Вы находитесь на странице: 1из 277

Copyright

by

Lianxiang Du

2001
Laboratory Investigations of Controlled Low-Strength Material

by

Lianxiang Du, B.S., M.S.C.E.

Dissertation
Presented to the Faculty of the Graduate School of

The University of Texas at Austin

in Partial Fulfillment

of the Requirements

for the Degree of

Doctor of Philosophy

The University of Texas at Austin


August, 2001
UMI Number: 3031045

Copyright 2001 by
Du, Lianxiang

All rights reserved.

________________________________________________________
UMI Microform 3031045
Copyright 2001 Bell & Howell Information and Learning Company.
All rights reserved. This microform edition is protected against
unauthorized copying under Title 17, United States Code.
____________________________________________________________

Bell & Howell Information and Learning Company


300 North Zeeb Road
P.O. Box 1346
Ann Arbor, MI 48106-1346
The Dissertation Committee for Lianxiang Du

Certifies that this is the approved version of the following dissertation:

Laboratory Investigations of Controlled Low-Strength Material

Committee:

Kevin J. Folliard, Supervisor

David W. Fowler

David Trejo

Ellen Rathje

Alan F. Rauch
Dedication

This dissertation is dedicated to my parents, my wife, Maohui Zhou, and other

people who encouraged and supported me during my study in China and in US.

This dissertation is especially dedicated to my dad, who always stayed late with

me until I finished my assignments when I was a kid.


Acknowledgements

This report would not have been possible without Dr. Kevin J. Folliard.

He spent countless time advising, discussing, and encouraging me. He opened

my eyes to the world of concrete technology. He was so patient that he spent

many hours with my English skills. I owe him so much.

Special gratitude goes to Dr. David Trejo at Texas A&M University, who

co-sponsored the research work presented in this dissertation. His spiritual

inspirations will benefit my whole life. Longhorns and Aggies can be not only

rivals, but also friends.

I also appreciate very much the cooperation and support of my other

committee members, Dr. David Fowler, Dr. Ellen Rathje, and Dr. Alan Rauch.

Their guidance and technical support were very valuable to me. Special thanks

also goes to Dr. Dov Leshchinsky at University of Delaware. He was always

there providing answers to issues in geotechnical engineering. The involving of

Lilma dos Santos Ribeiro and Jessica D. Reath in part of the research are

appreciated. I would also like to thank members of the Construction Materials

Research Group, specifically Mike Rung, Kerry Rothenberg, David Whitney,

v
Sherian Williams, and many undergraduate students who helped me with the

laboratory work.

vi
Laboratory Investigations of Controlled Low-Strength Material

Publication No._____________

Lianxiang Du, Ph.D.

The University of Texas at Austin, 2001

Supervisor: Kevin J. Folliard

This report presents the results of experimental studies on the properties of

controlled low-strength material (CLSM) and the corresponding test methods for

evaluating the characteristics of this material. A wide range of CLSM mixtures

were included in this study.

Fresh properties of CLSM were evaluated. The use of bottom ash,

foundry sand, and high-carbon fly ash was found to increase water demand to

achieve the desired flow in CLSM. Bottom ash tended to increase bleeding,

whereas foundry sand decreased bleeding tendencies. Two methods were used to

evaluate the setting and hardening of fresh CLSM mixtures, and their mechanisms

were evaluated. New methods were developed to measure the segregation and

subsidence of fresh CLSM mixtures.

The hardened properties of CLSM, including compressive strength,

excavatability, California Bearing Ratio, resilient modulus, water permeability,

vii
triaxial shear strength, and drying shrinkage were studied. Predictive models

were proposed for the strength development of CLSM mixtures. Modifications to

ASTM D 4832 were proposed to make it a more appropriate method for

measuring of the compressive strength of CLSM. The effects of alternative

capping materials, including neoprene pads, gypsum and sulfur, were studied. In

addition, factors affecting the strength gain of CLSM mixtures, including curing

temperature and humidity, were examined.

The resistance of CLSM mixtures to freezing and thawing was evaluated

using modified version of ASTM D 560. Higher air contents and compressive

strengths were found to improve the durability of CLSM mixtures exposed to

freezing-and-thawing cycles.

Although CLSM was found less corrosive than soil, the potential

formation of galvanic cells may be a concern to pipeline engineers. Factors

influencing the corrosion behaviors of ductile-iron samples were identified. For

coupled and uncoupled corrosion conditions, different contributing factors and

mechanisms were identified.

viii
Table of Contents

LIST OF TABLES XVI

LIST OF FIGURES XVIII

CHAPTER 1: INTRODUCTION 1

1.1 GENERAL ........................................................................................................ 1

1.2 RESEARCH OBJECTIVES .............................................................................. 3

1.3 OUTLINE OF THE DISSERTATION ............................................................. 4

CHAPTER 2: LITERATURE REVIEW 6

2.1 INTRODUCTION............................................................................................. 6

2.2 MATERIALS .................................................................................................... 9


2.2.1 Bonding Materials ................................................................................. 9
2.2.1.1 Portland Cement ........................................................................ 9
2.2.1.2 Fly ash ....................................................................................... 9
2.2.2 Aggregate ............................................................................................ 11
2.2.2.1 Concrete Aggregate ................................................................. 11
2.2.2.2 Foundry Sand ........................................................................... 12
2.2.2.3 Excess Fines from Aggregate Crushing Operation................. 13
2.2.2.4 Bottom Ash.............................................................................. 14
2.2.2.5 Other Aggregate Materials ...................................................... 15
2.2.3 Admixtures .......................................................................................... 15
2.2.3.1 Mineral Admixtures................................................................. 15
2.2.3.2 Chemical Admixtures .............................................................. 17
2.2.4 Water ................................................................................................... 17

ix
2.3 MIXTURE PROPORTIONS........................................................................... 17

2.4 BATCHING, MIXING, AND TRANSPORTATING .................................... 20

2.5 CLSM PROPERTIES AND TEST METHODS ............................................. 21


2.5.1 Fresh CLSM Properties ....................................................................... 22
2.5.1.1 Flowability............................................................................... 22
2.5.1.2 Segregation and Bleeding........................................................ 25
2.5.1.3 Hardening Time ....................................................................... 26
2.5.1.4 Subsidence ............................................................................... 27
2.5.2 Hardened CLSM Properties ................................................................ 27
2.5.2.1 Compressive Strength.............................................................. 27
2.5.2.2 Permeability............................................................................. 29
2.5.2.3 Shear Strength ......................................................................... 31
2.5.2.4 California Bearing Ratio (CBR) and Resilient Modulus......... 31
2.5.2.5 Drying Shrinkage ..................................................................... 33
2.5.2.6 Thermal Conductivity.............................................................. 34
2.5.2.7 Excavatability.......................................................................... 35
2.5.2.8 Freezing-and- Thawing Resistance .......................................... 36
2.5.2.9 Erosion..................................................................................... 37
2.5.2.10 Leaching and Environmental Impact..................................... 38
2.5.2.11 Corrosion............................................................................... 39
2.5.2.12 Utility Degradation................................................................ 41

2.6 ECONOMICS.................................................................................................. 41

2.7 CURRENT PRACTICE .................................................................................. 41


2.7.1 Case Histories ...................................................................................... 41
2.7.1.1 Introduction ............................................................................. 41
2.7.1.2 Backfills................................................................................... 44
2.7.1.3 Bridge Replacement ................................................................ 47
2.7.1.4 Structural Fill ........................................................................... 48

x
2.7.1.5 Utility Bedding ........................................................................ 51
2.7.1.6 Erosion Control ....................................................................... 57
2.7.1.7 Void Fill................................................................................... 57
2.7.1.8 Environmental Protection........................................................ 59

2.8 RESEARCH NEEDS ...................................................................................... 59


2.8.1 Introduction ......................................................................................... 59
2.8.2 Research Needs ................................................................................... 60
2.8.2.1 Standard Test Methods............................................................ 60
2.8.2.2 Effective Use of By-Product Materials in CLSM ................... 61
2.8.2.3 Excessive Long-Term Strength Gain ...................................... 61
2.8.2.4 Durability Aspects of CLSM ................................................... 62
2.8.2.5 Environmental Impact ............................................................. 62
2.8.2.6 Applicability of Soil Test Methods ......................................... 63

CHAPTER 3: MATERIALS AND MIXTURE PROPORTIONS 64

3.1 INTRODUCTION........................................................................................... 64

3.2 FINE AGGREGATE ....................................................................................... 64


3.2.1 Aggregate Type ................................................................................... 64
3.2.2 Aggregate Gradation ........................................................................... 65
3.2.3 Absorption Capacity and Specific Gravity of Fine Aggregate............ 66
3.2.4 Chemical Analysis of Foundry Sand ................................................... 68

3.3 MINERAL ADMIXTURES ............................................................................ 68


3.3.1 Fly ash ................................................................................................. 68
3.3.2 Chemical Analysis of Foundry Sand ................................................... 69

3.4 CHEMICAL ADMIXTURES ......................................................................... 70


3.4.1 Air-Entraining Agent ........................................................................... 70
3.4.2 Accelerating Agent .............................................................................. 71

xi
3.5 CEMENT......................................................................................................... 71

3.6 MIXING WATER ........................................................................................... 71

3.7 CLSM MIXTURE PROPORTIONS............................................................... 72


3.7.1 Mixture Proportioning ......................................................................... 72
3.7.2 Mixture Proportions ............................................................................. 74

CHAPTER 4: TEST METHODS AND PROCEDURES 82

4.1 INTRODUCTION........................................................................................... 82

4.2 FRESH CLSM TEST METHODS .................................................................. 85


4.2.1 Mixing Procedure and Measurement of Flow, Air Content, and
Unit Weight ......................................................................................... 85
4.2.2 Setting and Hardening ......................................................................... 87
4.2.3 Segregation.......................................................................................... 89
4.2.4 Subsidence ........................................................................................... 90

4.3 HARDENED PROPERTIES OF CLSM......................................................... 91


4.3.1 Unconfined Compressive Strength...................................................... 91
4.3.1.1 Specimen Preparation.............................................................. 91
4.3.1.2 Compression Machine ............................................................. 93
4.3.1.3 Effects of Loading Rate on Compressive strength.................. 95
4.3.1.4 Cylinder Curing and Conditions .............................................. 96
4.3.1.5 Effects of Curing Temperature and Humidity on
Compressive Strength................................................................. 98
4.3.1.6 Effects of Drainage Conditions on Compressive Strength...... 99
4.3.1.7 Alternative Capping Materials for Compression Testing...... 101
4.3.2 Triaxial Shear Strength...................................................................... 103
4.3.3 Water Permeability of CLSM............................................................ 104
4.3.4 Drying Shrinkage of CLSM mixtures ............................................... 105
4.3.5 California Bearing Ratio .................................................................... 106
4.3.6 Resilient Modulus.............................................................................. 106

xii
4.3.7 Excavation Study by CLSM Mixtures .............................................. 107

4.4 DURABILITY OF CLSM............................................................................. 109


4.4.1 Freezing-and- Thawing Resistance .................................................... 109
4.4.2 Wetting-and-Drying Resistance of CLSM ........................................ 110

4.5 CORROSION OF DUCTILE IRON IN CLSM ............................................ 110


4.5.1 Research Approach............................................................................ 110
4.5.2 Mass Loss Test .................................................................................. 111
4.5.2.1 Uncoupled Condition............................................................. 111
4.5.2.2 Coupled Condition (Galvanic) .............................................. 113
4.5.3 pH of Solutions Extruded from CLSM ............................................. 113
4.5.4 The Resistivity of CLSM................................................................... 115
4.5.5 Chloride Diffusion............................................................................. 116

4.6 LEACHING AND ENVIRONMENTAL IMPACT..................................... 116

CHAPTER 5: RESULTS AND DISCUSSIONS 120

5.1 INTRODUCTION......................................................................................... 120

5.2 CLSM MIXTURE PROPORTIONS AND FRESH PROPERTIES ............. 120


5.2.1 Water Demand/Flow Behavior .......................................................... 120
5.2.2 Air Contents of Fresh CLSM Mixtures............................................. 124
5.2.3 Unit Weight of Fresh CLSM Mixtures.............................................. 125
5.2.4 Setting/Hardening.............................................................................. 125
5.2.5 Bleeding of Fresh CLSM Mixtures ................................................... 133
5.2.6 Segregation of Fresh CLSM Mixtures .............................................. 134
5.2.7 Subsidence ......................................................................................... 136

5.3 HARDENED PROPERTIES OF CLSM....................................................... 137


5.3.1 Unconfined Compressive Strength.................................................... 137
5.3.1.1 Compressive Strength Development ..................................... 137
5.3.1.2 Effects of Cylinder Curing on Compressive Strength........... 146

xiii
5.3.1.3 Effects of Drainage Conditions on Compressive Strength.... 154
5.3.1.4 Specimen Preparation before Testing.................................... 157
5.3.1.5 Alternative Capping Materials for Compression Testing...... 159
5.3.1.6 Effects of Load Rate on Compressive Strength .................... 168
5.3.1.7 Effect of Specimen Size ........................................................ 172
5.3.1.8 Effects of Different Testing Machine .................................... 173
5.3.1.9 Effects of Curing Temperature and Humidity....................... 174
5.3.2 Triaxial Shear Strength...................................................................... 182
5.3.3 Water Permeability............................................................................ 184
5.3.4 Drying Shrinkage ............................................................................... 186
5.3.5 California Bearing Ratio .................................................................... 187
5.3.6 Resilient Modulus.............................................................................. 188
5.3.7 Excavation Study............................................................................... 189

5.4 DURABILITY OF CLSM............................................................................. 201


5.4.1 Freezing-and- Thawing Study............................................................ 201
5.4.2 Wetting-and-Drying Study................................................................ 204

5.5 CORROSION OF DUCTILE IRON IN CLSM ............................................ 206


5.5.1 Introduction ....................................................................................... 206
5.5.2 pH values of Extruded Solutions form CLSM Mixtures................... 207
5.5.3 Resistivity of CLSM Mixture ............................................................ 207
5.5.4 Chloride Concentration and Profiles ................................................. 209
5.5.5 Mass Loss of Ductile-Iron Specimens ............................................... 211
5.5.5.1 Mass Loss of Uncoupled Coupons ........................................ 211
5.5.5.2 Mass Loss of Coupled Coupons ............................................ 219

xiv
5.6 LEACHING AND ENVIRONMENTAL IMPACT..................................... 224

CHAPTER 6: CONCLUSIONS AND FUTURE WORK 228

6.1 CONCLUSIONS ........................................................................................... 228

6.2 FUTURE WORK .......................................................................................... 233

APPENDIX A: REFERCED STANDARD TEST METHODS 236

APPENDIX B: SETTING/HARDENING TEST RESULTS 240

REFERENCES 244

VITA 255

xv
List of Tables

Table 2.1: Four basic types of CLSM .................................................................. 18


