cc ratio) as a function of the testing
conditions (in particular cement content).
The main conclusions on the effects of cement treatment on the 1-D
consolidation behavior of a highly organic soil can be summarized as follows:
- as a result of cement treatment the virgin compression line of the soil shifts to
higher vertical effective reflecting the fact that the treated soil can sustain

406
higher vertical effective stress at any void ratio. The position of the virgin
compression line is independent of the initial void ratio/water content of the
soil. The greater the cement content, the more significant the shift in the
compression curve. For LR soil no further shift in the compression curve was
observed in going from 50% to 100% cement content.
- as a result of cement treatment the void ratio of the soil decreases;
- as a result of the shift of the compression curve to higher effective stresses,
the soil exhibits an increased value of the yield stress under 1-D drained
loading (preconsolidation stress), which is a result of the development of a
cementation structure. The value of the preconsolidation stress increases
with cement content, and, for a given cement content, is affected by the initial
void ratio (which determines at which stress level the reloading curve reaches
the virgin compression curve).
- below the preconsolidation stress the compressibility of the soil, measured by
the recompression index, decreases with increasing cement content.
- beyond the preconsolidation stress, the compressibility of the soil, measured
by the compression index, is not significantly affected by the treatment,
although for very high dosages of cement the compression index is observed
to increase slightly.
- continuous loading beyond the preconsolidation stress occurs with
approximately constant compression index, indicating the stable nature of the
structure created as a result of treatment with cement.
- the hydraulic conductivity, the constrained modulus and thus the coefficient of
consolidation all increase with cement treatment. The higher the cement
content, the more significant the change in these engineering properties.
- as a result of treatment with cement both the creep coefficient and the C/Cc
ratio decrease significantly. For LR soil at the highest cement content
examined in this research (~100% PC) C/Cc reaches values typical of
granular soils. The reduction in C/Cc is thought to reflect the replacement of
the soil with less creep susceptible cementitious products, the immobilization

407
of soil particles thanks to the bonding developed by cementitious product, and
the increase in the soil particle size. The fact that C/Cc decreases, while Cc
remains practically constant suggests that different mechanisms contribute to
the deformation of the soil during primary and secondary compression.
The effects on the engineering properties as a result of cement treatment
described above become more significant with increasing cement content. For
LR soil it was found that cement contents of approximately 8% and 18%
impacted the behavior of the soil in a negligible manner, and that dosages of
50% or greater were necessary to achieve a significant improvement in the
engineering properties.
The above described changes in the soil properties (in particular the increase in
hydraulic conductivity and coefficient of consolidation and the reduction in C/Cc)
attest to the evolution of the nature of the soil which for high cement contents
exhibits some characteristics typical of coarser grained soils (see more below in
b)
Based on the review of the available literature and the work conducted in this
research, it appears that the behavioral framework outlined above for cement
treated LR soil can at least qualitatively be extended to other organic and
inorganic soft soils. The changes in the engineering properties as well as the
conditions (e.g. cement content) required to achieve a certain level of
improvement in the engineering properties will however be soil specific.
b) Structure and chemistry of cement treated highly organic soil
Some conclusions on the structure and chemistry of cement treated highly
organic soils can be drawn from the consolidation tests as well as from the
results of the extraction tests and chemical tests.

408
As already mentioned above the mechanical and hydraulic properties measured
on the cement treated soils are consistent with a fabric formed of larger particles
created by the aggregation of smaller particles by the cementitious products.
With increasing cement content the particle size increases and the fabric of the
soil becomes more open, as reflected by the increasing stiffness (constrained
modulus), decreasing hydraulic conductivity and decreasing C (and C/Cc).
The results of the extraction tests and the chemical and functional
characterization of the different fractions extracted indicate that great part of the
soil organic matter reacts with the binder and/or is retained within the soil-cement
matrix that forms the aggregates. With increasing cement content the
aggregates become more compact (consistent with the increased stiffness and
reduced creep of the soil) effectively not allowing any of the organic matter to be
extracted.
c) Conditions for effectiveness of treatment
While the work performed for this research involved testing of only one soil, the
results obtained provide some insight into what might be the conditions
necessary for successfully treating highly organic soils through deep mixing.
In particular it is found that significant cement dosages (50-100% per dry mass of
the soil, corresponding to 120-230 kg of cement per 1m3 of treated soil) are
required to improve the properties of the soil to a degree that ensures the
effectiveness of the treatment. While these values are higher than those typically
used for mass treatment of other soft soils, they do fall within the range of values
used in field applications of deep mixing.
The values of cement required to improve the properties of LR soil (50-100%) are
specific to this geomaterial, and cannot be generalized to other organic soils.
Broader conclusions on the conditions required for cement treatment to be
effective can be, however, derived by observing that for LR significant changes in
the engineering properties (e.g. shift in compression curve, marked increase in