Table 2.2: Typical mixture proportions for low fly ash content CLSM............... 19
Table 2.3: Typical Mixture Proportions for High Fly Ash Content CLSM ......... 20
Table 2.4: Currently used or potential test methods for CLSM ........................... 22
Table 2.4: Currently used or potential test methods for CLSM (cont’d) .............. 23
Table 3.1: Aggregate gradations used in laboratory test program........................ 65
Table 3.2: Materials included in laboratory test program ................................... 67
Table 3.3: Chemical characteristics of foundry sand ........................................... 68
Table 3.4: Chemical composition of fly ashes ..................................................... 70
Table 3.5: Che mical analysis of ASTM Type I portland cement (weight percent) .
............................................................................................................................... 71
Table 3.6: Non-air entrained CLSM mixture proportions .................................... 75
Table 3.7: Air-entrained CLSM mixture proportions ........................................... 76
Table 3.8: Selected CLSM mixtures included in Stage One ................................ 77
Table 3.9: CLSM mixture proportions for load-rate study.................................... 79
Table 3.10: CLSM mixtures for cylinder curing/specimen conditioning study... 79
Table 3.11: CLSM mixtures for excavation study ............................................... 79
Table 3.12: CLSM mixtures used for freezing and thawing study....................... 80
Table 3.13: Mixture proportions for alternative capping materials study............ 80
Table 3.14: CLSM mixture proportions for drainage-condition study................. 80
Table 3.15: CLSM mixture proportions for cylinder storage-condition study..... 81
Table 3.16: Mixture proportions for effects of curing temperatures and humidity81
Table 3.17: CLSM mixture proportions for triaxial shear strength and water
permeability............................................................................................... 81
Table 4.1: Summary of initial test program for this study ................................... 83
Table 4.2: Overview of all properties and tests studied ....................................... 84
Table 4.3: Four curing conditions used in the study (using G-series mixtures)... 98
Table 5.1: CLSM mixture proportions and fresh properties .............................. 121
Table 5.2: Analysis of variance (ANOVA) for water demand ........................... 124
Table 5.3: Segregation data for selected CLSM mixtures.................................. 135
Table 5.4: Subsidence results of six selected CLSM mixtures .......................... 136
Table 5.5: Unconfined compressive strength of original 38 CLSM mixtures .... 139
Table 5.6: Comparison of capping materials and methods ................................ 140
Table 5.7: Comparison between predicted compressive strengths (using equations
shown in this section) and actual strengths from previous study (using
different constituent materials) ................................................................ 145
Table 5.8: Effects of curing methods on compressive strength (Mixture B-1) ... 147
Table 5.9: Compressive strength at 28 days using different curing methods...... 148

xvi
Table 5.10: Compressive strength at 91 days using different curing methods.... 149
Table 5.11: Compressive strength at 7 days under different drainage conditions .....
............................................................................................................................. 156
Table 5.12: Compressive strength at 28 days under different drainage conditions ..
............................................................................................................................. 157
Table 5.13: Influence of specimen air-drying on compressive strength (Mixture B-
2).............................................................................................................. 158
Table 5.14: Compressive strength results using different capping materials ..... 160
Table 5.15: Coefficients of variation for compressive strengths using different
capping materials ..................................................................................... 161
Table 5.16: Comparison of “a” values calculated by two methods.................... 165
Table 5.17: Effects of loading rates (wide range) on compressive strength ...... 169
Table 5.18: Effects of load rates (narrow range) on compressive strength........ 170
Table 5.19: Effects of loading rates (wide range) on deformation at peak load 170
Table 5.20: Effect of narrow range load rates (narrow range) on deformation at
peak load.................................................................................................. 171
Table 5.21: Effects of cylinder size on unconfined compressive strength......... 173
Table 5.22: Effects of different testing machines on compressive strength
(coefficient of variation is included in parentheses)................................ 174
Table 5.23: Compressive strength of specimens cured at different temperatures at
different ages ........................................................................................... 177
Table 5.24: Results of triaxial compression tests ............................................... 182
Table 5.25: Water permeability (hydraulic conductivity) of selected CLSM
mixtures ................................................................................................... 184
Table 5.26: Test results of water permeability after freezing-and-thawing cycles ...
............................................................................................................................. 185
Table 5.27: Drying shrinkage of selected CLSM mixtures ................................ 187
Table 5. 28: CBR values for six selected CLSM mixtures................................. 187
Table 5.29: Resilient modulus of six selected CLSM mixtures ......................... 188
Table 5.30: Results of excavation study............................................................. 190
Table 5.31: Comparison between field penetrometer values (field-cured samples)
and compressive strength values (field-cured cylinders) ........................ 191
Table 5.32: Results of excavatability study........................................................ 194
Table 5.33: Compressive and splitting tensile strengths at 7 and 28 days ......... 198
Table 5.34: Effects of temperature and drying conditions on splitting tensile
strength of CLSM.................................................................................... 200
Table 5.35: Freezing-and-thawing results for selected CLSM mixtures (after 7
days of curing) ......................................................................................... 202
Table 5.36: Freezing-and-thawing results for selected CLSM mixtures (after 28
days of curing) ......................................................................................... 202
Table 5.37: Freezing-and-thawing test results of D-series mixtures .................. 203

xvii
Table 5.38: Wetting-and-drying results for selected CLSM mixtures (after 7 days
of curing) ................................................................................................. 205
Table 5.39: Wetting and drying results for selected CLSM mixtures (after 28
days of curing) ......................................................................................... 206
Table 5.40: pH and resistivity values of CLSM mixtures.................................. 208
Table 5.41: Percent mass loss for samples embedded in CLSM only................ 213
Table 5.42: Statistical comparison of repeated CLSM mixtures (uncoupled
corrosion)................................................................................................. 216
Table 5.43: Statistical results of mass-loss test for uncoupled samples (CLSM
mixtures containing fly ash) .................................................................... 218
Table 5.44: Statistical results of mass-loss test for uncoupled samples (CLSM
mixtures containing no fly ash) ............................................................... 219
Table 5.45: Percent mass loss for coupled samples embedded in sand .............. 221
Table 5.46: Statistical comparison of repeated CLSM mixtures on corrosion of
coupled coupons embedded in sand ........................................................ 222
Table 5.47: Statistical results of mass-loss test for coupled samples (CLSM
mixtures containing fly ash) ...................................................................... 223
Table 5.48: Statistical results of mass-loss test for coupled samples (CLSM
mixtures without fly ash) ......................................................................... 224
Table 5.49: Analysis of heavy metal concentration of extract from raw materials
(mg/L) ...................................................................................................... 226
Table 5. 50: TCLP test results for Class C fly ash, Class F fly ash, and bottom ash
(mg/L) ...................................................................................................... 227
Table B-1: Needle penetration (NP) and soil penetrometer (SP) results............ 240
Table B-1: Needle penetrAtion (NP) and soil penetrometer (SP) results (cont’d)....
............................................................................................................................. 241
Table B-1: Needle penetration (NP) and soil penetrometer (SP) results (cont’d)242
Table B-2: Needle penetration (NP), soil penetrometer (SP) and vane shear tester
(VS) Results............................................................................................. 243

xviii
List of Figures

Figure 2.1: Applications of CLSM. Case A: backfill; Case B: utility bedding;


Case C: structural fill. ................................................................................ 44
Figure 2.2: Class D: impermissible beddings. load factor = 1.1. (From Soil
Engineering, 4th Edition, M. G. Spangler and R. L. Handy, Harper & Row,
Publisher, New York, 1982) ...................................................................... 52
Figure 2.3: Class C ordinary beddings. load factor = 1.5 (From Soil Engineering,
4th Edition, M. G. Spangler and R. L. Handy, Harper & Row, Publisher,
New York, 1982) ....................................................................................... 53
Figure 2.4: Class B first class beddings. load factor for ditch conduits =1.9
(From Soil Engineering, 4th Edition, M. G. Spangler and R. L. Handy,
Harper & Row, Publisher, New York, 1982) ............................................ 54
Figure 2.5: Class A: concrete-cradle bedding. load factor 2.25 to 3.4. (From Soil
Engineering, 4th Edition, M. G. Spangler and R. L. Handy, Harper & Row,
Publisher, New York, 1982) ...................................................................... 55
Figure 3.1: Gradations of three fine aggregates and ASTM C 33 limits.............. 66
Figure 4.1: A typical soil pocket penetrometer ..................................................... 88
Figure 4.2: Pocket vane shear tester used in this program .................................. 89
Figure 4.3: Plastic mold were pre-cut and taped before casting specimens ......... 93
Figure 4.4: A typical set-up of compression testing using geotechnical unconfined
compression machine. ............................................................................... 94
Figure 4.5: Drainage condition simulation......................................................... 101
Figure 4.6: Method of capping cylinders with gypsum compound .................... 103
Figure 4.7: Specimen set-up for testing of permeability of CLSM mixture after
freezing-and-thawing cycles.................................................................... 105
Figure 4.8: Corrosion test set-up for comparing corrosion performance of coupons
in CLSM and sand (uncoupled)............................................................... 112
Figure 4.9: Corrosion test set-up for comparing corrosion performance of
galvanically coupled coupons in CLSM and sand. ................................. 113
Figure 4.10: Set-up for resistance measurement ................................................. 114
Figure 4.11: Set-up for chloride ion diffusion.................................................... 115
Figure 4.12: Flow chart to study toxicity (heavy metals) of CLSM constituent
materials .................................................................................................. 119
Figure 5.1: Effects of variables on water demand of non-air-entrained CLSM
mixtures ................................................................................................... 123
Figure 5.2: Correlation between ASTM C 403 and soil penetrometer values ... 127
Figure 5.3: Stand for the pocket soil penetrometer ............................................. 128

xix
Figure 5.4: Comparison between needle penetration (NP) and soil penetrometer
reading (SP) values and the back calculated resistance from soil
penetrometer (BSP). ................................................................................ 130
Figure 5.5: Setting time needle and the wedge at the tip of the needle (following
ASTM C 403).......................................................................................... 132
Figure 5.6: Comparison between soil penetrometer and vane shear values ........ 133
Figure 5.7: Effects of variables on bleeding of non-air-entrained CLSM mixtures.
“+” means increasing the bleeding, “-“ means decreasing, and “X” means
the effect is not significant compared with variations. ........................... 134
Figure 5.8: Correlation between bleeding (weight percent) and subsidence....... 137
Figure 5.9: Load-deformation response of Mixture #12 at 3, 7, and 28 days .... 140
Figure 5. 10: Comparison of measured and predicted strengths of air-entrained
CLSM mixtures at the ages of 7, 28, and 91 days. .................................. 142
Figure 5.11: Comparison of measured and predicted strengths of non-air-entrained
CLSM mixtures ....................................................................................... 144
Figure 5.12: Effects of curing conditions on compressive strength for Mixtures G-
7 and G-8 ................................................................................................. 150
Figure 5.13: Effects of curing conditions on compressive strength for Mixtures G-
9 and G-10 ............................................................................................... 151
Figure 5.14: Curing condition effects on compressive strength of Mixture G-3
and G-6 .................................................................................................... 152
Figure 5.15: Curing condition effects on compressive strength of Mixture G-4
and G-5 .................................................................................................... 153
Figure 5.16: Effects of capping methods on coefficient of variation (7-day
compressive strength) .............................................................................. 162
Figure 5.17: Effects of capping methods on 163
Figure 5.18: Effects of capping methods on coefficient of variation (91-day
compressive strength) .............................................................................. 163
Figure 5.19: Load-deflection curves for mixture E-1 (compression testing at 28
days) ........................................................................................................ 167
Figure 5.20: Compressive strength of mixture H-1 for different curing conditions .
............................................................................................................................. 178
Figure 5.21: Compressive strength of mixture H-2 for different curing conditions .
............................................................................................................................. 178
Figure 5.22: Compressive strength of mixture H-3 for different curing conditions .
............................................................................................................................. 179
Figure 5.23: Compressive strength of mixture H-4 for different curing conditions .
............................................................................................................................. 179
Figure 5.24: Compressive strength of mixture H-5 for different curing conditions .
............................................................................................................................. 180
Figure 5.25: Compressive strength of mixture H-6 for different curing conditions .
............................................................................................................................. 180

xx
Figure 5.26: Stress-strain curves of mixture I-1 and I-4 at 28 days (triaxial
compression testing with an effective confining pressure of 69 kPa)..... 183
Figure 5.27: Penetration resistance (ASTM C 403) vs. compressive strength at
different ages (all C-series mixtures) ...................................................... 192
Figure 5.28: Cracking of CLSM mixture C-2 during DCP penetration............. 195
Figure 5.29: Correlation between stiffness (from GeoGauge) and DCP index.. 196
Figure 5.30: Correlation between DCP Index and Removibility Modulus (RE)
calculated using 28-day compressive strength (lab-cured cylinders) ...... 197
Figure 5.31: Correlation between DCP Index and Removibility Modulus (RE)
calculated using 240-day compressive strength (field-cured cylinders) . 197
Figure 5.32: A cylinder from mixture E-1, before and after being tested for
splitting tensile strength........................................................................... 199
Figure 5.33: Specimens after freezing-and-thawing cycles (CLSM mixtures
containing foundry sand) ......................................................................... 204
Figure 5.34: Chloride profiles for select CLSM mixtures after 22 weeks of
exposure................................................................................................... 210

xxi
Chapter 1: Introduction

1.1 GENERAL

Controlled low-strength material (CLSM) is a relatively new technology

whose use has grown in recent years. CLSM, often referred to as flowable fill, is

a highly flowable material typically comprised of water, cement, fine aggregates,

and often times, fly ash. Other by-product materials, such as foundry sand and

bottom ash, and chemical admixtures, including air-entraining agents, foaming

agents, and accelerators have also been used successfully in CLSM.

CLSM is typically specified and used in lieu of compacted fill in various

applications, especially for backfill, utility bedding, void fill, and bridge

approaches. Backfill includes applications such as backfilling walls (e.g.,

retaining walls) or trenches. Utility bedding applications involve the use of

CLSM as a bedding material for pipe, electrical, and other types of utilities and

conduits. Void-filling applications include the filling of sewers, tunnel shafts,

basements, or other underground structures. CLSM is also used in bridge

approaches, either as a subbase for the bridge approach slab or as backfill against

wingwalls or other elements.

There are various inherent advantages of using CLSM instead of

compacted fill in these applications. These benefits include reduced labor and

equipment costs (due to self-leveling properties and no need for compaction),

faster construction, and the ability to place material in confined spaces. The

relatively low strength of CLSM is advantageous because it allows for future

1
excavation, if required. Another advantage of CLSM is that it often contains by-

product materials, such as fly ash and foundry sand, thereby reducing the demand

on landfills, where these materials may otherwise be deposited.

Despite these benefits and advantages over compacted fill, the use of

CLSM is not currently as widespread as its potential might predict. One reason

for this is that CLSM is somewhat a hybrid material; that is, it is a cementitious

material that behaves more like a compacted fill. As such, much of the

information and discussions on its uses and benefits have fallen outside the

traditional specialties of concrete materials and geotechnical engineering.

Although there is considerable literature available on the topic, it is often not

given the level of attention it deserves by either group.

Many states have developed specifications (in some cases, provisional)

that govern the use of CLSM. However, these specifications differ from state to

state, and moreover, a variety of different test methods are currently being used to

define the same intended properties. This lack of conformity, both on

specifications and testing methods, has also hindered the proliferation of CLSM

applications.

There are also technical challenges that have served as obstacles to

widespread CLSM use. For instance, it is often observed in the field that

excessive long-term strength gain makes it difficult to excavate CLSM at later

ages. This can be a significant problem that translates into added cost and labor.

Other technical issues deserving attention are the compatibility of CLSM with

different types of utilities and pipes, the potential leaching of constituent materials

2
and elements, and the durability of CLSM subjected to freezing-and-thawing

cycles.

1.2 RESEARCH OBJECTIVES

This dissertation is directed towards identifying deficiencies in the state-

of-the-practice, modifying or developing new test methods, and understanding

factors affecting behaviors of CLSM mixtures. Specifically, the objectives of the

study are to:

1. Summarize the state-of-the-art and state-of-the-practice of CLSM

technology,

2. Identify and understand factors affecting the water demand of CLSM

mixtures to reach a target flow,

3. Evaluate bleeding and segregation of fresh CLSM mixtures, identify

factors influencing the amounts of bleed water, and develop test

method for measuring segregation,

4. Understand the setting and hardening behavior of fresh CLSM

mixtures and recommend test methods for laboratory and field testing,

5. Develop a suitable compression test method to measure strength of

CLSM mixtures, including providing guidance on loading rate,

cylinder curing conditions, and appropriate capping materials,

6. Develop predictive models for the short- and long-term compressive

strength of CLSM,

3
7. Study the effects of curing temperature, humidity, and other

parameters on the strength gain of CLSM mixtures,

8. Study the suitability of certain standard test methods on CLSM,

including triaxial shear test, water permeability, California Bearing

Ratio (CBR), and resilient modulus,

9. Evaluate drying shrinkage of CLSM mixtures,

10. Evaluate methods to predict future excavation of CLSM mixtures,

11. Study durability of CLSM mixtures, including freezing-and-thawing

resistance and wetting-and-drying resistance,

12. Evaluate corrosion behaviors of ductile-iron specimens in CLSM,

including coupled and uncoupled cases, and identify potential

constituents affecting corrosion rate,

13. Propose a procedure to evaluate the leaching and environmental

impact of CLSM mixtures.

1.3 OUTLINE OF THE DISSERTATION

The dissertation is comprised of six chapters. A brief summary of each

chapter is provided below.

Chapter 1 explains the overall objectives of this study.

Chapter 2 summarizes the state-of-the-art and current practice related to

CLSM. Typical applications of CLSM are also provided with case histories.

Chapter 3 describes the CLSM mixture proportions used in the study,

which included 38 mixtures in the Stage One study and 67 mixtures in Stage Two.

4
Three types of aggregates (concrete sand, foundry sand, and bottom ash), three

types of fly ash (Class C, Class F, and High carbon), and American Society for

Testing and Materials (ASTM) Type I cement were included in the study.

Relevant physical and chemical properties of these materials were also measured.

Chapter 4 provides the detailed test program in this research. The

program included the measurements of bleeding and segregation, flow, unit

weight, air content, compressive strength, CBR, resilient modulus, traixial shear

strength, water permeability, drying shrinkage, freezing-and-thawing resistance,

wetting-and-drying resistance, corrosion of ductile-iron samples, and

environmental impact. Factors affecting the testing and development of

compressive strength are included in the test program.

Chapter 5 provides the results and discussions. Models predicting strength

development are provided. Methods predicting setting and hardening are

compared. Effects of curing temperature and humidity are discussed. Factors

affecting the testing of compressive strength of CLSM are studied and

recommendations to modify current ASTM standard test methods are provided.

Methods predicting excavatability are compared. Factors potentially affecting

corrosion behaviors are identified. A systematic procedure was developed to

assess the potential toxicity of raw materials used in CLSM.

Chapter 6 summarizes the main conclusions of the study and identifies

technical issues that deserve attention in future research.

5
Chapter 2: Literature Review

2.1 INTRODUCTION

Controlled low-strength material (CLSM) is, as defined by American

Concrete Institute (ACI) Committee 229, a self-compacted, cementitious material

used primarily as a backfill in lieu of compacted fill.1 Several terms are currently

used to describe this material, including flowable fill, unshrinkable fill, controlled

density fill, flowable mortar, plastic soil-cement, soil-cement slurry, K-Krete and

other various names.

Controlled low-strength materials are defined by “Cement and Concrete

Terminology (ACI 116R)” as materials that result in a compressive strength of 8.3

MPa or less.1 However, most current CLSM applications require unconfined

compressive strength of 2.1 MPa or less. Some researchers consider the range of

0.3 to 1.1 MPa as a good index of sufficient strength and easy future excavation.

For applications where future excavations are expected, the excavatability of

CLSM is critical and this may determine the success of CLSM in practices such

as utility bedding.

Soil-cement has been a widely used material in geotechnical-engineering

practices for a long time. CLSM is relatively new and is different from

conventional soil-cement. Compaction is usually required for soil-cement, which

is not the case for CLSM. One of its earliest applications was carried out in 1964

by the US Bureau of Reclamation as the bedding of the 515-km long pipelines in

the Canadian River Aqueduct Project, which runs from north of Amarillo to south

6
of Lubbock, Texas.2 The material used in that project was called plastic soil-

cement. The soil used consisted of local blow sand deposits. A new construction

procedure was introduced, and the cost of this project was estimated to be 40

percent less than using conventional backfilling techniques. The productivity was

increased from about 120 to 305 m of pipe placed per shift.

In the early 1970s, Detroit Edison Company, in cooperation with Kuhlman

Corp., a ready-mix concrete producer in Toledo, Ohio, investigated an alternative

to compacted granular fill utilizing fly ash and concrete batching techniques. This

new backfill material, called flowable fly ash, was used in several applications in

the late 1970s.3-5 This material was composed principally of fly ash and typically

4 to 5 percent cement, along with an appropriate amount of water. In the Belle

River project, it was estimated that more than $1 million was saved by using this

new material.3 One impressive and exciting feature of this material was that it

remained cohesive when being placed. Another characteristic was the steep angle

of repose when it was placed either above or under water.

Eventually, a company known as K-Krete Inc. was formed. In 1977, four

patents for K-Krete were issued to “Brewer et al.”.6 A typical K-Krete mixture

consisted of 1305 to 1661 kg of sand, 166 to 297 kg of fly ash, 24 to 119 kg of

cement, and up to 0.35 to 0.40 m3 of water per m3 of product. The four patents

included mixture design, backfill technique, pipe bedding, and dike construction.

These patents were sold to Contech, Inc., in Minneapolis, Minn., who ceded the

patent rights later to the National Ready Mix Concrete Association (NRMCA)

with the stipulation that those rights may not be used in a proprietary manner.6,7

7
Since then, ready-mixed concrete producers and contractors have used similar

materials to K-Krete without violating patent restrictions.

Following K-Krete’s emergence as a replacement material for

conventional compacted fill, similar materials have been developed and used

throughout the United States and Canada. However, the lack of a centralized

source for obtaining and disseminating information within the marketplace

appeared to cause confusion and reluctance on the part of the engineering

community to use these materials. In response to the proposal of Brewer, ACI

Committee 229 was established in 1984 under the title “Controlled Low-Strength

Materials (CLSM)”. After years of efforts, in 1994, the committee published a

report called “Controlled Low Strength Materials (CLSM),”1 which has been

referenced widely. In 1999, the revised edition was published with the same title.

In 1998, the American Society for Testing and Materials (ASTM)

published a book titled “The Design and Application of Controlled Low-Strength

Materials (flowable fill).”8 The articles in this book represented the state of art

and practice of CLSM in the field and in the research laboratory at that time.

Different types of waste materials were included in CLSM mixtures to recycle

waste and reduce the cost. Currently, there are five ASTM testing standards

available for CLSM. Specific information and description on these and other test

methods relevant to CLSM are provided later in this dissertation. For

convenience, these tests are also listed in Appendix A.

Many universities and state Departments of Transportation (DOTs) have

conducted comprehensive studies on CLSM.9-19 Some of these are still in

8
progress. For instance, The University of Louisville and Purdue University have

conducted extensive research on CLSM. State DOTs have used CLSM in

constructing of small-span bridges and culverts.20-22 The Iowa DOT used it as

backfill for bridge approaches and bridges replacement.23 Certain companies,

such as W. R. Grace Construction Products and Master Builders Technologies,

have developed powerful air-entraining agents to produce CLSM with high air

content. The air content of the air-modified CLSM can be as high as 15 to 35%,

and the corresponding density is significantly decreased.24

A comprehensive investigation of CLSM, funded by the National

Cooperative Highway Research Program (NCHRP) under Project 24-12 and 24-

12(1) (“Controlled Low-Strength Material for Backfill, Utility Bedding, Void Fill,

and Bridge Approaches”) has culminated in two reports. One report summarized

the state-of-the-art and current practice related to CLSM in 1999 and the other

report summarized the findings of a comprehensive laboratory investigation of

CLSM in 2001.25,26 Much of the laboratory study described in Chapter 4 of this

dissertation was part of these NCHRP-funded efforts.

2.2 MATERIALS

To compete with conventional compacted fill, the cost of CLSM plays a

key role. The materials used in CLSM are thus of crucial importance for this

technology. One significant benefit of CLSM is its ability to use a wide range of

local materials, including by-product materials. Portland cement, fly ash,

9
aggregates (e.g., foundry sand), chemical admixtures, and other by-product

materials can be included in the production of CLSM.

2.2.1 Bonding Materials

2.2.1.1 Portland Cement

Although there is no restriction placed on the types of portland cement

used in CLSM, ASTM Type I is most commonly used. The prevailing criteria are

the local availability and cost of cement, and Type II or Type I/II cements may be

more common in certain regions of the United States. Because of the

comparatively low cement contents in CLSM, common concrete durability

problems, such as alkali-aggregate reaction and sulfate attack, appear quite

unlikely. Type III portland cement has been successfully used in CLSM to

achieve higher early strength and to reduce subsidence.25

2.2.1.2 Fly Ash

Class C fly ash can be used as a binder because of its cementitious

property. The ash typically contains 15 to 35 percent analytical CaO. With

similar proportions, CLSM mixtures with Class C fly ash demonstrate higher

compressive strength than those with Class F fly ash. One example is Flash Fill,

which contains predominantly Class C fly ash, concrete sand and water.27 No

portland cement is added for binding. Its quick-setting, self-leveling, self-

compacting characteristics make trench road repairs faster, easier, and more

economical. These characteristics make it a promising material. Other uses of


10
flash fill include building foundations, fill around pipes, gas lines, and manholes,

and replacement of weak subgrade beneath footings. Flash fill can be hand-

excavated without power-assisted tools or machinery. However, the quick-setting

property of flash fill may cause potential construction difficulties.

2.2.2 Aggregate

Unlike conventional soil-cement, CLSM does not use local soil as an

ingredient. Because of the low cement content of CLSM, the cost of aggregate

materials actually dominates the total cost. It is common that CLSM producers

use whatever suitable aggregate materials that are locally available in the

mixtures, such as concrete sand, foundry sand, crushed glass, by-product fines,

and other aggregates.

2.2.2.1 Concrete Aggregate

Although a wide range of aggregate materials may be used successfully in

CLSM, conventional concrete sand (satisfying ASTM C 33) is most commonly

utilized, especially for CLSM produced at ready-mix concrete plants (the

dominant source of CLSM). Sand that barely meets the ASTM C 33 requirements

can be used in CLSM production as long as specified flowability is satisfied.

CLSM have mostly evolved using only sand as aggregate. CLSM in the

Pacific Northwest has developed differently in that many mixtures use gravels up

to 25-mm top size.28 The reasons for the use of gravel center around economy

and performance. Concrete technology demonstrates that if the largest top-size

11
aggregate is used, the lowest void content in the combined aggregates will be

achieved. Reduced voids result in a lower paste requirement, which

correspondingly reduces the cost on cementitious materials. Gravel can be a

viable material as aggregate in CLSM proportions. Economics are likely to

determine whether gravel is used or not. Performance of CLSM mixtures with

gravel may be expected to be similar to those with sand only. Subsidence can be

reduced as a result of low mixing water requirement.

2.2.2.2 Foundry Sand

Foundry sand, a by-product of the metal casting industry, has been studied

and used successfully in CLSM, and its use has increased in recent years.10,30-32

Foundry sand is a viable candidate for use in CLSM because of its lower cost,

increasing availability, and satisfactory performance.

It is estimated that a typical foundry generates approximately one ton of

waste sand for every ton of metal castings produced and shipped.33 The most

commonly used waste foundry sand in CLSM is “green sand,” a term applied

when the original uniform sand is treated with a bonding agent (usually clay) to

optimize the efficiency of the sand in the molding process. After molding is

completed, the sand is discarded, often in landfills, and in 1994 typical costs were

$22/tonne to $44/tonne.8

The Federal Highway Administration (FHWA) issued a report, “User

Guidelines for Waste and By-Product Materials in Pavement Construction,”

which covers in detail the use of foundry sand (and fly ash) in CLSM and

12
provides guidelines for its proper usage.34 The Environmental Protection Agency

(EPA) has also recognized foundry sand, along with fly ash, as suitable materials

for CLSM.35

There has been a perception that using foundry sand in CLSM may have a

detrimental environmental impact because heavy metals may leach from the

foundry sand. Ferrous foundry sands are more commonly used in CLSM because

there are concerns about the contents of heavy metal in non-ferrous foundry sand.

The EPA does not recommend using non-ferrous foundry sand in CLSM because

of the concerns over the potential leaching of phenols and heavy metals, such as

cadmium, lead, copper, nickel, and zinc.35 Bhat et al. reported the results of their

environmental tests.10 Bio-assay tests were performed as part of this study to

screen out potentially hazardous waste foundry sands. The test was performed

both on expressed CLSM pore solution and raw foundry sand. No firm

conclusions were presented in that report.

2.2.2.3 Excess Fines from Aggregate Crushing Operation

With the development of performance-enhancing admixtures (air-

entraining foam), high-fines limestone screenings have been used as aggregate for

CLSM.36,37 It was estimated that more than 81.6 million tonnes of quarry by-

products were generated in 1993.36 It is difficult to find viable uses for by-

product fines, which account for 15- to 20-percent of aggregate produced, because

most construction and highway material specifications limit the aggregate finer

than 0.075 mm to 5 or 7 percent or less specified by ASTM C 33. A study

13
showed that 14 to 30 percent of air worked well for CLSM mixtures containing

high-fines aggregate.37 Screenings that contain up to 21 percent finer than 0.075

mm particles can be used as aggregate to produce CLSM mixtures meeting the

National Ready Mixed Concrete Association (NRMCA) performance

recommendations.

Another source of high-fines aggregate is recycled concrete. Current

practice usually involves using recycled concrete as coarse aggregate, leaving an

abundance of fines (passing 300-ìm sieve), which may be well suited for use in

CLSM.

2.2.2.4 Bottom Ash

Bottom ash and fly ash are both by-product materials of coal combustion.

Bottom ash is formed by large noncombustible particles that cannot be carried by

the hot gases. These particles descend on hoppers or conveyors, at the bottom of

the furnace, in a solid or partially molten condition. The particles gradually cool

down to form bottom ash. Bottom ash particles are typically porous and angular

in shape. As a by-product material, bottom ash is commonly disposed of in

ponds. In this process, bottom ash is passed through a crusher to reduce the size

of large particles, and it is transported hydraulically through pipelines to the pond

site. The typical range of particle sizes falls between 75 microns and 25

millimeters. Bottom ash can be used in compacted fill when combined with fly

ash.38 Researchers have successfully used bottom ash in CLSM.39

14
2.2.2.5 Other Aggregate Materials

In addition to the aggregate materials previously described, there are other

materials used in CLSM as aggregates. Colored glass that cannot be recycled by

local bottle manufacturers has been crushed to pass a 12.5-mm (½ in) sieve and

successfully used in CLSM as an aggregate.40 A special process was utilized so

that the crushed glass could be handled with bare hands. Phosphogypsum is a by-

product of the production of phosphoric acid and has been shown to be a viable

aggregate for CLSM.41 Toxicity characteristic leaching procedure (TCLP) testing

showed that the toxic contents of the individual materials and final design pass

mixtures were well below the EPA leachate standards. Ground granulated blast

furnace slag (GGBFS) and bentonite have also been used in CLSM.59

2.2.3 Admixtures

2.2.3.1 Mineral Admixtures

Fly ash is a by-product of coal combustion and has been used in a wide

range of construction applications, including CLSM. Fly ashes are broadly

classified into two classes based on the calcium content and pozzolanic properties

(ASTM C 618). Fly ash normally produced from lignite or subituminous coals is

classified as Class C fly ash, which has cementitious properties in addition to

pozzolanic properties. Fly ash normally produced from anthracite or bituminous

coal is classified as Class F fly ash. It has only pozzolanic properties and has no

self-cementitious characteristics because of the absence of sufficient amount of

calcium oxide.

15
Fly ash is used mostly in portland cement concrete, but its use in CLSM

has increased considerably in recent years. Even so, approximately 70 to 75

percent of fly ash generated annually is still disposed in landfills.42 A majority of

this unused fly ash does not meet the specifications for use in portland cement

concrete (the dominant application), sometimes because of the high percentages

of unburned carbon, as measured by the loss on ignition (LOI) test. Values of

LOI of fly ash used in concrete are typically limited to six percent for Class C and

Class F fly ash according to ASTM C 618, whereas certain bituminous fly ashes

may have LOI values in excess of 15 to 20 percent. Higher unburned-carbon

contents increase water demand in concrete and can significantly raise chemical

admixture demand (especially air-entraining agents and superplasticizers).

Despite the limitations placed on fly ash for conventional concrete, it has been

demonstrated that CLSM can be successfully produced using a wide variety of fly

ash types and sources, including high-carbon fly ash that is not permitted in

concrete.

Because ASTM C 618 Class F and Class C fly ash are commonly

available commercially, they are common mineral admixtures used in CLSM.

There are numerous benefits of using fly ash in CLSM, including improved

flowability, reduced segregation and bleeding, and in many cases decreased

material cost.

16
2.2.3.2 Chemical Admixtures

Air-entraining agents are the most commonly used chemical admixtures in

CLSM, and set accelerators have been used to a lesser extent to increase the speed

of construction (e.g., earlier opening of traffic) and to minimize subsidence.

CLSM is often produced with relatively high air contents (e.g. 20-35 percent),

either through the use of potent, liquid air-entraining agents or with foaming

agents applied through a foaming gun.24,43,44 The advantages of such CLSM

include low density, improved insulation properties, reduced segregation and

bleeding, decreased water and/or cement content, improved frost resistance, and

lower material cost. Also, high air contents may be used to limit long-term

strength gain to assure future excavatability.

2.2.4 Water

There are no special requirements for water to be used in CLSM. As a

general rule, any water that is suitable for concrete will work well for CLSM.

2.3 MIXTURE PROPORTIONS

Currently, there is no standard mixture proportioning method that has been

widely adopted for CLSM. A group of studies have focused on CLSM mixture

proportions,29-30,45-46 but most CLSM producers specify mixtures based on past

experience with locally available materials. Because a variety of by-product and

non-standard materials are typically used in CLSM, it is difficult to find a

standard approach to mixture proportioning.

17
In general, there are four typical types of CLSM used in current practice,

as shown in Table 2.1.

Table 2.1: Four basic types of CLSM

Type Binder Flow-driving factors Aggregate

I (K-Krete) Cement, fly ash Water, fly ash Filler*

II (high fly ash) Cement, fly ash Water, fly ash Fly ash

III (high air) Cement High air, water Filler

IV (flash fill) Fly ash Water, fly ash Filler


* Note: Filler includes concrete sand, bottom ash, foundry sand, and high -fine aggregate, depending on the mixture
proportions.

Type I CLSM was the earliest type that appeared in the construction field.

The mixture proportions tend to vary with different applications, especially the

cement content, which is the main factor affecting the strength of CLSM. Type II

CLSM mixtures usually have cement contents of 4 to 5 percent of the total solid

materials. Because of its excellent flowability, Type II CLSM is used widely in

backfilling spaces where access for compaction is limited, if not impossible.

Type II CLSM could achieve compressive strength values as high as 690 kPa at

28 days. To produce Type III CLSM, potent chemical admixtures are required to

entrain large volume of air voids. Type III CLSM is efficient in reducing fill load

because a density as low as 673 kg/m3 can be reached by using foaming agents.47

Type IV CLSM is patented and little is known about its proportions, but the

18
mixtures are typically high in Class C fly ash content. Type I and III CLSM

mixtures are the most commonly used mixtures by state DOTs.25

Although the unconfined compressive strength of CLSM can be as high as

8 MPa, the vast majority of applications are designed for future excavatability and

have strengths of less than 1 to 1.4 MPa. For these mixtures, low cement contents

are used (e.g., 30 to 60 kg/m3 ), with or without fly ash. Certain mixtures use very

high fly ash contents (Type III), where fly ash effectively serves both as a binder

and as an aggregate. Tables 2.2 and 2.3 show what the FHWA considers to be

typical mixture proportions for low and high fly ash contents. Mixtures that do

not contain fly ash are often produced with high air contents (20 to 35%), where

the high air contents provide certain similar benefits as fly ash, including

improved flow, reduced segregation and bleeding, and decreased cement content.