409
preconsolidation stress) are observed when the cement content is such that the
pH rises above the value required (12.6) for the C-S-H gel formed by the
hydration of the cement to remain stable. This suggests that pH measurements
could be used at least in a preliminary stage to perform a first estimate of the
cement content needed to treat a certain soil. This also indicates that for higher
acidity organic soils the cement content required to reach a pH of 12.6 may
greatly exceed the values used in this research.
The results of the tests for extraction of humic acid and fulvic acid from the
cement treated soils and the FT-IR tests conducted on the different fractions
obtained after extraction also offer the opportunity for drawing some broader
conclusions as to what organic soils may be more effectively treated using
cement. These tests demonstrate that certain acids (phenolic OH and carboxyl
acid) are more easily extracted from the cement-soil matrix, and thus soils rich in
these acids may be less amenable to be treated using cement or the treatment
may require a significant increase in the cement content
d) Applicability and advantages of deep mixing method for the improvement of
highly organic soils
As discussed earlier in this thesis, despite the successful application of deep
mixing methods for the improvement of inorganic soft soils, there has been,
especially in the US, some reluctance to using this method to address
construction on highly organic deposits, primarily due to concerns about the
interference of organic matter with the cement hydration process. The work
conducted for this research on LR soils clearly demonstrates that deep mixing
can be a viable option for construction on highly organic soils.
Moreover, the work presented in this research leads to important conclusions on
the practical advantages in the application of this technique to highly organic
soils in the field. It is, in fact, shown that treatment by cement not only reduces
the final settlement due to loading of the soil (already recognized), but also leads

410
to an increase in the rate of consolidation (thus mitigating stability and scheduling
problems associated with delayed consolidation) and a reduction in the creep
settlements (which for organic soil often represent the main design concern). In
particular the effect of the treatment on the secondary compression behavior has
gone largely unrecognized in the literature.
7.3. Recommendations for Future Research
This research investigated the effects of cement treatment on the 1-D
consolidation behavior of a highly organic soil and the changes in the organic
matter with cement treatment. The testing methods, test results obtained, and
the general framework proposed in this research provide significant insight on the
effects of cement treatment on the 1-D consolidation behavior of cement treated
organic soils and peats, and on the chemical interactions between organic
matter and cement. However, there exist numerous opportunities for continued
work. The recommendations for future research can be grouped into the
following five categories briefly discussed below:
a) additional investigation of the 1D consolidation behavior of LR soil;
b) investigation of the microstructure of cement treated LR soil;
c) investigation of the effects of cement treatment on the 1D consolidation
behavior of a broader range of organic soils;
d) assessment of the effects of the treatment on the strength properties;
e) evaluation of other binders for improving the engineering properties of
highly organic soils.
a) additional investigation of the 1D consolidation behavior of LR soil
The experimental program conducted on LR soil was ultimately limited by
time and resource availability, and there remain many interesting aspects of the
behavior of this soil both in its intact natural state, as well as following treatment
that warrant additional investigation.

411
With regard to the natural material, there remain questions in particular with
regard to the mechanisms responsible for the very high C/Cc of this soil. Further
investigation of the creep behavior of this soil is recommended.
With regard instead to the cement treated material, the following issues in
particular would benefit from additional work:
- effect of curing time: in this research the curing time was maintained constant
(14 days) in the majority of the tests and increased only up to 28 days in select
cases. While the results obtained for these two curing durations indicate that
this testing parameter has negligible influence on the 1D consolidation behavior,
this observation cannot, be extrapolated to longer curing periods. FT-IR
analyses conducted as part of this research on the natural soil indicate that the
organic matter of LR soil contains organic compounds, such as carboxyl acid,
phenolic OH and polysaccharides, which are known to act as retarders. This
suggests that for longer curing periods both the structure and the mechanical
properties will evolve further with time. In view of the above, it is recommended
that further tests, similar to those described in this thesis be conducted on the
treated soil after a more prolonged curing phase (90 days or more). This
approach should be applied to soil treated with 50% or more Portland cement,
i.e. for conditions that even under standard curing led to significant
improvement in the 1D consolidation behavior.
- structure development during creep: results were presented for tests conducted

on the cement treated specimens involving a creep phase in the normally


consolidated region (e.g. Figure 6.74). It was found that this creep phase gave
rise to a cementation induced preconsolidation stress greater than the creep
stress. The increased structure responsible for this preconsolidation stress was
found to be metastable and rapidly destroyed upon subsequent loading. This
behavior is clearly very different from that observed for the same cement
treated soil during the first loading stage in the CRS tests. The difference in the
type of structure generated as a result of curing outside the triaxial cell versus
that created as a result of aging in the normally consolidated region definitely

412
warrants additional work. The results are also likely to be relevant to
understanding aging and cementation processes in natural soils.
b) investigation of the microstructure of cement treated LR soil
In this research conclusions on the structure of both the natural and cement
treated LR soil with cement were mostly inferred from the consolidation tests as
well as from the results of the extraction tests and the chemical analyses. Only
for the natural soil was an attempt made, via cryogenic SEM (cryo-SEM), to
directly observe the structure of the soil (see chapter 3). Primarily due to time
constraints it was not possible to perform similar observations on the cement
treated soil.
It is recommended here that future work further explore the microstructure of the
cement treated soil using scanning electron microscopy (SEM), or, preferably,
given the significant water content that continues to characterize even the
cement treated soil, using an environmental SEM. Elemental analyses which are
today routinely combined with SEM/ESEM would provide additional insight into
the chemical makeup of the treated soil. Similarly, XRD analyses may be used
to determine the mineralogy of the new geomaterial generated by treatment of
LR soil with cement.
c) investigation of the effects of cement treatment on the 1D consolidation
behavior of a broader range of organic soils;
Considering the wide range of engineering properties of organic soils and peats,
and the effects of geological environment on the chemical and mineralogical
compositions of organic soils and peats, the author recommends the following,
as future research direction: Considering the wide range of engineering
properties of organic soils and peats, and the effects of geological environment
on the chemical and mineralogical composition of these soils, it is apparent that
the results from this study cannot be easily extrapolated to other soils. The