It also minimizes the subsidence of backfill layers because of little bleeding.

Table 2.2: Typical mixture proportions for low fly ash content CLSM 42
Material Range (kg/m3) Typical mixture proportions (kg/m3)
Fly ash (6-14%)* 119 - 297 178
Cement 30 - 119 59
Sand 1483 - 1780 1542
Water 198 - 494 297
Total: 2076
* Higher calcium fly ash is used in lower amounts than low calcium fly ash

19
Table 2.3: Typical Mixture Proportions for High Fly Ash Content CLSM 42
Material Range (kg/m3) Typical mixture proportions (kg/m3)
Fly ash (95%) 949 - 1542 1234
Cement 47 – 74 62
Water 222 - 371 247
Total: 1543

ACI Committee 229 opposes the use of plastic fines such as clay because

they can produce deleterious results, such as increased shrinkage.1

Brewer categorized CLSM mixtures according to their specific

applications.48 CLSM mixture used in backfill is referred to as CLSM-CDF,

pavement base as CLSM-CPB, structural fill as CLSM-CSF, thermal fill as

CLSM-CTF, anti-corrosion fill as CLSM-ACF , and permeability fill as CLSM-

CPF. The term CDF often appears in literature. It is a term created by Brewer to

represent Controlled Density Fill, which indicates a uniform density from the

trench’s bottom to its top. The density of the CDF material can be adjusted by

changing its components.48

2.4 BATCHING, MIXING, AND TRANSPORTING

CLSM is typically batched, mixed, and transported in similar fashions as

concrete. Most CLSM is batched at ready-mix plants and mixed in truck mixers.

Because of the amount of fines included and the fact that CLSM does not usually

contain coarse aggregate (which would assist in breaking up cement clumps and

dispersing the mixture), it is important to mix CLSM thoroughly. There is no


20
standard mixing procedure for CLSM, and mixing time varies with CLSM types.

Different mixing procedures have been described in the literature.

The high fluidity of CLSM may create difficulties in transporting full or

near-full loads in ready mixed trucks. To address this potential problem, certain

producers hold back part of the mixing water and/or liquid air-entraining agents

for on-the-job-site additions rather than at the plant, thereby reducing the mixture

volume in the truck on route to the site. The high fluidity of CLSM affects

several aspects of placement (e.g., floating of pipes) and construction.

2.5 CLSM PROPERTIES AND TEST METHODS

Table 2.4 summarizes test methods and critical CLSM properties that

affect its performance in backfill, utility bedding, void fill, bridge approach, and

other applications.25 Currently, there are five ASTM test methods specifically

designed for CLSM. The table lists various test methods that are currently being

used to assess CLSM or tests that may have the potential to serve as a suitable

methods. CLSM is a hybrid material (falling between soil and concrete), and as

such, soil and concrete test methods may be adopted for use. Focus should be

placed on the most important CLSM properties, their effects on performance, and

current or potential test methods that are most applicable to CLSM.

21
2.5.1 Fresh CLSM Properties

2.5.1.1 Flowability

One of the most important attributes of CLSM is its capacity to easily flow

into confined areas and obtain required strength and/or density without the need

for conventional placing and compacting equipment.1 The self-leveling properties

of CLSM significantly reduce labor and increase construction speed and safety.

Because the enhanced flow properties of CLSM are critical to successful

placement and performance, flowability is measured regularly as an important

quality-control parameter.

Table 2.4: Currently used or potential test methods for CLSM


CLSM Test Methods
Property
AASHTO T 276, “Developing Early Age Compression Test Values and Projecting Later
Age Strengths"
ASTM D 4832, “Preparation and Testing of Controlled Low-Strength Material (CLSM) Test
Cylinders”
ASTM D 4429, “CBR (California Bearing Ratio) of Soils in Place”
Mechanical AASHTO T 193, “Standard Method of Test for The California Bearing Ratio”
properties AASHTO T 24 “Obtaining and Testing Drilled Cores and Sawed Beams of Concrete”
AASHTO T 222, “Nonrepetitive Static Plate Load Tests of Soils and Flexible Pavement
Components for Use in Evaluation and Design of Airport and Highway Pavements”
ASTM D 6024, “Test Method for Ball Drop on Controlled Low Strength Material to
Determine Suitability for Load Application”
AASHTO T 22, “Compressive Strength of Cylindrical Concrete Specimens”
AASHTO T 106, “Compressive Strength of Hydraulic Cement Mortar”
AASHTO T 292, “Standard Method of Test for Resilient Modulus of Subgrade Soils and
Untreated Base/Subbase Materials”
ASTM D 6103, “Standard Test Method for Flow Consistency of Controlled Low-Strength
Flow Material”
ASTM C 939, “Flow of Grout for Preplaced-Aggregate Concrete (Flow Cone Method)”
CRD-C611, “Method of Test for Flow of Grout Mixtures”
Unit weight ASTM D 6023, “Standard Test Method for Unit Weight, Yield, and Air Content
yield and (Gravimetric) of Controlled Low Strength Material”
air content AASHTO T 121, “Unit Weight, Yield, and Air Content (Gravimetric) of Concrete”

22
Table 2.4: Currently used or potential test methods for CLSM (continued)
CLSM Test methods
property
Hardening AASHTO T 197, “Time of Setting of Concrete Mixtures by Penetration Resistance”
time
Sampling ASTM D 5971, “Standard Practice for Sampling Freshly Mixed Controlled Low-
Strength Material”
AASHTO T 141, “Sampling Fresh Concrete”
Segregation ASTM C 940, “Standard Test Method for Expansion and Bleeding of Freshly Mixed
and bleeding Grouts for Pre- Placed Aggregate Concrete in the Laboratory”
Drying AASHTO T 160, “Shrinkage of Portland Cement Concrete”
shrinkage
Consolidation ASTM D 2435, “Standard Test Method for One-Dimensional Consolidation Properties
of Soil”
ASTM D 5084, “Measurement of Hydraulic Conductivity of Saturated Porous Materials
using a Flexible Wall Permeameter”
Permeability AASHTO T 277, “Rapid Determination of the Chloride Permeability of Concrete”
ASTM D 4525, “Standard Test Method for Permeability of Rocks by Flowing Air”
Freezing-and- AASHTO T 161, “Standard Test Method for Resistance of Concrete to Freezing and
thawing Thawing”
resistance ASTM D 560, “Standard Test Methods for Freezing and Thawing of Compacted Soil-
Cement Mixtures”
ASTM D 5334, “Standard Test Method for Determination of Thermal Conductivity of
Soil and Soft Rock by Thermal Needle Probe Procedure”
Thermal ASTM D 5335, “Standard Test Method for Linear Coefficient of Thermal Expansion of
properties Rock Using Bonded Electrical Resistance Strain Gages”
ASTM D 4612, “Standard Practice for Calculating Thermal Diffusivity of Rocks”
ASTM G 1, “Standard Practice for Preparing, Cleaning, and Evaluating Corrosion Test
Specimens”
Corrosion of
ASTM G 51, “Standard Test Method for Measuring pH of Soil for Use in Corrosion
metals in Testing”
CLSM Modified ASTM G 59, “Standard Practice for Conducting Potentiodynamic Polarization
Resistance Measurements”
Modified ASTM G 109, “Standard Test Method for Determining the Effects of Chemical
Admixtures on the Corrosion of Embedded Steel in Concrete Exposed to Chloride
Environments”
ASTM G 57, “Standard Test Method for Field Measurement of Soil Resistivity Using
the Wenner Four-Electrode Method”
Leaching ASTM D 4874, “Standard Test Method for Leaching Solid Material in a Column
Apparatus”
EPA SW-846, Method 1311, "Toxicity Characteristic Leaching Procedure (TCLP)"
Utility ASTM D 543, “Standard Practices for Evaluating the Resistance of Plastics to Chemical
degradation Reagents”

ASTM D 6103, “Standard Test Method for Flow Consistency of

Controlled Low Strength Material,” has become the most common test to measure

CLSM flow since its adoption by ASTM. The test method uses a 75-mm × 150-

mm cylinder, open at both ends, which is filled with CLSM and lifted between 2
23
to 4 seconds allowing the CLSM to slump and to increase in diameter. The final

diameter of “pancake” is typically used to differentiate between various degrees

of flowability. A final diameter of 200 mm or higher is typical for a highly

flowable mixture. Some researchers have found that this test method is not as

sensitive to consistency as the conventional slump test.49 Water and fine material

may run out of the cylinder, leaving a sand cylinder standing. When the slump

cone was used, the sand mix tended to flow with the water and fines.

ASTM C 143, “Standard Test Method for Slump of Hydraulic Cement

Concrete”, has been used to a lesser extent to measure the flowability of CLSM.

The Washington Aggregate and Concrete Association defines flowability ranges

associated with the slump cone as follows:50

Low flowability: less than 150 mm


Normal flowability: 150 to 200 mm
High flowability: greater than 200 mm

ASTM C 939, “Standard Test Method for Flow of Grout for Preplaced-

Aggregate Concrete,” measures the efflux time of CLSM as it passes through a

flow cone. Several state DOTs have specified this test method for CLSM, and the

Florida and Indiana Departments of Transportation require an efflux time of 30 ±

5 seconds.1

Similar to the workability of concrete, flowability is not a fundamental

property of CLSM. To be meaningful, it must be related to the type of

construction and requirement of flowability. There are many factors that

influence the flowability of CLSM, including CLSM components, aggregate

24
gradation and shape, air content, water content, fly ash type and quantity, and

others. Specific methods of performing ASTM D 6103 affect the value of

flowability, and the lifting force has been found to be particularly important.51

To achieve the desired flow, trial mixtures are recommended. Bhat

proposed the concept of a “flow curves” to study possible factors affecting the

flow of CLSM.10 He found that when proper contents of fly ash were used,

adequate flow was obtained. But when too much fly ash was used, more water

was required. The study was limited mainly to CLSM using foundry sand as

aggregate. Gray used rheology to study the flow behavior and regarded the grout

as a Herschel-Bulkley (HB) or Bingham type fluid.52 Unfortunately, this

approach may be too complicated for field application. Anti-washout admixture

(AWA) and high-range water-reducing admixture (HRWRA) were found to help

underwater placement and reduce cross-contamination.53 Janardhanam et al. have

found that superplasticizers could significantly reduce the water requirement of

flowable fly ash, but the high cost of this admixture usually limits its use in

CLSM. 54

2.5.1.2 Segregation and Bleeding

Because of the high-water content requirement for the high fluidity of

CLSM mixtures, there is the potential for excessive segregation and bleeding,

especially for Type I and II CLSM mixtures. Because all components of CLSM

are in suspension when being placed, it is natural that segregation and bleeding

tend to occur when mixtures consolidate. When water is the main flow-enhancing

25
ingredient and there is insufficient surface area (fines), segregation and bleeding

are likely to occur. Generally, the use of fly ash and air-entraining agents reduce

the water demand, thus minimizing the potential for segregation and excessive

bleeding. Using ASTM Test Method C 940, Hoopes found that the 30% air-

modified mix had 0% bleed water, while the non-air-modified CLSM mixture

yielded 2.4% bleed water at the top of the sample.24 Little has been reported on

segregation and bleeding of Type IV CLSM mixtures.

CLSM mixtures, even with high air content, are generally pumpable with

careful proportioning.

2.5.1.3 Hardening Time

ACI Committee 229 defines hardening time as the approximate period of

time required for CLSM to develop from the plastic state to a hardened state, with

sufficient strength to support the weight of an average person. 1 Hardening time

can be as short as one hour, but it generally takes three to five hours under normal

conditions.55 There are many factors affecting the hardening of CLSM. These

factors include binder type and content, aggregate type, drainage conditions,

proportioning of CLSM, temperature, humidity, and depth of fill.

Hardening time, as defined above, is approximate and likely originated

from the concept of “Standing-On Time” in concrete technology. The most

common and easiest approach is to use penetration resistance as an index of

hardening. ASTM C 403, “Standard Test Method for Time of Setting of Concrete

Mixtures by Penetration Resistance,” has been frequently utilized. The soil

26
pocket penetrometer has also been used to study the hardening of CLSM.

Terzhaghi’s bearing-capacity equation of a circular footing is the theoretical basis

of this method.10 Resistance values from the soil pocket penetrometer generally

correlate well with those of ASTM C 403. Penetration values are often specified

to schedule construction practices and the time to opening of traffic.

2.5.1.4 Subsidence

Subsidence or volume reduction of CLSM results from the loss of water

(through bleeding and absorption into surrounding environment) and/or entrapped

air. CLSM mixtures with significant bleed water may have considerable

subsidence, up to approximately 1 to 2 percent of the fill depth. 56 Subsidence

generally occurs during CLSM placement, up until the fresh material sets.

Accelerating admixtures or high early-strength cement may decrease subsidence

by reducing bleed water. Currently, little information is available on subsidence.

2.5.2 Hardened CLSM Properties

2.5.2.1 Compressive Strength

The strength components of CLSM can be considered in terms of

particulate and non-particulate components.10 The non-particulate component of

strength results from the cementitious (and pozzolanic) reaction of cement and fly

ash with water. The particulate component of strength is composed of pure

friction and dilatancy, similar to granular soil. For unconfined compressive

strength, the non-particulate component is dominant because granular soil has no


27
shear strength without confining pressure. Water cement ratio is considered as

the single most important parameter that determines the unconfined compressive

strength of CLSM mixtures.10 Some researchers have deemed cement content as a

more convenient index of strength development, possibly because of the

insensitivity of strength to water cement ratio variations at high levels (>3).20 The

type and amount of fly ash (if used) also have important effects on the strength

development of CLSM mixtures. Many mixture proportions have been studied in

laboratory testing.57,58 Because of the variety of component materials in CLSM, it

is difficult to develop a general model to predict the strength development of

CLSM mixtures.

Compressive strength is one of the most important CLSM properties and it

is generally measured by most practitioners. CLSM mixtures with compressive

strengths of 0.35 to 0.69 MPa (50 to 100 psi) have similar bearing capacities as

well-compacted granular soil. CLSM strength values are often specified for

future excavation (e.g., maximum allowed values of 0.35 to 1 MPa), pavement

construction, and opening of traffic.

ASTM D 4832, “Preparation and Testing of Controlled Low-Strength

Material (CLSM) Test Cylinders,” was adopted in 1995. It has gained acceptance

with state agencies and commercial testing laboratories. The relatively low

strength of CLSM may cause potential problems, such as damaging the test

cylinder when stripping of CLSM cylinders from plastic molds. Because large-

capacity concrete compression machines usually have poor accuracy in the

required low load ranges, the results often show significant variations.26 The

28
capacities of many load frames used to test CLSM specimens in research

laboratories are typically in the range of 1,300 kN to 2,220 kN. To fail a 150-mm

x 300-mm CLSM specimen with a compressive strength of 1 MPa, a load of only

about 18 kN is required. This load is only about one percent of the load frame

capacity in a typical concrete compression machine. According to one survey,

most state DOTs used concrete machines in the testing of CLSM compression

specimens.25 Researchers using a loading rate of 0.24 MPa/sec (35 psi/sec) failed

CLSM specimens in a few seconds and could not satisfy the minimum

requirement of two-minute failure time specified in the ASTM standard.10

2.5.2.2 Permeability

In certain applications, the permeability of CLSM to both liquids and

gases is essential for construction and field performance. Drainage

characteristics, durability and leaching potential are influenced by permeability.

Because of the relatively low permeability of CLSM, certain specifications limit

the use of CLSM as utility bedding for gas pipelines. Difficulties have been

reported in detecting gas-leaks in pipelines buried in CLSM mixtures.25 Typical

water permeability values of CLSM are in the range of 10-4 to 10-5 cm/sec and

may achieve 10-7 cm/sec in mixtures with high strength or high fine contents.1

CLSM mixtures with 30-percent and 21-percent air contents have yielded

permeability values of 1.7 x 10-2 and 1.2 x 10-3 cm/sec, respectively. 24

ASTM D 5084, “Standard Test Method for Measurement of Hydraulic

Conductivity of Saturated Porous Materials Using a Flexible Wall Permeameter,”

29
is often adopted for CLSM mixtures. In experimental work of Bhat et al, a

backpressure of around 550 kPa was applied and maintained.10 In his study, the

“B” value was checked. The “B” value is the ratio of pore water pressure change

to confining pressure change under undrained conditions:

∆u
B= ( 2.1)
∆σcon

Where ∆u = change of pore water pressure due to change of

confining pressure

∆σcon = change in confining pressure

A “B” value of 0.97 or higher was obtained and a constant head

permeability test was conducted with a hydraulic gradient of approximately 10 in

the study.10

When soil specimens are saturated, water is assumed to fill up the voids in

the soil skeleton and water becomes connected. Although the bulk modulus water

changes with temperature and confining pressure, it is much higher than that of

soil. If a soil sample is saturated and under undrained conditions, the change in

confining pressure will be carried mainly by the water rather than by the soil, as

demonstrated by high “B” values. For a CLSM samples, the skeleton has a

relatively higher bulk modulus, resulting from the bonding strength of the

hydration products. It can carry a portion of the change of confining pressure.