413
results of this research established specific patterns of behavior for one particular
soil treated with cement, and determined the conditions (pH and cement content)
required to significantly improve its properties. However, it remains unproven that
a similar improvement in properties can be achieved with any organic soil, and if
so under what conditions this can occur.
It appears therefore of great importance that future experimental work in this field
involve testing of a broader range of organic soils characterized by different
organic content, pH, and degree of humification.
The work should be founded on the characterization of the mechanical behavior
of the soil both in the natural (intact and reconstituted) state and treated state
and should be aimed at mapping the variation in engineering properties as a
function of the characteristics (e.g. organic content and degree of humification) of
the natural soil. As discussed above in b), efforts should be made to investigate
through SEM or ESEM the microstructure generated by treating the various soils
with cement.
Finally, this research has attempted to address the complexity of organic matter,
moving beyond the use of organic content and fiber content as the sole means to
characterize an organic soil. Future work should continue in this direction, and
include as an integral part of the experimentation a careful characterization of the
inorganic and organic fractions of both the natural soil and the cement treated
soil employing the types of methods (fractionation, elemental analyses, etc.)
used in this research (consistency in the organic matter fractionation and
extraction methods is critical since organic matter is highly sensitive to the
change in chemical environment). This will allow quantification of the various
organic compounds present, especially those that are known to act as retarders,
and lead to more fundamentally link the outcome of the treatment to the chemical
make-up of the soil. Note that beyond the interest in cement treatment of organic

414
soils, such a study would shed great insight into the factors and mechanisms
responsible for the behavior of natural organic soils.

d) assessment of the effects of the treatment on the strength properties


The work presented in this research focused exclusively on investigating the
effects of cement treatment on the 1D consolidation behavior of a highly organic
soil, while the impact of the treatment on the strength properties of the soil was
not addressed at all. This is, perhaps, the top priority for future research, as
highly organic soils pose not only settlement but also stability concerns and
design on these deposits requires consideration of both stability and settlement
constraints.
Evaluation of the shear behavior of the cement treated organic soils should rely
on consolidated undrained tests, following an approach similar to that used by
Kang (2007) for the study of the behavior of cement treated (inorganic) clays.
e) evaluation of other binders for improving the engineering properties of highly
organic soils;
Portland cement was used in this research as unconfined compression tests
performed by Humphrey (2001) on LR soil mixed with a few different binders
showed that the greatest improvement in strength was observed with this binder.
Field applications of deep mixing in inorganic soil often involve the use of other
binders such as fly ash and lime (often in combination with cement), and in some
countries (e.g. Sweden) there is considerable experience with the use of lime for
treating organic soils. These alternative binders should be considered in future
experimental work on deep mixing of organic soils.

LIST OF REFERENCES

415

LIST OF REFERENCES

Adams, J. (1965). The engineering behavior of a Canadian muskeg, Proc. 6th.


int. Conf. Soil. Mech. Found. Engrg., University of Toronto Press, Toronto,

Canada, 1, 3-7.
hnberg, H. (1996). Stress parameters of cement and line stabilized soils,
Grouting and Deep Mixing, vol1, p 387-392.

hnberg, H. (1996). Stress dependent parameters of cement and lime stabilized


soils. Grouting and Deep Mixing, Proceedings of IS-Tokyo96/ The second
International Conference on Ground Improvement Geosystems, Tokyo, 14-17
May 1996, 387-392.
hnberg, H. (1996). Measured permeabilities in stabilized Swedish soils,
Grouting and Ground Improvement, Proceedings of the Third International
Conference, Geotechnical Special Publication No. 120, Vol. 1. 622-633.
hnberg, H., and G. Holm. (1999). Stabilization of some Swedish organic soils
with different types of binder. Dry Mix Methods for Deep Soil Stabilization. Proc.
Of international conference on dry mix methods for deep soil stabilization., 13-15
Oct. 1999, Stockholm. Ed. Hkan Bredenberg, Gran Holm, and Bengt B. Broms.
Rotterdam: A.A. Balkema, 101-108.

416
Al-Shamrani, M.A. and Dhowian, A.W. (1997). Preloading for reduction of
compressibility characteristics of sabkha soil profiles, Engineering Geology, n 12, p 19-41.
American Society for Testing and Materials. (1999). 1999 Annual Book of ASTM
Standards. ASTM.
Amorosi, A. Rampello, S. (1998). The influence of natural soil structure on the
mechanical behaviour of a stiff clay, The geotechnics of hard soils - soft rocks
(eds Evangelista & Picarelli), p. 395-401
Al-Rawas, A.A. (2002). Microgabric and mineralogical studies on the
stabilization of an expansive soil using cement bypass dust and some types of
slags, Can. Geotech. J., 39 (5), 1150-1167.
Aziz, M.A. (1984). A New Method of Stabilization of Soft Soils. International
Conference on Case Histories in Geotechnical Engineering. Proceeding of
International Conference on Case Histories in Geotechnical Engineering. 6-11
May 1984, St. Louis. Ed. Shamsher Prakash. Rolla: University of Missouri-Rolla,
1215-1219.
Becker, D.E., Crooks, J.H.A., Been, K. and Jefferies, M.G. (1987). Work as a
criterion for determining in situ and yield stresses in clays, Can. Geotech. J., 24
(4), 549-564.
Berry,P.L. and Poskitt, T.J. (1972). The consolidation of peat, Geotchnique, 22
(1), 27 - 52
Berry,P.L. and Vickers, B. (1975). Consolidation of fibrous peat, Geotchnique,
17, 2-26.