Only under high pressure can water enter certain small sized voids in the

hydration products. In Bhat’s study, a backpressure around 550 kPa was used and

it was higher than the strength of certain CLSM mixtures.10 The confining

pressure was not provided in that study. The pore structure of CLSM specimens
30
can be damaged by such a high pressure. A hydraulic gradient of 10 can seldom

be reached in field conditions. Precautions must be taken when evaluating

permeability values from different sources. CLSM mixtures may best be treated

as rock and the change of “B” value should be checked with the change in

confining pressure, rather than using the “B” value itself.

2.5.2.3 Shear Strength

The main objective of using CLSM is as a viable alternative to

conventional compacted fill. As such, it is natural to characterize the engineering

properties of CLSM using geotechnical parameters familiar to field engineers.

Both the triaxial shear test and direct shear test are commonly performed in soils

laboratories, using ASTM D 3080 for the direct shear test and USACE EM 1110-

2-1906 for the consolidated-drained triaxial shear test. The applicability of these

methods to CLSM deserves significant attention.

There have been only a few published studies focusing on the triaxial

shear strength testing of CLSM.10,11,24,43,45,57,59 The reported internal frictional

angles from triaxial compression tests range from 20 to 40 degrees. There are

even fewer published results of direct shear testing studies.16,24,60 The possible

reason is the difficulty in preparing direct shear samples.

2.5.2.4 California Bearing Ratio (CBR) and Resilient Modulus

As first proposed and used in the 1920s by O. J. Porter, an engineer of the

California State Highways Department, CBR is a penetration-type test developed

31
to evaluate the strength of subbase and subgrade materials. Through years of

development, AASHTO T 193, “The California Bearing Capacity”, is now a

standard method. To evaluate the potential of using CLSM as a subbase or

subgrade material, researchers have frequently evaluated CLSM mixtures by

measuring their CBR values.11,18,60-63

The resilient modulus is defined as the initial tangent modulus of the

stress-strain curve in a triaxial test. After conditioning, a deviator stress, ∆σ1 ,

approximating the working stress, is repeated applied until certain criteria are met.

This test was initially used to evaluate pavement subgrades. The resilient

modulus of CLSM can be evaluated if it is to be used as a subgrade material.

Currently, AASHTO method T 274, “Resilient Modulus of Unbound Granular

Base/Subbase Materials and Subgrade Soils (SHRP Protocol P46)”, is commonly

used. There have been very few studies on the resilient modulus of CLSM

mixtures.

2.5.2.4 Settlement and Consolidation

CLSM may reduce its volume as it releases its free water and entrapped

air through consolidation of the mixture. The consolidation of soil consists of

primary and secondary consolidation. Primary consolidation is mainly caused by

seeping of water. Secondary consolidation is caused by rolling, slipping, sliding,

to some extent by crushing at the particle contact points, and elastic distortions.

Similarly, the initial settlement of CLSM mixture within several hours after

placement maybe regarded as primary consolidation. The further reduction of

32
CLSM volume due to applied load is similar to secondary consolidation of soil.

Because of the relatively strong and rigid skeleton of CLSM after hardening,

consolidation is expected to be very small and corresponding settlement of

backfill will be negligible.

ASTM D 2435, “Standard Test Method for One-Dimensional Properties of

Soil”, can be used to measure the consolidation of CLSM mixtures. The test can

estimate both the rate and total amount of settlement of CLSM used in various

applications. For Type III CLSM mixture, the coefficient of volume

compressibility is in the range of compacted dense gravel fill.24 For non-air-

modified CLSM mixtures, the coefficients are expected to be significantly

smaller.

2.5.2.5 Drying Shrinkage

High water-cement ratio and high water content are two factors known to

cause excessive drying shrinkage in concrete. Although CLSM has higher water

cement ratios and higher water contents than concrete, limited research on drying

shrinkage has shown that CLSM may not exhibit much more shrinkage than

concrete. Typical reported linear-shrinkage values of CLSM are in the range of

0.02 to 0.05 percent, which is similar to concrete with low drying shrinkage.1 The

shrinkage and expansion of CLSM tend to continue varying throughout testing.

Gandham et al found that the maximum shrinkage and expansion values of CLSM

were generally less than the acceptable limit established for concrete.41 Lucht

also showed that shrinkage of CLSM was minimal when a shrinkage-ring method

33
was used.13 A shrinkage ring is often used to measure the cracking of concrete

cast around a steel ring. This approach represents 100% restraint and can be used

to assess different materials and mixtures.

Most published studies have used the conventional concrete method to

measure shrinkage of CLSM specimens.41 This method specifies embedding gage

studs at both ends of the specimens and measuring the length change. Careful

handling of the shrinkage prisms during form removal and subsequent

measurements are required. CLSM specimens could be damaged when using this

approach because of the lower strengths of CLSM. Thus, this approach may not

be appropriate for CLSM. Lucht have also used the shrinkage ring method, which

is not an adopted standard.13

2.5.2.6 Thermal Conductivity

Moisture content and dry density are two major factors affecting thermal

conductivity of CLSM. Less important factors include mineral composition,

particle shape and size, gradation characteristics, organic content and specific

gravity. Low density, air-modified CLSM (Type III) is particularly suitable for

pipe backfill because of its enhanced insulating properties. The choice of thermal

conductivity characteristics depends on the specific application.

Methods to measure thermal conductivity of CLSM include ASTM D

5334, “Determination of Thermal Conductivity of Soil and Soft Rock by Thermal

Needle Probe Procedure”, and ASTM C 177, “Steady-State Heat Flux

Measurements and Thermal Transmission Properties by Means of the Guarded-

34
Hot-Plate Apparatus”. Institute of Electrical and Electronics Engineers (IEEE)

Standard 442-1981 has also been used to measure thermal resistivity of CLSM.27

2.5.2.7 Excavatability

To compete with conventional compacted fill in many applications, CLSM

must be excavatable at later ages. Researchers often use unconfined compressive

strength of CLSM as the index of excavatability.6 In general, CLSM with a

compressive strength of 350 kPa or less can be excavated manually and CLSM

with the range of 690 to 1400 kPa can be removed by mechanical equipment,
2
such as backhoes. Engineers in Hamilton County, Ohio, specify a Removability

Modulus to estimate future excavation of CLSM. It is a function of both the

thirty-day unconfined compressive strength and the dry unit weight. If the

calculated value of the Removability Modulus is less than 1.0, the specific

mixture is deemed to be removable in the future.

To predict excavatability, it is important to predict the long-term strength

development of CLSM. When foundry sand was used in CLSM, the unconfined

compressive strength increased up to 30% from 28 days to 91 days. 10 When the

cement content was high, the strength of CLSM more than doubled from the age

of four to thirteen weeks.64 . Fly ash types and contents also affect the ultimate

strength of CLSM.

CLSM mixtures with high quantities of coarse aggregate may become

very difficult to remove manually even when its strength is low. On the contrary,

Type III CLSM with strengths of 2.1 MPa can be removed with a backhoe.1

35
There is no existing laboratory test method to estimate the actual

excavatability/digibility of CLSM in the field. Actual field excavation is needed

to assess the ease of CLSM removal.

2.5.2.8 Freezing-and-Thawing Resistance

Freezing-and-thawing resistance is an important durability parameter for

construction materials. There have been several laboratory and field studies on

the freezing and thawing resistance of CLSM.4,11,17,57,65 Type II CLSM broke into

pieces about the size of a hand when it was placed in a zone of total water

saturation and subjected to severe winter freezing and temperatures below –18

°C.4 Without water saturation, Type II CLSM appears to perform well under

freeze-and-thaw conditions in the field.4 Although CLSM was considered a

suitable backfill material, Gress proposed that the top 5 to 150 mm of CLSM

mixture fill be replaced by a frost heave-compatible material for a uniform

heaving of pavement and trench after 24 hours of CLSM placement.65 Type III

CLSM and Type II CLSM with a high cement content (10%) are recommended

for construction applications where severe freezing is expected. According to

Hoopes, Type I CLSM specimens failed in only five cycles during the testing

process of ASTM D 560.24 Type III CLSM mixtures with 21% and 30% air

contents performed satisfactorily when they were evaluated with ASTM D 560.

Conventional concrete test methods for freezing-and-thawing resistance

have been found to be severe for CLSM. The low strength and the high

permeability of CLSM make it susceptible to damage under testing conditions for

36
concrete. The high permeability allows water to enter CLSM samples easily and

ice lenses to form inside the specimens. As a result, the material is damaged

internally. This phenomenon is very similar to the frost heave of soil. The

internal hydraulic pressure generated during water freezing also damages the

skeletal structure of CLSM. Nantung have modified existing concrete test

methods by exposing CLSM specimens to less severe environments to simulate

field conditions.17 ASTM D 560 is a method to measure the ability of soil-cement

mixtures to resist freezing-and-thawing cycles. Because of the similarity of

CLSM and soil-cement, the method is beginning to gain acceptance in the

research community.45,65 This method is much less severe than those for concrete

and requires only 12 freezing-and-thawing cycles.

2.5.2.9 Erosion

There is no standard procedure for testing the erodibility of CLSM. Dolen

have tested erodibility according to the guidelines of Corps of Engineers

publication REMR-GT-15.59 Tap water is injected through a formed 3.2-mm

diameter hole, lengthwise through the center of the cylindrical test sample. The

water has a pressure of approximately 275 kPa and the test is continued for 28

days. Then, the total mass of eroded material is determined. Specimens with

sufficiently high strength have demonstrated an erodibility where there was a

mass loss of approximately 0.03%, based on eroded material retained on the 75

µm (No. 200) sieve after 28 days.

37
Krell found that flowable fly ash resisted erosion in both its plastic and

hardened states.4 A tank was constructed with water being pumped continuously

across a sample. This test was developed to compare materials and determine

how they affect water quality. CLSM was shown to be superior to the other

backfill materials. This property eliminates the need for silt curtains, containment

by weirs, and other environmental protection systems when CLSM is used as a

backfill under water.

Type I CLSM was also used in Burlington, Iowa to prevent erosion. 6 A

CLSM mixture with unconfined compressive strength of 3,448 kPa (500 psi) was

used to fill the voids of riprap in a drainage ditch. The voids in front of and

behind the concrete floodwall along the Mississippi River were filled with CLSM

having a compressive strength of 690 kPa (100 psi). Similar material was also

used as an erosion control mat along the Iowa River.

2.5.2.10 Leaching and Environmental Impact

With more and more waste materials or by-product materials, such as

foundry sand and fly ash, introduced into CLSM mixtures, some environmental

concerns have been raised. Heavy-metal ions may leach into the environment

more easily from CLSM than from concrete because of the higher permeability of

CLSM. There are two approaches to assess this problem. One approach is to test

the leachate of source materials and the other is to observe the long-term field

performance of actual CLSM applications. Because CLSM is a relatively new

technology, research has mainly focused on the first approach.

38
The Extraction Procedure (EP) toxicity method, Toxicity Characteristic

Leaching Procedure (TCLP) method, American Foundry Society (AFS) method,

and ASTM method are well known laboratory test procedures to determine the

leaching characteristics of waste materials. Heavy metals include arsenic, barium,

cadmium, chromium, lead, mercury, selenium, and silver. Bhat used “Bio-assay”

testing on foundry sand to predict the effects of foundry sand components,

including organics.10

Bhat also measured the pH values and toxicity of expressed pore solutions

from both source materials and CLSM mixtures. Only one of the eleven mixtures

evaluated was found to be toxic.10 Naik et al. and Gandham et al. studied the pore

solution leachate characteristics of different mixtures using the TCLP test (EPA

SW-846, Method 1311).39,41 They found that the toxic contents of the mixtures

were well below the EPA leachate standards.

2.5.2.11 Corrosion

Although corrosion of metal pipes buried in CLSM has not yet surfaced as

a serious problem in field applications, corrosion is a long-term process that may

appear as a problem in the future. Folliard et al. have discussed the potential

mechanisms of corrosion likely to occur in CLSM.26

Most information on the corrosion of metals in CLSM mixtures has been

obtained from laboratory testing results. Abelleira et al. found that CLSM

significantly improved corrosion performance of buried iron. Metal coupons

were buried in CLSM mixtures and were then submerged under corrosive water

39
containing sulfate,, chloride and bicarbonate ions.60 Ramme et al. tested the

resistivity of CLSM mixtures and found that the corrosion potential of CLSM (fly

ash slurry) was significantly less than that of typical soils used for trench

backfill.61 Krell considered that the alkalinity of fly ash and cement could inhibit

the rusting of iron and steel.4 The pH of extruded pore solution from CLSM

mixtures were also tested.

Several ASTM test methods may be utilized to evaluate the corrosion

potential of steel contacting with CLSM mixture. They include ASTM G 1,

“Standard Test Method for Preparing, Cleaning, and Evaluation Corrosion Test

Specimens”, ASTM G 51, “Standard Test Method for Measuring pH of Soil for

Use in Corrosion Testing”, and modified ASTM G 59, “Standard Practice for

Conducting Potentiodynamic Polarization Resistance Measurements”.

Although the corrosion potential of metals in CLSM is significantly

decreased with respect to soil, there may be a concern with galvanic corrosion.

Galvanic cells may form when a pipe contacts both soil and CLSM, two different

environments. If CLSM is only used as utility bedding and the upper portion of

the backfill is conventional compacted fill, a large galvanic cell will be formed

with a small anode and a large cathode. Even though these pipes are buried in

CLSM, a galvanic cell will form between sections in CLSM and sections in the

surrounding soils.

40
2.5.2.12 Utility Degradation

Degradation of non-metallic utilities in CLSM has not been reported as a

problem in field applications. Some plastics can degrade in high pH solutions.

Because CLSM is basic in nature, the long-term durability of various utilities in

CLSM needs to be investigated.56

2.6 ECONOMICS

A recent survey showed that most state DOTs considered the cost of

CLSM as the largest barrier for widespread use.25 Prices as high as $130 per

cubic meter were reported. When filling the same trench, CLSM was slightly

more expensive than conventional granular backfill.66 At the same time, because

of the flowability of CLSM, it is possible to reduce the trench dimensions and

reduce the cost below that of conventional backfill. Bhat et. al stated that savings

in cost could be achieved when waste foundry sand was used as fine aggregate in

CLSM mixtures.10

2.7 CURRENT PRACTICE

2.7.1 Case Histories

2.7.1.1 Introduction

In general, an embankment refers to a volume of earthen material that is

placed and compacted for the purpose of raising the grade of a roadway (or

railway) above the level of the existing surrounding ground surface. A fill refers

41
to a volume of earthen material that is placed and compacted for the purpose of

filling in a hole or depression.

To categorize the typical applications of CLSM, the following words are

defined. These definitions are used in the discussion to follow in this dissertation.

Bridge Replacement

Bridge replacement refers to the use of CLSM as construction material in

the cases of converting deteriorated bridges. Those bridges are not demolished

but changed into culverts or other structures by using CLSM to fully fill the

original space beneath the bridge.

Backfill

Backfill refers to the use of CLSM in filling the openings formed during

the installation or repair of utility. The material is only utilized to cover the utility

components and transfer loads above. No structural support is provided to

utilities in this case.

Structural Fill

Structural fill refers to the use of CLSM in construction or repair, where

CLSM is to act as a load-bearing and/or load-transferring unit. Applications

include foundation support, pavement bases, bridge approaches, soil retaining

structures, and other components.

Insulation and Isolation Fill

42
Insulation and isolation fill refers to the use of CLSM as insulation and/or

isolation material to take the advantage of its appropriate thermal and acoustic

properties.

Utility Bedding

Utility bedding refers to the use of CLSM as bedding material in utility

installation or repair to provide uniform support. Strictly speaking, utility

bedding belongs to the category of structural fill, but is listed separately here to

indicate its importance.

Erosion Control

Erosion control refers to the use of CLSM in anti-erosion applications,

such as filling the voids of riprap.

Void Filling

Void filling refers to the use of CLSM in the practice of filling voids from

settlement or abandoned space. The material is not intended to bear or transfer

significant loads.

Figure 2.1 illustrates the typical applications of CLSM.

43
2.7.1.2 Backfills

One common application of CLSM is to replace conventional compacted

backfill. The high flowability of CLSM eliminates or decreases the dependence

on accessibility of compaction of the equipment, crew productivity, and degree of

compaction verification. Granular or site-excavated backfill, even when

compacted properly in the required layer thickness may not achieve the

uniformity of CLSM. Folliard et al. found that the production rates when placing

CLSM could be substantially higher than conventional compacted backfill.25

CLSM

A B C

Figure 2.1: Applications of CLSM. Case A: backfill; Case B: utility


bedding; Case C: structural fill.

The following are case studies where CLSM has been used as backfill:

Case history 1: Because of severe settlement problems with soil backfill,

the city of Peoria, Illinois, tested CLSM for backfilling utility trenches in 1988.1

As a result, the city changed its backfilling procedure to require the use of CLSM

on all street openings.

44
Case history 2: At one location in Iowa, a laid-up limestone block
67
retaining wall bulged out, indicating the initial stage of failure. A new “H”-pile

wall with wood planks was built in front of the old wall and flowable mortar was

used to fill the gap between the two walls.

Case history 3: In a Maine DOT project, “Evaluation of Flowable Fill for

Use in Highway and Bridge Applications”, it was found that unshrinkable backfill

(25 kg cement per cubic meters, conventional concrete aggregates, 160 to 200 mm

slump, usually air entrained, maximum 28 days compressive strength of 0.4 MPa)

satisfied the technical requirements for utility cut restorations and this material

was cost effective compared to properly compacted and inspected granular

backfill.68

Case history 4: In 1988, during reconstruction of the New England

Thruway (I-95) between Pelham Parkway and Bartow Avenue, Albany, New

York state, lightweight CLSM (672kg/m3 , minimum 28 days compressive

strength of 0.69 MPa) was specified for backfilling certain utilities, which was

placed above a soft organic silt containing a strata of highly organic peat and

organic clay in a random pattern.47

Case history 5: During the construction of Edison’s 1350 MW Belle River

plant, 122,570 m3 of flowable fly ash were used in trench fill and 84,267 m3 was

used in the power house area. The project estimates indicated that more than $1

million were saved through the use of the cement-stabilized ash.3

45
Case history 6: In one tilt-up construction project in Denver, it was

estimated that approximately two days of construction time were saved by using

CLSM for floor construction. 69

Case history 7: The Minnesota District #9 Maintenance Department used

CLSM to fill voids under bridge pier footings in the Robert Street Bridge in St.

Paul, Minn. This avoided building cofferdams or replacing the bridge.

Approximately 126 m3 (165 yd 3 ) of K-Krete was placed under two piers at a cost

of $107,000.6

Case history 8: CLSM provided the solution when severe spring flooding

washed all of the fill from above and around a multiplate steel arch pipe in

Hutchinson County, South Dakota. The pipe was repaired in nine days at a cost

of $12,000. Replacing the pipe was originally estimated to cost $40,000. 6

Case history 9: In Des Moines County, Iowa, CLSM was used to fill a

void along a wooden culvert, where access for conventional repairs was not

available. 6

Case history 10: Controlled-density fill was utilized to completely fill an

underslab void for the US Navy at Rough & Ready Island, Stockton, California.

The total cost was less than 20 percent of the amount authorized by the owner for

placing and compacting conventional granular backfill.70

Case history 11: A hollow sidewalk cavity containing locker room

facilities was backfilled with CLSM in downtown Milwaukee during the summer

of 1984.71

46
Case history 12: CLSM was used in backfill excavations to remove oil-

contaminated soil adjacent to foundations of existing structures. CLSM was also

placed underwater at Harbor in Plymouth, Massachusetts.72

2.7.1.3 Bridge Replacement

The Iowa DOT developed a creative process to replace abandoned

bridges. This process involved installing metal culverts, pipes, or reinforced box

culverts under the bridges. Soil was used as forms at both ends of the bridge.

The space between the bottom of the bridge and the top of the culvert was

backfilled with CLSM.23

Case history 13: Iowa DOT engineers developed flowable mortar, which

includes sand, portland cement, fly ash and water, to convert 10 bridges into

culverts. US Route 30 was kept open to normal traffic most of the time. The cost

was one-third less than that of routine repair (1985).23

Case history 14: Six narrow bridges were replaced in a 9-mile length of U.

S. Route 30 between Woodbine and Logan in Harrison County, Iowa. Four of

these bridges were modified by using flowable mortar with concrete pipe culverts.

One was modified by using flowable mortar with a reinforced-concrete box

culvert. One was replaced with a new bridge. The installation of pipe culvert

with CLSM cost only one-third of a reinforced-concrete box culvert and one-

fourth of conventional bridge replacement.67,6

47
2.7.1.4 Structural Fill

Because of the high strength and low-consolidation potential of CLSM

compared with compacted fill, CLSM is used widely as a structural fill. In

structural fill, the compressive strength of CLSM can vary from 0.69 MPa to 8.3

MPa (100 psi to 1200 psi) depending upon the application. The structural

coefficient of CLSM layer is estimated to range from 0.16 to 0.28 for compressive

strengths from 2.8 MPa to 8.3 MPa (400 psi to 1,200 psi).1

There are five primary causes of bridge-approach settlement, which causes

“the bump at the end of the bridge”73 :

1. time-dependent consolidation of the embankment foundation soil under

the embankment,

2. time-dependent consolidation of the approach embankment,

3. poor compaction of abutment backfill caused by restricted access of

standard compaction equipment,

4. erosion of soil at the abutment face, and

5. poor drainage of the embankment and abutment backfill

State DOTs are especially interested in the application of CLSM in bridge

approach construction because use of CLSM may eliminate causes No. 3 and No.

4. Possibly, if CLSM is used for the bridge approach embankment, cause No. 2

can be avoided.

Case history 15: Near Boone, Iowa, 2,145 cubic meters of CLSM were

used to provide proper bearing capacity for the footing of a grain elevator that was

built on load bearing soil.7

48
Case history 16: In June 1979, a small peninsula about 48.8 meters long

and 9.1 meters wide was constructed of flowable fly ash at Edison's Monroe

Power Plant.3

Case history 17: CLSM was used to fill between a spread footing and the

bearing soil during a cast-in-place hospital parking garage construction. It was

also used for updating the manufacturing plant without interrupting the

manufacturing process. The mix was line pumped over 305 m during some

placements.74

Case history 18: In a micro tunneling project in Newark, California,

CLSM was used to stabilize the soil surrounding the sheet-piled shaft that would

be used to launch a micro tunnel boring machine. The use of this fly-ash-based

CLSM greatly improved the stability of the soils and increased the safety of the

shaft during the launch. The use of CLSM provided a saving of more than

$100,000 in the Newark project.75

Case history 19: In July 1995, a contractor placed more than 306 cubic

meters (400 cubic yards) of CLSM as backfill for the abutments of a bridge

located along Colorado State Highway 135 near Crested Butte, Colorado.

Approximately 3060 m3 (4000 yd 3 ) of CLSM was used on a large commercial

distribution center in Loveland, Colorado for foundation wall backfill in 1991.69

Case history 20: In the Muroran Bay Bridge project, a man-made island of

67 m diameter was built by underwater placement of a self-hardening slurry on a

soft seabed. From October 1988, 53,600 m3 of slurry, containing 46.1 x 106 kg of

fly ash, 22.2 x 106 kg of volcanic ash, and 2.6 x 106 kg of cement, was filled into a

49
steel-pile cofferdam during a three-month period. Observations included 1)

reduced lateral earth pressure on the cofferdam walls, resulting in smaller bending

moments and lateral displacements; and 2) higher strength of the island for

subsequent construction work. In spite of the low temperatures in winter, heat

supplied from the hydration of the slurry kept the fill temperature above 10 ºC and

strength development of the mixture was higher than expected.76

Case history 21: Pockets of soft material of an underlying soil layer were

excavated and then backfilled with flowable fly ash (Class F ash). The next

morning, the contractor continued his work on building the six-story apartment

building.4

Case history 22: CLSM mitigated roadway settlement above buried

culvert and pipes in Wyoming.21

Case history 23: The Oklahoma DOT and Oklahoma State University

School of Civil and Environmental Engineering conducted (1998) a research

program to evaluate options to minimize approach embankment settlement.77 The

research program involved construction of three new bridges on U.S. 177 north of

Stillwater, Oklahoma. They found that the use of CLSM as construction material

for approach embankment was a simple and reasonably cost-effective method to

reduce the potential for “the bump at the end of the bridge”.

Case history 24: Flowable fly ash was used to save Detroit's Northeast
78
Raw Water Tunnel, Detroit, Michigan, 1985-6. CLSM was used to support the

tunnel before the grouting was finished.

50
2.7.1.5 Utility Bedding

Inadequate bedding support is known to reduce the service life of pipes or

utilities. Improper bedding preparation can decrease the desired pipe carrying

capacity or cause excessive backfill settlement. There are several methods to test

the inherent strength of pipe sections. These methods include the two-edge

bearing test, the three-edge bearing test, the sand bearing test, and the Minnesota

bearing test. The three-edge bearing method is the simplest to perform. The ratio

of the strength of a pipe under any stated condition of loading, whether in the field

or in the laboratory, to its strength by the three-edge bearing test is called the load

factor or strength ratio for the stated condition and is designated as Lf. There are

four classes of bedding.79

Impermissible bedding (Class D), Figure 2.2, is the case where little or no

effort is taken to shape the foundation to fit the lower part of the conduit exterior

or to refill all spaces under and around the conduit with granular materials, at least

partially compacted.

51
Figure 2.2: Class D: impermissible beddings. load factor = 1.1. (From Soil
Engineering, 4th Edition, M. G. Spangler and R. L. Handy, Harper &
Row, Publisher, New York, 1982)

Ordinary bedding (Class C), shown in Figure 2.3, is the case where the

conduit is bedded with “ordinary” care in an earth foundation, pre-shaped to fit

the lower part of the conduit exterior with reasonable closeness for a width of at

least 50% of the conduit breadth. The remainder of the conduit is surrounded to a

height of at least 0.15 m above its top by granular materials that are shovel-placed

and shovel-tamped to completely fill all spaces under and adjacent to the conduit.

All this work must be done under the general direction of a competent engineer.

52
H > 7.3 m-13 mm per 305 mm of H

Figure 2.3: Class C ordinary beddings. load factor = 1.5 (From Soil
Engineering, 4th Edition, M. G. Spangler and R. L. Handy, Harper
& Row, Publisher, New York, 1982)

First class bedding (Class B), shown in Figure 2.4, is the case where the

conduit is carefully bedded on fine granular materials in an earth foundation that

is carefully pre-shaped by means of a template to fit the lower part of the conduit

exterior for a width of at least 60% of the conduit breadth. The remainder of the

conduit is entirely surrounded to a height of at least 0.3 m above its top by

granular materials that are carefully placed to completely fill all spaces under and

adjacent to the conduit. The granular materials are thoroughly tamped on each

side and under the conduit as far as practicable in layers not exceeding 0.15 m in

thickness.

53
Figure 2.4: Class B first class beddings. load factor for ditch conduits =1.9 (From
Soil Engineering, 4th Edition, M. G. Spangler and R. L. Handy,
Harper & Row, Publisher, New York, 1982)

Concrete-cradle bedding (Class A), Figure 2.5, is the case where the lower

part of the conduit exterior is bedded in plain or reinforced concrete of suitable

thickness under the lowest part of the conduit and extending upward on each side

of the conduit for a distance not greater than one-fourth the outside diameter.

When CLSM is used for utility bedding, there is no or only minimal preparation

needed. Labor-intensive compaction in the haunch zone is eliminated. The

flowability and strength characteristics of CLSM may make this material a cost-

effective alternative to conventional bedding materials.

Case history 25: The US Bureau of Reclamation (USBR) used CLSM

(soil-cement slurry) in 1964 as bedding material for 381- to 2438- mm diameter

concrete pipe along the entire Canadian River Aqueduct Project, which stretches

515 km (322 miles) from Amarillo to Lubbock, Texas. It was estimated that soil-

54
cement slurry reduced bedding costs by 40 percent. Production increased from

122 linear meters to 305 linear meters of pipe placed per shift.2

Figure 2.5: Class A: concrete-cradle bedding. load factor 2.25 to 3.4. (From
Soil Engineering, 4th Edition, M. G. Spangler and R. L. Handy,
Harper & Row, Publisher, New York, 1982)

Case history 26: In the research project FHWA/OH-93/014, Brewer

installed a pipe arch culvert (381 cm × 241 cm, 2745 cm long with only 840 mm

of CLSM top cover) in Paulding County, 1989 and a long span arch culvert (10.5

m long) in Cuyahoga County in 1990. He found that: 1): the minimum wall

thickness required for the structural capacity of flexible culverts, pipe arches and

long span arches, could be reduced when using CLSM-CDF as the backfill

because of a reduction in required handling stiffness; 2) a relationship was

established between the horizontal and vertical stress of CLSM-CDF backfill,

which was accompanied by the recognition of CLSM-CDF's cohesion

55
development; 3) more triaxial shear information is needed for wider CLSM-CDF

strength ranges, (0.35 to 10.4 MPa (50 to 1500 psi)) to help develop more

accurate determination of CLSM-CDF's cohesion development; and, 4) initially,

conventionally placed backfill exerts less vertical and horizontal load on a culvert

than does a CLSM-CDF backfill, but after CLSM-CDF's cohesion buildup the

reverse is true.20

Case history 27: Quick-setting CLSM was used for replacement of

approximately 305 m of 914-mm water main under numerous train tracks in

downtown Seattle.74

Case history 28: In 1995, for a project at Kent County International

Airport, CLSM was used to backfill a 1.5 m diameter rigid concrete pipe in a

narrow 3.1 m trench, where the material was placed up to the pipe springline. The

culvert pipe has been inspected three times since installation and it is performing

well.80

Case history 29: CLSM was specified as bedding material for a 32,000-m

pipe drainage system at the new Denver International Airport (constructed

between 1991 and 1993). The flowable backfill was designed to completely

surround the pipe and extend a minimum of 152 mm above the top of the pipe.69

Case history 30: A roadway over a myriad of petroleum pipes was needed

to accommodate construction traffic for a project at the SOHIO petroleum plant in

Toledo, Ohio. A rectangular area was excavated and backfilled with a monolithic

mass of CLSM to totally envelope the pipes. After a three-day curing period, a

56
100-mm granular layer was placed over the area as a wearing course for the

roadway. 6

2.7.1.6 Erosion Control

Erosion of soil or backfill is known to cause riverbank failure, bridge

abutment movement, bridge approach settlement, and other damage. The use of

CLSM may help eliminate or alleviate these problems.

Case history 31: 1. Two projects in Burlington, Iowa, used CLSM to

prevent erosion. 2. Wapello, Iowa's business district, utilized CLSM as an

erosion control mat to avoid further erosion of the Iowa River bank.6

Case history 32: Riprap was put into the erosion cavity and the spaces

between the rocks were filled with flowable mortar in Iowa counties.7

2.7.1.7 Void Fill

Because of its flowability, CLSM is an excellent material to fill the voids

inside and/or between objects. In practice, CLSM is used widely to fill tunnel

shafts, sewers, basements, and other underground structures.

Case history 33: CLSM was used to fill an abandoned sewer and a

deserted tunnel that passed under the Menomonee River in downtown Milwaukee.

The material was reported to flow 72 meters and 92 meters, respectively. Only

four hours were needed to fill an exploratory shaft 36.6 m deep, 3.7 m in diameter

with a 9.2 m long branch tunnel. A total of 602 cubic meters of CLSM were

cast.1

57
Case history 34: To verify the practicality of the hydraulic backfilling

concept, a small-scale field injection of about 765 m3 of ash-bentonite-water

slurry into an abandoned mine was completed at the Fairfax mine in Preston

County, West Virginia. It demonstrated that the hydraulic backfilling method is a

practical solution to ash disposal with collateral benefits in preventing Acid Mine

Drainage (AMD) and subsidence.52

Case history 35: CLSM was chosen to fill the an abondoned pipeline

beneath a critical segment of Interstate 70 near Officers Gulch, immediately east

of Copper Mountain, Colorado.69

Case history 36: Flowable fly ash was used to fill abandoned gasoline

tanks, tunnels, pits, and sewers.4

Case history 37: 1.CLSM was used to fill the a void under a 30.5-meter

long metal culvert undermined by water passing beneath. 2. At the University of

Iowa at Iowa City, old stream tunnels were filled with CLSM. 3. Many

abandoned fuel tanks have been filled with CLSM. 7

Case history 38: 1. Wisconsin Electric filled two obsolete steam service

tunnels in downtown Milwaukee in December 1983. One tunnel was 1.8 m in

diameter by 88.5 m long (1.8 x 88m) and the other was a 1.5 m high by 1.2 m

wide ellipsoid section. 2. A number of abandoned steam utility facilities in the

Menomonee River Valley of Milwaukee were filled using CLSM. 3. The

abandoned Indian Creek Parkway Sewer was filled with CLSM.71

Case history 39: In LaSalle, Illinois, 306 m3 of CLSM was pumped to fill

the basement of a building. In Seattle, Washington, 19,150 m3 voids over each

58
bus station in a tunnel was filled with CLSM. CLSM has also been used to fill

abandoned underground tanks.1

Case history 40: CLSM was used to fill old basements full of loose

materials including bricks, broken concrete, and timber.4

Case history 41: Iowa DOT filled two 7.6-m3 and one 3.8-m3 abandoned

fuel tanks with approximately 20 m3 of CLSM near Ames in 1987. The project

was completed at a cost of $1,140. The cost of removing the tanks was estimated

at approximately $8,000.6

Case history 42: With CLSM, the City of Des Moines repaired the grade

under the street pavement and filled an undermined culvert for Pope John Paul's

visit.7

2.7.1.8 Environmental Protection

Case history 43: In Seattle, Washington, a CLSM mixture containing a

contaminated soil was designed to encapsulate it during the construction of a new

manufacturing plant.74 The alternative choice of the owner is to haul the

contaminated soil to a landfill over 640 km away,

2.8 RESEARCH NEEDS

2.8.1 Introduction

Although CLSM has gained popularity in recent years, there are several

obstacles for which some engineers are reluctant to specify it for construction

projects. The purpose of this section is to identify these challenges and


59
corresponding research needed for CLSM. One important obstacle is that

engineers are not familiar with this material. Education is extremely important

for increased use and acceptance of CLSM.