417
Brady, N.C. and Weil, R.R. (1999). Elements of the Nature and Properties of
Soils. Prentice Hall.
Brandl, H. (1981), Alteration of soil parameters by stabilisation with lime,
Proceedings 10th International Conference on Soil Mechanics and Foundation
Engineering, Stockholm, Sweden, 587-594.

Broms, B.B (1999). Design of lime, lime/cement and cement columns, Dry Mix
Methods for Deep Soil Stabilization, Proceedings of the International Conference
on Dry Mix Methods for Deep Soil Stabilization, Stockholm, Sweden, 13-15 Oct,

1999, 125-153.
Broms, B. and Boman, P. (1977). Stabilization of soil with lime columns, Royal
Institute of Technology, Department of Soils and Rock Mechanics, Sweden,
Design Handbook.
Bruce, D.A. and Bruce, E.C. (2003). The practitioners guide to deep mixing,
Geotechnical Special Publication No. 120, ASCE, 474-488.
Buensuceso, B.R. (1990). Engineering behavior of lime treated soft Bangkok
Clay, Ph.D. thesis, Asian Institute of Technology, Bangkok.
Burland, J.B. (1990). On the compressibility and shear strength of natural clays,
Geotchnique, 46 (3), 329-378.

Burland, J.B., Rampello, S., Georgiannou, V.N., and Calabresi, G. (1996). A


laboratory study of the strength of four stiff clays, Geotchnique, 46 (3), 491-514.
Casagrande, A. (1948). Classification and Identification of soils, Transactions of
the American Society of Civil Engineers, 113, 901-930.

418
Chew, S.H., Kamuzzaman, A.H.M. and Lee, F.H. (2004). Physiochemical and
Engineering Behavior of Cement Treated Clays, J. Geotech. and Geoenviron.
Engrg., 130 (7), 696-706.

Colleselli, Francesco, Ciampaolo Cortellazzo, and Simonetta Cola (2000).


Laboratory Testing of Italian Peaty Soils. Geotechnics of High Water Content
Materials: ASTM Special Technical Publication 1374. Ed. Tuncer B. Edil and
Patrick Fox. West Conshohocken: ASTM, 226-240.
Collepardi, M.M. (1995). Chapter 6. Water Reducers/Retarders in Concrete
Admixtures Handbook: Properties, Science, and Technology, Second edition.
Noyes Publications.
Conner, J.R. (1990). Chemical Fixation and Solidification of Hazardous Wastes.
Van Nostrand Reinhold.
Cook, P.M. (1956). Consolidation Characteristics of Organic Soils. Proc. Ninth.
Can. Soil. Mech. Conf., 82-87.
Cortellazzo, G. and Cola, S. (1999). Geotechnical characteristics of two Italian
peats stabilized with binders, Dry Mix Methods for Deep Soil Stabilization,
Proceddings of the International Conderence on Dry Mix Methods for Deep Soil
Stbilization, Stockholm, Sweden, 13-15 Oct. 1999, 93-100.
Den Haan, E.J. (1996). A compression model for non-brittle soft clays and peat.
Gotechnique., 46 (1), 1-16.

Deng, Y. and Dixon, J.B. (2002) Soil Organic Matter and Organic-Mineral
Interactions, Soil Mineralogy with Environmental Applications, SSSA Book
series No. 7. Soil Science Society of America, Inc. 69-107.

419
Dhowian, A.B., and Edil, E.B. (1980). Consolidation Behavior of Peats, Geotech.
Testing J., 3 (3), 105-114.

Earth Exploration, INC. (1993). Preliminary geotechnical evaluation: Proposed


improvements to Lindberg road in West Lafayette, Tippecanoe County, Indiana
Edil, Tuncer B. and Xiaodong Wang. (2000). Shear Strength and Ko of Peats
and Organic Soils. Geotechnics of High Water Content Materials: ASTM Special
Technical Publication 1374. Ed. Tuncer B. Edil and Patrick Fox. West
Conshohocken: ASTM, 209- 225.
Esrig, M.I. (1999). Keynote lecture: Properties of binders and stabilized soils.
Dry Mix Methods for Deep Soil Stabilization. Proc. of international conference on
dry mix methods for deep soil stabilization., 13-15 Oct. 1999, Stockholm. Ed.
Hkan Bredenberg, Gran Holm, and Bengt B. Broms. Rotterdam: A.A. Balkema,
67-72.
Esrig, M.I., Mac Kenna, P.E., and Forte, E.P. (2003) Ground stabilization in the
United States by the Scandinavian lime cement dry mix process, Geotechnical
Special Publication No. 120, ASCE, 501-514.
Federal Highway Administration. (2000). An introduction to the deep mixing
method as used in geotechnical applications, Prepared by Geosystems, L.P.,

FHWA-RD-99-138.
Filz, G.M., Hodges, D.K., Weatherby, D.E., and Marr, W.A. (2005). Standard
definition and laboratory procedure for soil-cement specimens applicable to the
wet method of deep mixing, Proc. Geo-Frontier 2005 Congress on Innovation in

Grouting and Soil Improvement, Geotechnical Special Publication, No. 136,


ASCE, Austin, TX, p1643-1655.