2.8.2 Research Needs

The following are some important issues that deserve additional emphasis

in current and future research project.

2.8.2.1 Standard Test Methods

To date, there are only five ASTM test methods and no AASHTO method

on CLSM. The five ASTM test methods are:

• ASTM D 4832-95, “Standard Test Method for Preparation and Testing

of Controlled Low Strength Material (CLSM) Test Cylinders”;

• ASTM D 5971-96, “Standard Practice for Sampling Freshly Mixed

Controlled Low-Strength Material”;

• ASTM D 6023-96, “Standard Test Method for Unit Weight, Yield,

Cement Content, and Air Content (Gravimetric) of Controlled Low Strength

Material (CLSM)”;

• ASTM D 6024-96, “Standard Test Method for Ball Drop on Controlled

Low Strength Material (CLSM) to Determine Suitability for Load Application”;

• ASTM D 6103-97, “Standard Test Method for Flow Consistency of

Controlled Low Strength Material (CLSM)”.

60
Obviously, there is a lack of standard test methods for CLSM. Only a few

CLSM properties are routinely measured by state DOTs and commercial testing

laboratories, and even these properties are measured with different test methods.25

These different methods may yield different results. It is important to identify

suitable methods if they exist, and to create new methods, if necessary, to

characterize the relevant properties of CLSM.

2.8.2.2 Effective Use of By-product Materials in CLSM

One outstanding characteristic of CLSM is its ability to consume various

and large amounts of by-product materials. Different by-product materials are

used in different locations throughout the United States. These materials include

fly ash of different carbon contents, bottom ash, crushed concrete, high fines, and

crushed recycle glasses. Unfortunately, there is no general agreement among

CLSM users about how to best use and specify these materials, or how to address

specific issues that may arise from their inclusion in CLSM.25 Procedures or test

methods to identify suitable by-product materials to be used in current and future

CLSM applications are needed.

2.8.2.3 Excessive Long-Term Strength Gain

A technical and practical concern is the uncertainty of long-term strength

gain for CLSM because the majority of CLSM applications may eventually

require excavation (e.g., to repair or replace utilities). Strength of CLSM is

considered as the most important factor that determines the excavatability or

61
diggibility of CLSM mixtures placed in field. Density and aggregate type also

influence the actual ease of excavation. There is little published data or

information on this aspect.

To study the long-term strength gain of CLSM, it is important to study the

effects of constituent materials and different mixture proportions. For instance,

CLSM containing fly ash demonstrates more long-term strength gain because of

the pozzolanic and/or cementitious reaction of fly ash. Because CLSM is usually

made with locally available materials and the properties of these materials may

vary significantly, it is necessary to perform accelerated or long-term testing to

predict the long-term strength gain of CLSM mixtures.

2.8.2.4 Durability Aspects of CLSM

It is essential to study the durability characteristics of CLSM as a

construction material. There have not yet been significant durability problems

identified in field applications.25 But this may be due to the relatively short

period of time that CLSM has been used. Issues of particular interest include

freezing and thawing, corrosion, and chemical attack.

2.8.2.5 Environmental Impact

Fly ash, foundry sand, and other by-product materials are often used in

CLSM. These materials may contain heavy metals and other potentially harmful

materials that may leach into the surrounding environments. The leaching

potential and consequent results deserve further attention. Research is needed to

62
develop suitable test approaches and to provide specifications to warrant the safe

and effective use of by-product materials in CLSM.

2.8.2.6 Applicability of Soil Test Methods

As CLSM is mainly used to replace conventional compacted granular fill,

it may be desirable to use soil testing methods to determine the properties of

CLSM. Many parameters describing soil can be used to define properties of

CLSM, such as resilient modulus, CBR values, water permeability, cohesion,

internal friction angle, consolidation and settlement, and bedding factors for

pipelines. However, the applicability of these methods should be verified in

practice.

Soil test methods that may be applied to CLSM should be simple methods,

using equipment available at most soil laboratories. Testing should only be

performed when specifically required for a given application, especially when

design parameters are needed as part of the design process.

63
Chapter 3: Materials and Mixture Proportions

3.1 INTRODUCTION

Component materials used in CLSM include fine aggregates, fly ash,

portland cement, and water. Chemical admixtures are sometimes used to obtain

different objectives, such as air-entraining and setting-accelerating. In this

chapter, materials utilized in this study are described. In addition, a description of

the research program is presented.

Thirty-eight mixes were initially cast to study the important properties of

CLSM. Additional follow-up mixtures were cast to further investigate certain

technical issues. Because of the importance of compressive strength in CLSM

specifications, particular emphasis was placed on evaluating parameters affecting

CLSM strength.

3.2 FINE AGGREGATE

3.2.1 Aggregate Type

One advantage of CLSM is its ability to accommodate waste and by-

product materials. Foundry sand and bottom ash are two of the most common by-

product materials used in CLSM, and both were included in this study. A natural

sand, typically used in concrete, was also used throughout the investigation.

64
3.2.2 Aggregate Gradation

The gradations of the three fine aggregates were determined in accordance

with ASTM C 136-96a, “Standard Test Method for Sieve Analysis of Fine and

Coarse Aggregates.” Because of the large-sized particles in the as-received

bottom ash, only materials passing the 6.35 mm sieve were used in this study.

The gradations of the three fine aggregates are shown in Table 3.1 and Figure 3.1.

For comparison, the upper and lower limits set by ASTM C 33, “Standard

Specification for Concrete Aggregates,” are also shown in the graph. Although

the river sand met the specification of ASTM C 33-93, it approached the coarse

limit of the gradation band. The bottom ash was found to be slightly coarser and

the foundry sand slightly finer than ASTM C 33 gradation limits.

Table 3.1: Aggregate gradations used in laboratory test program


Sieve # Concrete sand Bottom ash* Foundry sand
Cumulative %Cumulative % Cumulative % Cumulative % Cumulative %Cumulative %
retained passing retained passing retained passing
#4 (4.75 mm) 2.73 97.27 9.26 90.74 0.38 99.62
#8 (2.36mm) 20.81 79.19 31.70 68.30 4.26 95.74
#16 (1.18mm) 39.45 60.55 35.78 64.22 11.25 88.75
#30 (600µm) 59.11 40.89 49.66 50.34 30.46 69.54
#50 (300µm) 83.08 16.92 74.98 25.02 71.07 28.93
#100 (150µm) 96.02 3.98 87.85 12.15 96.12 3.88
#200 (75µm) 99.09 0.91 95.39 4.61 98.68 1.32
<#200 (<75 µm) 100 0 100 0 100 0
Fineness 3.0 2.89 2.14
modulus
* The bottom ash gradation reflects the particle size after “scalping” off the material retained above the 6.35
mm sieve

65
100

90

80

70
Percent passing (%)

60

50

40 River sand

30 Bottom ash
Foundry sand
20 ASTM C33 (lower limit)
ASTM C33 (upper limit)
10

0
0.1 1 10
Sieve opening size (mm)

Figure 3.1: Gradations of three fine aggregates and ASTM C 33 limits.

3.2.3 Absorption Capacity and Specific Gravity of Fine Aggregates

The absorption capacity and specific gravity of each aggregate were found

in accordance with ASTM C 128-97, “Standard Test Method for Specific Gravity

and Absorption of Fine Aggregate.” The surface-saturated dry (SSD) specific

gravity and absorption capacity of each aggregate were determined. The results

are shown in Table 3.2. The Fineness Modulus (FM) of the river sand was

slightly higher than those used in concrete. FM is an empirical factor used in

66
concrete technology to assess the fineness of an aggregate. The cumulative

percentages retained sieves: No. 100 (150 ìm), No. 50 (300 ìm), No. 30 (600

ìm), No. 16 (1.18 mm), No. 8 (2.36 mm), No. 4 (4.75 mm), and larger –

increasing in the ratio of 2 to 1, are calculated and divided by 100. The value

obtain is called FM. The absorption capacities of foundry sand and bottom ash

were found to be significantly higher than aggregates typically used in concrete.

Loss on ignition (LOI) of fly ashes were also measured.

Table 3.2: Materials included in laboratory test program*

Material Type or class


Portland cement ASTM Type I (S.G.=3.15)
Fly ash ASTM Class F (LOI = 2.9%, S.G.=2.41)
ASTM Class C (LOI = 0.37%, S.G.=2.51)
High-carbon fly ash (LOI = 14.44%, S.G.=2.09)
Fine aggregate ASTM C 33 river sand (S.G.=2.60, Absorption=1.0%, FM = 3.0)
Foundry sand (ferrous) (S.G.=2.36, Absorption=5.6%, FM = 2.14)
Bottom ash** (S.G.= 2.28, Absorption = 8.9%, FM = 2.89)
Chemical admixtures Air-entraining agent (liquid)
Accelerating admixture (non-chloride)
* Different materials were used to measure permeability and triaxial shear strength, ASTM
Type II cement and a non-classified fly ash were used.
** Bottom ash is classified as a fine aggregate because of similar particle sizes.

The aggregates used in this study were based on saturated surface-dry

(SSD) state. The following equation was used to perform moisture correction:

Wbatch=WSSD(1+p)/(1+a) (3.1)

Where, p – moisture content

A – aggregate absorption capacity


67
3.2.4 Chemical Analyses of Foundry Sand

The chemical analysis of foundry sand was performed by a commercial

analytical laboratory and the results are shown in Table 3.3. The Material Safety

Data Sheet (MSDS) from the producer indicated that this material might comprise

0 to 4 % bentonite and 0 to 4 % coal fines.

Table 3.3: Chemical characteristics of foundry sand

Chemical composition, % by weight Foundry sand


Silicon Dioxide, SiO2 85.20
Aluminum Oxide, Al2 O3 3.92
Iron Oxide, Fe2 O3 3.46
Total (SiO2 + Al2 O3 + Fe2 O3 ) 92.58
Calcium Oxide, CaO 0.79
Magnesium Oxide, MgO 0.58
Sodium Oxide,Na 2 O 0.98
Potassium Oxide, K2 O 0.17
Titanium Dioxide, TiO2 0.21
Manganese Dioxide, MnO2 0.11
Phosphorus Pentoxide, P2 O5 0.00
Strontium Oxide, SrO 0.01
Barium Oxide, BaO 0.07
Sulfur Trioxide, SO3 0.20
Loss On Ignition (LOI) 4.30

3.3 MINERAL ADMIXTURES

3.3.1 Fly Ash

The CLSM mixtures in this study included three types of fly ash, ASTM C

618 Class F, Class C and high-carbon fly ashes. Fly ash with less than 70% SiO 2

and Al2 O3 , resulting in 15-35%CaO is classified as Class C fly ash. On the other

68
hand, fly ash with more than 70% SiO 2 and Al2 O3 , and lower CaO contents, is

classified as Class F fly ash. According to ASTM C 618–94a, “Standard

Specification for Coal Fly Ash and Raw or Calcined Natural Pozzolan for Use as

a Mineral Admixture in Portland Cement Concrete”, the LOI limit maximum is

six percent for Class C and Class F fly ashes. The LOI of each fly ash was

measured in accordance with ASTM C 311-98b, “Standard Test Method for

Sampling and Testing Fly ash or Natural Pozzolans for Use as a Mineral

Admixture in Portland-Cement Concrete.” Higher LOI values in fly ash typically

reflect higher carbon contents. The high-carbon fly ash used in this study would

not typically be allowed for use in conventional concrete. High-carbon fly ash is

known to cause increased water demands and difficulties in entraining air in

concrete, but high-carbon fly ash has been allowed in CLSM mixtures by state

DOTs according a survey. 25 The LOI results of the fly ashes are shown in Table

3.2.

3.3.2 Chemical Analyses of Fly Ashes

Chemical analyses of the four fly ash used in this study were performed by

a commercial analytical laboratories, and the results are shown in Table 3.4.

Class F fly ash had high proportions of silica and alumina (a total of 85%) where

Class C had only 55% of these oxide. As a result, the Class C fly ash had a much

higher calcium oxide content. Because of the presence of free lime, some Class C

fly ashes rapidly set and harden when being mixed with water, and are considered

to be true hydraulic cements. The high-carbon fly ash had a higher calcium oxide

69
content than the Class F fly ash, but this by itself does not necessarily indicate that

high-carbon fly ash has higher reactivity than the Class F fly ash. The fly ash

used in water permeability and triaxial shear tests was also analyzed.

Table 3.4: Chemical composition of fly ashes


Chemical composition Class F Class C fly High- Fly ash
fly ash ash carbon fly used in
ash triaxial test
Silicon Dioxide, SiO2 55.24 34.40 47.29 56.41
Aluminum Oxide, Al2 O3 29.43 20.20 19.53 26.40
Iron Oxide, Fe 2 O3 5.19 5.76 5.21 3.48
Total (SiO 2 + Al2 O3 + Fe 2 O3 ) 89.86 60.36 72.03 86.29
Calcium Oxide, CaO 1.59 26.73 6.01 1.51
Magnesium Oxide, MgO 0.93 5.15 2.05 0.58
Sodium Oxide,Na2 O 0.24 1.58 2.47 0.28
Potassium Oxide, K2 O 2.18 0.36 0.83 1.57
Titanium Dioxide, TiO 2 1.44 1.32 0.75 1.43
Manganese Dioxide, MnO 2 0.03 0.06 0.04 0.03
Phosphorus Pentoxide, P2 O5 0.28 1.08 0.40 0.10
Strontium Oxide, SrO 0.10 0.39 0.31 0.12
Barium Oxide, BaO 0.07 0.60 0.30 0.06
Sulfur Trioxide, SO 3 0.38 1.98 0.38 0.12
Loss on Ignition (LOI) 2.90 0.37 14.44 7.92

3.4 CHEMICAL ADMIXTURES

3.4.1 Air-Entraining Agent

A liquid air-entraining agent (AEA), designed specifically for use in

CLSM, was used in selected CLSM mixtures. It is a viscous solution of organic

compounds. The average specific gravity of the agent is 1.0 and the average total

solids are 95%. This admixture is capable of entraining stable air contents

ranging from 15-35%.

70
3.4.2 Accelerating Agent

A liquid, non-chloride accelerating admixture, typically used in concrete,

was utilized in this study for a small number of mixtures. This agent belongs to

ASTM C 494 Type C, and the average specific gravity is 1.34. It has a solid

contents of 41.5%.

3.5 CEMENT

ASTM Type I portland cement was used in this study. ASTM Type II

cement was used only in the study of water permeability and triaxail shear tests.

Its chemical composition, provided by the manufacturer, is shown in Table 3.5.

Table 3.5: Chemical analysis of ASTM Type I portland cement (weight


percent)
Chemical Analysis Type I Type II
(%) (%)
SiO2 21.0 20.14
Al2 O3 4.9 4.54
Fe2 O3 2.3 3.73
CaO 64.8 62.11
MgO 1.7 4.08
Available alkalis as Na 2 O 0.3 0.85
Compound Composition Weight percent (%)
C3 S 62.0
C2 S 13.0
C3 A 9.0
C4 AF 7.0

3.6 MIXING WATER

Laboratory tap water was used throughout this study.

71
3.7 CLSM MIXTURE PROPORTIONS

3.7.1 Mixture Proportioning

This section summarizes the mixture proportions used in this study. There

were two distinct stages of mixing for this study. In the first stage, a total of 38

mixtures were cast and tested. In the second stage, additional mixtures were cast

to study certain issues in greater detail.