420
Fontana, P., Lagioia, R., and Nova, R. (1998). Subsidence and wellbore stability
during the excavation of oil wells, The geotechnics of hard soils -= soft rocks
(eds Evangelista & Picarelli), p. 519-527
Force, E.A. (1998). Factors controlling pore pressure generation during Ko
consolidation of laboratory tests MS Thesis, MIT Dept. of Civil and Env. Eng.,
Cambridge, MA
Fox, P. J. and Edil, T.B. (1996). Effects of stress and temperature on secondary
compression of peat. Can. Geotech. J., 33, 405-415.
Fox, P. J., Edil, T.B., and Lan, L.T. (1992). C/Cc concept applied to
compression of Peat. J. Geotech. Engr., 118 (8), 1256-1263.
Fox, P. J., Roy-Chowdhury, N., and Edil, T.B.. (1999). Secondary Compression
of Peat With or Without Surcharging. J. Geotech. and Geoenviron. Engrg., 125
(2), 160-162.
Ghazali, F.M., Baghdadi, Z.A., and Khan, A.M. (1991). Overconsolidated
behavior of phosphoric acid and lime-stabilized kaolin clay, Transportation
Research Record 1295, 62-69.
Gruen, H.A. and Lovell, C.W. (1984). Compression of peat under embankment
loading, Transportation Research Record, p 56-59.
Hanrahan, E.T. (1954). An investigation of some physical properties of peat,
Gotechnique., 4 (2), 108-123.

Harpstread, M.I. and Hole, F.D. (1980). Soil Science Simplified. Iowa State
University Press, Ames, IOWA, U.S.A.

421
Hebib, S., and E.R. Farrell. (1999). Some experience of stabilizing Irish organic
soils. Dry Mix Methods for Deep Soil Stabilization. Proc. of international
conference on dry mix methods for deep soil stabilization., 13-15 Oct. 1999,
Stockholm. Ed. Hkan Bredenberg, Gran Holm, and Bengt B. Broms.
Rotterdam: A.A. Balkema, 81-84.
Hebib, S. and Farrel, E.R. (2003). Some experiences on the stabilization of Irish
peats, Can. Geotech. J., 40 (1), 107-120.
Hebib, S. and Farrel, E.R. (2003). Deep soil mixing for organic soils, Engineers
Journal, 56 (2).

Hobb, N.B. (1986). Mire morphology and the properties and behavior of some
British and foreign peats. Q. J. Eng. Geol., London, V.19, No.1, 7-80.
Holm, Goran. (2003). State of practice in dry deep mixing methods,
Geotechnical Special Publication, n 1201, p 145-163.

Humphrey, A.M. (2001). Improvement of a soil with considerable organic


content, MS Thesis, Purdue University
Huttunen, E. and K. Kujala. (1996). On the stabilization of organic soils.
Grouting and Deep Mixing. Proc. of IS-Tokyo 96/ The second international
conference on ground improvement geosystems., 14-17 May 1996, Tokyo. Ed.
Ryozo Yonekura, Masaaki Terashi, and Mitsuhiro Shibazaki. Rotterdam: A.A.
Balkema., 411-414.

422
Huttunen, E., K. Kujala, and H. Vesa. (1996). Assesment of the quality of
stabilized peat and clay. Grouting and Deep Mixing. Proc. of IS-Tokyo 96/ The
second international conference on ground improvement geosystems., 14-17
May 1996, Tokyo. Ed. Ryozo Yonekura, Masaaki Terashi, and Mitsuhiro
Shibazaki. Rotterdam: A.A. Balkema., 607- 612.
Indiana Department of Transportation. (1999). Division of Contracts and
Construction Standards Section. 1999 Standards. Indianapolis: Indiana.
Janbu, N., Tokheim, O. and Semmeset, K. (1981). Consolidation tests with
continuous loading. 10th Int. Conf. Soil Mechanics & foundation Eng. Stockholm,
Vol. 1. 645-654.
Jameson, R. N. J., (1996). Mechanical Properties of Soil Cement and
Applications in Excavation Support. Unpublished MAs Thesis, MIT.
Jamiolkowski, M., Ladd, C.C., Germaine, J.T. and Lancellotta, R. (1985). "New
developments in field and laboratory testing of soils" Proc. 11th Int. Conference
on Soil Mech. and Found. Engrg., San Francisco, USA, 1, 57-153.

Jelisic, Nenad and Leppanen, Mikko (2003). Mass stabilization of organic soils
and soft clay, Geotechnical Special Publication, n 1201, p 552-561.
Johnston, C.T. and Tombcz, E. (2002) Surface Chemistry of Soil Minerals,
Soil Mineralogy with Environmental Applications, SSSA Book series No. 7. Soil
Science Society of America, Inc. 37-67.
Jolicoeur, C. and Simard, M.A. (1998). Chemical Admixture-Cement
Interactions: Phenomenology and Physio-chemical Concepts, Cement and
Concrete Composites, 20, 87-101.

423
Kang, Y. (2007). A fundamental study on the mechanical behavior of a cementtreated clay, Ph.D. thesis, in preparation, Purdue University.
Kassim, K.A. and Clarke, B.G. (1999). Constant rate of strain consolidation
equipment and procedure for stabilized soils, Geotech. Testing J., 22, 13-21.
Keene, P., and Zawodniak, C.D. (1968). Embankment construction on peat
utilizing hydraulic fill, Proc. 3rd. Int. Peat Contr., National Research Council of
Canada, Ottawa, Canada, 45-50.

Koda, E., Szymanski, A., and Wolski, W. (1993), Field and laboratory
experience with the use of strip drains in organic soils, Canadian Geotechnical
Journal, v 30, n2, p 308-318.
Koda, E., Wolski, W. (1994). The influence of strip drains on the consolidation
performance of organic soils. Advances in Understanding and Modelling the
Mechanical Behavior of Peat. Ed. Evert den Haan, Ruud Termaat & Tuncer B.
Edil. A.A. Balkema: Rotterdam, 347-359.
Kogure, K., Yamaguchi, H., Ohira, Y., and Ono, H. (1986). Experiments on
Consolidation Characteristics of a Fibrous Peat. Proceedings Advances in
Peatlands Engineering, Carleton University, August 25-26, 1986.
Kononova, M.M. (1966). Soil Organic Matter: Its Nature, Its Role in Soil
Formation and in Soil Fertility. 2nd English edition Translated from the Russian by
Nowakowski, T.Z. and Newman, A.C.D. Pergamon Press

424
Kirov, B. (1994). Experience of peat preloading in the Varna West Harbour.
Advances in Understanding and Modelling the Mechanical Behavior of Peat. Ed.
Evert den Haan, Ruud Termaat & Tuncer B. Edil. A.A. Balkema: Rotterdam, 341345.
Kujala, K., M. Mkikyr, and O. Lehto. (1996). Effect of humus on the binding
reaction in stabilized soil. Grouting and Deep Mixing. Proc. of IS-Tokyo 96/ The
second international conference on ground improvement geosystems., 14-17
May 1996, Tokyo. Ed. Ryozo Yonekura, Masaaki Terashi, and Mitsuhiro
Shibazaki. Rotterdam: A.A. Balkema., 415-420.
Ladd, C.C., Young, G.A., Kraemer, S.R., and Burke, D.M. (19991). Engineering
properties of Boston Blue Clay from special testing program, Geotechnical
Special Publication, n 91, 1999, 1-24.

Lahtinen, P., H. Jyrv, and K. Kuusipuro. (1999). Development of binders for


organic soils. Dry Mix Methods for Deep Soil Stabilization. Proc. of international
conference on dry mix methods for deep soil stabilization., 13-15 Oct. 1999,
Stockholm. Ed. Hkan Bredenberg, Gran Holm, and Bengt B. Broms.
Rotterdam: A.A. Balkema, 109-114.
Lambe, William T., and Robert V. Whitman. (1969). Soil Mechanics. New York:
John Wiley & Sons.
Lambrechts, J.R., Ganse, M.A, and Layhee, C.A. (2003). Soil mixing to stabilize
organic clay forI-95 Widening, Alexandria, VA. Geotechnical Special Publication
No. 120, ASCE, 575-585.

425
Landva, A.O., Korpojaakko, E.O., and Pheeney, P.E. (1982). Geotechnical
Classification of Peats and Organic Soils. Testing of Peats and Organic Soils.
ASTM Special Testing Publication 820. 37-51
Landva, Arvid O., Peter E. Pheeney, and Donald E. Mersereau. (1983).
Undisturbed Sampling of Peat. Testing of Peats and Organic Soils: ASTM
Special Technical Publication 820. Proc. of a symposium sponsored by ASTM
committee D-18 on Soil and Rock., 23 June 1982, Toronto. Ed. P.M. Jarrett.
Philadelphia: ASTM, 141-156.
Landva, A.O., and P.E. Pheeney. (1980). Peat Fabric and Structure. Can.
Geotech. J., 17, 416-435

Lea, F.M. (1956). The Chemistry of Cement and Concrete, St. Martins press
Inc.
Lea, N.D. and Brawner, C.O. (1963). Highway design and construction over peat
deposits in lower British Columbia Hwy Res. Rec. 7, 1-32.
Lefebvre, G., Langlois, P, Lupien, V. and Lavalle, J.G. (1984). Laboratory
Testing and in situ Behavior of Peat as Embankment Foundation, Can. Geotech.
J., 21 (2), 322-337.

Leroueil S., Tavenas F., Samson L. and Morin P. (1983). Preconsolidation


pressure of Champlain clays. Part II: Laboratory determination, Can Geotech. J.,
20 (4), 803-816.

Lewis, W.A. (1956). The settlement of the approach embankment to a new road
bridge at Lockford, West Suffolk, Gotechnique., 6 (3), 106-114.

426
Liu, M.D. and Carter, J.P. (1999). Virgin compression of structured soils,
Geotchnique, 49 (4), 43-57.

Liu, M.D. and Carter, J.P. (2000). Modeling the destructuring of soils during
virgin compression, Geotchnique, 50 (4), 479-483.
Liu, M.D. and Carter, J.P. (2002). A structured Cam Clay model, Can Geotech.
J., 39 (6), 1313-1332.

Locat, J. Tremblay, H. and Leroueil, S. (1996). Mechanical and hydraulic


behavior of a soft inorganic clay treated with lime, Can Geotech. J., 33 (4), 654669.
Lorenzo, G.A., and Bergado, D.T. (2004) Fundamental parameters of cementadmixed clay: New approach, Journal of Geotechnical and Geoenvironmental
Engineering, ASCE, v. 130, n 10, p 1042-1050.

Maccarini, M. (1987). Laboratory investigations on artificial soil. Ph.D. thesis,


University of London.
MacFarlane, I.C. (1969). Muskeg Engineering Handbook. Ed. Ivan C.
MacFarlane. Toronto, University of Toronto Press.
MacFarlane, I.C. and Williams, G.P. (1974). Some Engineering Aspects of Peat
Soils, Histosols: Their Characteristics, Classification, and Use. SSSA Special
Publication Series No. 6. 79-93.
McCallister, L.D., and Petry, T.M. (1992). Leach tests on lime-treated clays,
Geotech. Testing J., 15, 106-114.

427
Mesri, G. (1973). Coefficient of Secondary Compression, J. Geotech. and
Geoenviron. Engrg., 99 SM1, 123-137.

Mesri, G. and Casto, A. (1987). C/Cc concept and Ko during secondary


compression, J. Geotech.. Engrg., 113 (3), 230-247.
Mesri, G. and Godlewski, P.M. (1977). Time- and Stress-Compressibility
Interrelationship. J. Geotech.. Engrg., 103, GT5, 417 - 430.
Mesri, G., Stark, M.D, Ajlouni, M.A. and Chen, C.S. (1997). Secondary
Compression of Peat with or Without Surcharging, J. Geotech. and Geoenviron.
Engrg., 123 (5), 411- 421.

Mesri, G., Stark, M.D, and Chen, C.S. (1994). C/Cc Concept Applied to
Compression of Peat-Discussion, J. Geotech. and Geoenviron. Engrg., 120 (4),
764-767.
Mesri, G. and Rokhsar, A. (1974). Theory of Consolidation For Clays, J.
Geotech. and Geoenviron. Engrg., 100, GT8, 889-904.

Nalbantoglu, Z. and Tuncer E. R. (2001). Compressibility and hydraulic


conductivity of a chemically treated expansive clay, Can Geotech. J., 38 (1),
154-160.
National Resources Conservation Service (NRCS). (2004)
HTTP://WWW.FTW.NRCS.USDA.GOV/OSD/DAT/HOUGHTON.HTML

428
Ng, Samuel Y., and James C. Rudd. (1984). Embankments Built Over Swamps.
International Conference on Case Histories in Geotechnical Engineering.
Proceeding of International Conference on Case Histories in Geotechnical
Engineering. 6-11 May 1984, St. Louis. Ed. Shamsher Prakash. Rolla: University

of Missouri-Rolla, 609-618.
Nicholson, P.J. and Chu, E.K. (1994). Boston Central Artery tunnel excavation
support remediation by ground treatment, ASCE conference on Soil Structure
Interaction and Environmental Issues, Hershey, PA, September 12-14.

ORourke, T.D. and McGinn, A.J. (2004). Case history of deep mixing soil
stabilization for Boston Central Artery, Geotechnical Special Publication, n 1261,
Geotechnical Engineering for Transportaion Projects: Proceddings of Geo-Trans
2004, p 77-136.

Rao, S.N and Rajasekaran, G. (1996). Reaction Products Formed in LimeStabilized Marine Clays, J. Geotech.. Engrg., 122 (5), 329-336.
Russel, J.B., Lawrence, C.A., Paddock, M.W., and Ross, D. (1999) Monitoring
preload performance, ASTM Special Technical Publication, n 1358, p 236-249.
Ryan C.R. and Jasperse, B.H. (1989). Deep soil mixing at Jackson Lake Dam,
Proc. ASCE 1989 Foundation Engineering Congress, Foundation Engineering:
Current Principles and Practices, v.1, p 354-367.

Tan, K.H. (2003). Humic Matter in Soil and the Environment: Principles and
Controversies. Marcel Dekker, Inc.
Tan, K.H. (2005). Soil Sampling, Preparation, and Analysis. Second edition.
Taylor & Francis.

429
Terashi, M. (1997). Deep mixing method brief state of the art. 17th
International Conference on Soil Mechanics and Foundation Engineering,
Hamburg, Germany, September, 4.
Pollard, S.J.T., Montgomery, D.M., Sollars, C.J. and Perry, R. (1991). Organic
compounds in the cement-based stabilization/solidification of hazardous mixed
wastes-Mechanistic and process considerations, Journal of Hazardous Materials,
28, 313-327.

Pousette, K., J Mcsik, A. Jacobson, R. Andersson, and P. Lahtinen. (1999).


Peat soil samples stabilized in laboratory Experiences from manufacturing and
testing. Dry Mix Methods for Deep Soil Stabilization. Proc. of international
conference on dry mix methods for deep soil stabilization., 13-15 Oct. 1999,
Stockholm. Ed. Hkan Bredenberg, Gran Holm, and Bengt B. Broms.
Rotterdam: A.A. Balkema, 85-92.
Ramachandran, V.S. (1995). Concrete Admixtures Handbook: Properties,
Science and Technology. Second edition. Noyes Publications.
Rao, S.N. and Rajasekaran, G. (1996) Reaction products formed in limestabilized marine clays, J. Geotech. Engrg., 122 (5), 329-336.
Radforth, N.W. (1969). Classification of Muskeg. Muskeg Engineering
Handbook. Ed. Ivan C. MacFarlane. Toronto, University of Toronto Press, 31-39.
Samson, L., and LaRochelle, P. (1972). Design and performance of an
expressway construicted over peat by preloading, Can. Geotech. J., Ottawa,
Canada, 9(4), 447-466.

430
Santangata, M.C. (1994). Investigation of Sample Disturbance in Soft Clays
Using Triaxial Element Tests, S.M. Thesis, Department of Civil and
Environmental Engineering, MIT, Cambridge, MA, 289p.
Sasaki, Harumi. (1985). Effectiveness and applicability of the methods of
foundation improvement for embankments over peat deposits. Recent
Developments in Ground Improvement Techniques. Proc. of Symposium on
Recent Developments in Ground Improvement Techniques ., 29 Nov. 3 Dec.
1982, Bangkok. Ed. A.S. Balasubramaniam. Boston: A.A. Balkema, 543-561.
Schnitzer, M. and Khan, S.U. (1972). Humic Substances in the Environment.
Marcel Dekker, Inc.
Schnitzer, M. and Khan, S.U. (1978). Soil Organic Matter. Elsevier Scientific
Publishing Company.
Schulze, D.G. (2002) An Introduction to Soil Mineralogy, Soil Mineralogy with
Environmental Applications, SSSA Book series No. 7. Soil Science Society of
America, Inc. 1-35.
Sheahan, T.C. and Watters, P.J. (1997) Experimental Verification of CRS
Consolidation Theory, J. Geotech. and Geoenviron. Engrg., 123 (5), 430 437.
Soil Taxonomy: A Basic System of Soil Classification for Making and Interpreting
Soil Surveys. (1999). Second edition. United States Department of Agriculture
Natural Resources Conservation Service, Agriculture Handbook Number 436.
Sparks, D.L. (2003). Environmental Soil Chemistry, Second edition. Academic
Press.

431
Spencer, R.D. (1993). Chemistry and Microstructure of Solidified Waste Forms.
Lewis Publishers
Stevenson, F.J. (1994) Humus Chemistry: Genesis, Composition, Reactions. 2nd
edition. John Wiley & Sons, Inc.
Terashi, M. Tanaka, H. Niidome, Y., and Sakanoi, H. (1980). Permeability of
treated soils, Proceedings 15th Japan Conference on Soil mechanics and
Foundation Engineering, 773-776.
Terashi, M. and Tanaka, H. (1983). Settlement analysis for deep mixing
methods, Proceedings 8th Eur. Conf. on Soil Mech. and Found. Eng., Helsinki,
1983, p 955-960.
Terzaghi, K., Peck, R.B. and Mesri, G. (1996). Soil Mechanics in Engineering
Practice, 3rd edition, John Wiley & Sons, Inc.
Townsend, D.C. and Klym, T.W. (1996). Durability of lime-stabilized soils,
Highway Research Board Bulletin 139, 25-41.
Tremblay, H, Leroueil, S. and Locat, J. (2001). Mechanical improvement and
vertical yield stress prediction of clayey soils from eastern Canada treated with
lime or cement, Can Geotech. J., 38, 567-579.
Tremblay, H, Duchesne, J., Locat, J., and Leroueil, S. (2002). Influence of the
nature of organic compounds on fine soil stabilization with cement, Can Geotech.
J., 39, 535-546.

Unites States Department of Agronomy (USDA). (2003). Keys to soil taxonomy,


ninth edition.

432
Unites States Department of Agronomy (USDA). (1998). Soil survey of
Tippecanoe County, Indiana.
Venema, Thomas P., C.R. Waletzko, and Frank J. Swekosky. (1990). Ground
Modification for a 14-story Structure over a Deep Peat Deposit. Hayward Baker.
Ground Modification Case Histories. Odenton: Hayward Baker.
Walmsley, M.E. (1973) Physical and Chemical Properties of Peat, Muskeg and
the Northern Environment in Canada. Ed. N.W. Radforth and C.O. Brawner,

University of Toronto Press, 82-129.


Weber, W.G. (1969). Performance of embankments constructed over peat, J.
Soil. Mech. And. Found. Div., ASCE, 95(1), 53-76.

Wissa, A E. Z., Christian, J. T., Davis, E.H. and Heiberg, S. (1971).


Consolidation at Constant Rate of Strain, J. Soil Mech. and Found Div., ASCE,
97 (10), 1393-1413.

VITA

433

VITA

Joonho Hwang was born in Seoul, Republic of Korea, on July 23, 1972, as the
second son of Chung Soo Hwang and Chung Sun Lee. After completing his
degree at Pohang High School, Pohang, Republic of Korea, in 1991, he entered
the Yonsei University, Seoul, Republic of Korea, receiving a Bachelor of Science
degree in February, 1997. He entered the Graduate School in the Department of
Earth and Atmospheric Sciences at Purdue University, Indiana, in August of 1999,
receiving a Master of Science degree in December, 2001. He then entered the
School of Civil Engineering, Purdue University, Indiana, in January 2002 for his
Ph.D. He is married to Yoon Kyung Kim, with whom he had a daughter.

You might also like