To determine the mixes for the first-stage study, important CLSM

properties (and corresponding test methods) were prioritized and classified into

the following three groups, depending on the research objectives of this study:

I. Important CLSM properties


- Flow, setting time, unconfined compressive strength, and corrosion
measured for all 38 mixtures
II. Potentially important CLSM properties
- Segregation and bleeding, subsidence, triaxial shear, CBR, resilient
modulus, water permeability, drying shrinkage, excavatability, freezing
and thawing, leaching/environmental impact
- Measured for selected mixtures only
- Only seeking “Order of Magnitude” differences
III. Less important CLSM properties
- Direct shear strength, air/gas permeability, consolidation,
thermal conductivity
- Not included in laboratory study
- Literature-based and existing-practice-based coverage only

The most common types of CLSM materials and mixtures were selected

for the initial laboratory study. For each of these mixtures, the types and amounts

of cement, fly ash, and aggregates were selected prior to mixing, and the water

content of each mixture was then adjusted to achieve a flow of about 200 mm, as

measured by ASTM D 6103-97, “Standard Test Method for Flow Consistency of

Controlled Low-Strength Material (CLSM).” The three main groups of mixtures

and the range of materials chosen for the initial 38 mixtures included:

72
CLSM (with fine aggregates)
Type I portland cement: 1 type, 2 levels - 30 kg/m3 , 60 kg/m3
Fly ash: 3 types, 3 levels - 0 kg/m3 , 180 kg/m3 , 360 kg/m3
Fine aggregate: 2 types, 1 level - 1500 kg/m3
Air content: 3 levels - Entrapped air only, 15-20% air, 25-30% air
(Air-entraining agents will not be used for CLSM containing fly
ash)

CLSM (without fine aggregates)


Type I portland cement: 1 type, 1 level - 60 kg/m3
Fly ash: 3 types, 1 level - 1200 kg/m3
Air content: 1 level - Entrapped air only

CLSM (with set accelerator)


Selected mixtures from the test matrix

After selecting the rationale for the mixture proportions, as previously

defined, and specifying the desired dosage range for the key components, a

statistical program (ECHIP) was used to generate portions for a test matrix of the

non-air-entrained CLSM mixtures. This software uses experimental design

concepts to produce statistically significant results with a minimal number of

trials. In other words, rather than producing CLSM with every possible

combination of material and dosage, which would not be practical, an optimized

test matrix was produced that could be used to predict test results across the entire

spectrum of variables. In addition, the program can be used to statistically

compare the results of one test to another or to determine the effects of individual

or combined variables on test results. The program is also designed to assess the

repeatability of test results by requiring duplication of certain mixtures within the

test matrix. The repeatability of a test method greatly affects its acceptance as a

standard test method.

73
Although an air-entraining dosage 13 times the manufacture’s

recommended dosage was used in the CLSM mixtures containing foundry sand,

less than 3% of air was introduced in these mixtures. This may be attributed to

the presence of coals and organic materials in the foundry sand. Thus, foundry

sand was not included in air-entrained CLSM mixtures. For the air-entrained

mixtures, a full test matrix was used. Seven CLSM mixtures other than the above

mentioned mixtures were chosen for the stage 1 study. As such, a total of 38

mixtures were tested.

In stage 2, based on the results of stage 1, more specific CLSM properties

were studied in further detail. For each aspect studied, specially selected CLSM

mixtures were cast.

3.7.2 Mixture Proportions

To maximize the practicality of the research program, two separate

mixture series were generated using the statistical software, one for non-air-

entrained CLSM (with fly ash), and one for air-entrained CLSM (without fly ash).

The non-air-entrained mixtures are shown in Table 3.6. Mixture proportions for

air-entrained mixes are shown in Table 3.7. To maintain continuity in mixture

numbering, two new mixtures, mixture #20 and #21, were substituted for the

originally proposed foundry sand mixtures.

74
Table 3.6: Non-air entrained CLSM mixture proportions
Mixture* Type I cement Fly ash type Fly ash content Fine aggregate
content type**
(kg/m3 ) (kg/m3 )
1 30 Class C 180 River sand
2 60 Class C 180 River sand
1r 30 Class C 180 River sand
9 60 Class F 360 Foundry sand
15 30 Class C 360 Foundry sand
4 30 Class F 360 River sand
5 60 Class F 180 Bottom ash
3 60 Class C 360 Bottom ash
8 60 High carbon 180 Foundry sand
10 30 High carbon 180 Bottom ash
12 30 Class C 360 Bottom ash
6 30 High carbon 360 River sand
3r 60 Class C 360 Bottom ash
4r 30 Class F 360 River sand
7 30 Class F 180 Foundry sand
11 60 High carbon 360 Bottom ash
14 60 Class F 360 River sand
13 60 Class C 360 Foundry sand
5r 60 Class F 180 Bottom ash
2r 60 Class C 180 River sand
* ECHIP randomizes order of mixtures and provides for duplicates.
** Fine aggregate content was held constant at 1500 kg/m3.

Because of the modifications to the initially planned matrix for the air-

entrained mixtures, the statistical program for designing the air-entrained mixtures

was not used. Mixtures covering all of the selected variables were cast and

evaluated. That is, two cement contents (30 kg/m3 and 60 kg/m3 ), two target air

contents (15-20% and 25-30%), and two aggregate types (concrete sand and

bottom ash) were used in all combinations to create a total of eight mixtures.

From these eight mixtures, three were selected for replicate mixtures, bringing the

total number of mixtures in Table 3.7 from 7 to 11.

75
Table 3.7: Air-entrained CLSM mixture proportions
Mix # Cement content Air content Fine aggregate
(kg/m3 ) type**
23 60 25-30% Bottom Ash
17 30 15-20% Bottom Ash
19 30 25-30% Bottom Ash
18 60 15-20% Concrete Sand
16 30 15-20% Concrete Sand
22 60 25-30% Concrete Sand
20* 60 15-20% Bottom Ash
22 60 25-30% Concrete Sand
21* 30 25-30% Concrete Sand
16 30 15-20% Concrete Sand
20* 60 15-20% Bottom Ash
* These mixes were substituted for the originally proposed mixtures because of difficulty in
entraining air in mixtures containing foundry sand. The originally proposed mixtures
containing foundry sand were still cast, but without entrained air. Table 3.8 shows these
mixtures (Mix # 29 and #30).
** Fine aggregate content was held constant at 1500 kg/m3.

The mixtures shown in Table 3.8 were strategically chosen to investigate

specific mixture types that may be of interest. The mixtures represent typical

CLSM paste mixtures (i.e., 5% cement, 95% fly ash) and also include mixtures

containing an accelerating admixture. Lastly, this table includes non-air-entrained

CLSM mixtures containing foundry sand (selected after the challenges

encountered in entraining air in mixtures containing foundry sand).

Mixtures #24 and #25 were CLSM mixtures without fine aggregates. The

effect of acceleration admixtures was studied on mixtures #26, #27 and #28.

Because of the high fine content of foundry sand, it was possible to design CLSM

mixtures without fly ash or high air contents. Two mixtures, #29 and #30, were

proposed to for further investigation.

76
Table 3.8: Selected CLSM mixtures included in Stage One
Type I cement Fly ash
Mixture content Fly ash content Air content Fine aggregate
# (kg/m3 ) type (kg/m3 ) type*
24 60 Class F 1200 Entrapped air only None
25 60 High 1200 Entrapped air only None
carbon
26** 60 None 0 25-30% River sand
27** 60 Class F 1200 Entrapped air only None
28** 60 Class F 180 Entrapped air only River sand
29 60 None 0 Entrapped air only Foundry sand
30 30 None 0 Entrapped air only Foundry sand
* For mixtures 26, 28, 29, and 30, the fine aggregate content was held constant at 1500 kg/m3
** Mixtures contain accelerating admixture

After casting and testing the initially proposed mixtures (as summarized in

Table 3.6 through 3.8), additional mixtures were cast to further investigate or

refine selected test methods or to study selected CLSM properties in more detail.

The mixtures were based in most cases on previously cast mixtures (from the

original 38). However, there were several variations used to evaluate certain

properties or tests.

Because the compressive strength of CLSM is the most important property

commonly measured (and often the only hardened property measured), particular

emphasis was placed on developing a refined test that is more reliable and

reproducible. Issues such as load rate, curing condition, temperature effects, and

capping methods were studied in detail. The development of an improved

compressive strength test method is also critical because of the inclusion of

strength in most specifications, especially as it relates to excavatability.

77
Nine sets of additional mixtures were cast and will be referred to

throughout this dissertation by the designation of Mixture Series A through I, as

shown below.

Mixture Number of
Series Description Mixtures
A Effects of load rate on compressive strength (Table 3.9) 7
B Effects of curing and air-drying on compressive strength (Table 3.10) 2
C Long-term strength gain and excavatability (Table 3.11) 9
D Freezing and thawing resistance (Table 3.12) 11
E Alternative capping materials for compression cylinders (Table 3.13) 8
F Effects of drainage on compression cylinders (Table 3.14) 8
G Effects of storage conditions on compression cylinders (Table 3.15) 10
H Effects of temperature and humidity on compressive strength (Table 3.16) 6
I Permeability and triaxial shear strength (Table 3.17) 6

Tables 3.9 through 3.17 show the mixture proportions for the additional

investigations. To be consistent with the other mixtures in this study, the

aggregate content was held constant at 1500 kg/m3 . These tables also contain

selected information on the fresh CLSM properties. Later in Chapter 5, the

findings of these investigations are provided. These specialty studies were

designed to assist in refining and optimizing the test methods.

78
Table 3.9: CLSM mixture proportions for load-rate study
Mixture Cement Fly ash Fly ash Fine Water Flow Air Unit
3
content type content aggregate type (kg/m ) (mm) content weight
3 3
(kg/m ) (kg/m ) (%) (kg/m3 )
A-1 60 None None River sand 156 175 25.0 1739
A-2 60 Class F 360 Foundry sand 520 200 1.9 1755
A-3 60 Class F 1200 None 486 213 1.6 1620
A-4 30 Class C 180 River sand 265 330 1.7 2161
A-5 30 Class C 180 River sand 213 216 Not 2226
measured
A-6 60 Class F 1200 None 501 216 Not 1635
measured
A-7 60 None None River sand 156 165 24.5 1740

Table 3.10: CLSM mixtures for cylinder curing/specimen conditioning study


Mixture Cement content Fly ash Fly ash content Fine aggregate Water Flow
(kg/m3 ) Type (kg/m3 ) type (kg/m3 ) (mm)
B-1 30 Class C 180 Concrete Sand 203 250
B-2 30 Class C 180 Concrete Sand 189 200

Table 3.11: CLSM mixtures for excavation study


Mixture Cement River sand Fly ash Fly ash type Water Flow Air content Unit
(kg/m3 ) content content (kg/m3 ) (mm) (%) weight
3 3
(kg/ m ) (kg/m ) (kg/m3 )
C-1 60 None 1195 Class F 485 200 Entrapped 1637
C-2 0 2000 275 Class C 252 229 Entrapped 2148
C-3 30 1500 None None 112 178 28% 1642
C-4 15 1500 180 Class F 177 200 Entrapped 2192
C-5 30 1500 180 Class F 175 200 Entrapped 2158
C-6 15 1500 180 High-carbon 224 216 Entrapped 2095
C-7 30 1500 180 High-carbon 224 216 Entrapped 2115
C-8 15 1500 180 Class C 170 206 Entrapped 2190
C-9 45 1500 None None 103 178 25.5 1652

79
Table 3.12: CLSM mixtures used for freezing and thawing study
Mixture Cement Sand Fly ash Water Flow Air content Unit weight
(kg/m3 ) (1500 kg/m3 ) (180kg/m3 ) (kg/m3 ) (mm) (%) (kg/m3 )
D-1 30 River None 119 180 27% 1630
D-2 30 River Class F 205 200 Entrapped 2196
D-3 30 River High-carbon 256 229 Entrapped 2078
D-4 30 River Class C 200 216 Entrapped 1980
D-7 30 Foundry Class F 425 238 Entrapped 1835
D-6 30 Foundry High-carbon 481 229 Entrapped 1757
D-5 30 Foundry Class C 399 200 Entrapped 1800
D-8 30 Bottom ash Class F 357 200 Entrapped 1870
D-9 30 Bottom ash High-carbon 407 200 Entrapped 1733
D-10 30 Bottom ash Class C 282 200 Entrapped 1896
D-11 45 River None 96 152 30% 1569

Table 3.13: Mixture proportions for alternative capping materials study


Cement Fly ash Fly ash Concret Water Flow Total Air Unit
Mixture content type content e sand (kg/m (mm) bleeding content weight
(kg/m3) (kg/m3) (kg/m3) 3
) (%) (%) (kg/m3)
E-1 60 Class F 1140 None 480 200 2.30 1.4 1630
E-2 None Class C 206 1500 155 200 0.00 1.2 1626
E-3 30 None None 1500 109 175 0.00 29 1651
E-4 15 Class F 180 1500 184 200 2.07 1.7 1693
E-5 30 Class F 180 1500 176 200 0.87 2.5 2180
E-6 15 High- 180 1500 202 200 2.50 1.8 2122
carbon
E-7 30 High- 180 1500 229 225 3.04 - 2136
carbon
E-8 15 Class C 180 1500 109 200 0.85 1.75 2235

Table 3.14: CLSM mixture proportions for drainage-condition study


Cement Fly ash Fly ash River Water Flow Total Air Fresh unit
Mixture content type content sand (kg/m3) (mm) bleeding content weight
(kg/m3) (kg/m3 (kg/m3) (Weight %) (%) (kg/m3)
)
F-1 60 Class F 1140 None 480 200 2.30 1.4 1630
F-2 None Class C 180 2000 250 200 0.00 1.2 1626
F-3 30 None None 1500 109 175 0.00 29 1651
F-4 15 Class F 180 1500 184 200 2.07 1.7 1693
F-5 30 Class F 180 1500 176 200 0.87 2.5 2180
F-6 15 High- 180 1500 202 200 2.50 1.8 2122
carbon
F-7 30 High- 180 1500 229 225 3.04 Not 2136
carbon measure
d
F-8 15 Class C 180 1500 109 200 0.85 1.75 2235

80
Table 3.15: CLSM mixture proportions for cylinder storage-condition study
Mixture Cement Sand Fly ash Fly ash Aggregate Water Flow Air content
(kg/m3 ) (kg/m3 ) (kg/m3 ) type type (kg/m3 ) (mm) (%)
G-1 60 - 1140 Class F - 485 200 Entrapped
G-2 0 2000 275 Class C River sand 252 200 Entrapped
G-3 30 1500 - - River sand 112 187 29
G-4 15 1500 180 Class F River sand 177 200 Entrapped
G-5 30 1500 180 Class F River sand 175 200 Entrapped
G-6 45 1500 - - River sand 103 190 30
G-7 30 1500 180 Class F Foundry sand 349 216 Entrapped
G-8 30 1500 180 Class C Foundry sand 352 190 Entrapped
G-9 30 1500 180 Class F Bottom ash 424 140 Entrapped
G-10 30 1500 180 Class C Bottom ash 367 152 Entrapped

Table 3.16: Mixture proportions for effects of curing temperatures and


humidity
Cement Fly ash Fly ash Concrete Water Flow Air Unit
Mixture content type content sand (kg/m3 ) (mm) content weight
(kg/m3 ) (kg/m3 ) (kg/m3 ) (%) (kg/m3 )
H-1 60 Class F 1200 none 492 220 2.35 1631
H-2 15 Class F 240 1500 197 240 1.15 2191
H-3 15 Class C 240 1500 175 240 1.35 2212
H-4 30 Class C 180 1500 181 200 1.2 2163
H-5 30 Class F 180 1500 188 220 1.35 2210
H-6 60 none 0 1500 123 190 25.5 1603

Table 3.17: CLSM mixture proportions for triaxial shear strength and water
permeability
Mixture Cement Fly ash Sand Water Flow Air Fresh unit Dry unit Moisture
(kg/m3 ) (kg/m3 (kg/m3 (kg/m3 ) (mm) content weight weight content
) ) (%) (kg/m3 ) (kg/m3 ) (%)

I-1 30 180 1500 283 200 1.30 2036 1746 16.55


I-2 60 180 1535 327 210 1.25 2077 1748 18.79
I-3 120 180 1485 335 220 0.65 2087 1759 18.61
I-4 60 - 1471 212 190 22.5 1607 1415 13.59
I-5 60 - 1181 172 210 27.0 1569 1381 13.65
I-6 60 120 - 420 200 0.5 1529 1133 35.00

81
Chapter 4: Test Methods and Procedures

4.1 INTRODUCTION

The originally proposed testing program is summarized in Table 4.1.

Table 4.2 provides a summary of all the tests ultimately performed in this study,

which includes the originally proposed tests, in addition to supplemental studies

that were later added to the program (using the mixtures in Tables 3.9 to 3.17).

This section provides information on the specific test methods used in this study,

including various modifications to selected methods. Test methods are grouped

into three categories based on the properties that they are intended to measure:

fresh CLSM properties, hardened CLSM properties, and durability properties.

Some properties were studied in more detail than others. As previously

mentioned, significant emphasis was placed on the unconfined compressive

strength of CLSM.

82
Table 4.1: Summary of initial test program for this study
CLSM Test Specimen Test ages
property method(s) details
IMPORTANT PROPERTIES - TESTING FOR ALL MIXTURES
Flow ASTM D 6103 75 mm x 150 mm After mixing
Setting/Hardening Accordingly (function of early
ASTM C 403 150 mm x 150 mm
time stiffness)
Compressive 75 mm x 150 mm & 3 cylinders each @ 3, 7, 28,
ASTM D 4832
strength 150 mm x 300 mm and 91 days
CLSM -sand ASTM G 1 75 mm x 150 mm 182 days
comparative study
ASTM G 1 &
Galvanic cells 100 mm x 200 mm 182 days and monthly
ASTM G 109
50 mm diameter x 100
pH ASTM G 51 7, 28, and 182 days
mm height
100 mm x 150 mm x
Resistivity ASTM G 57 28 and 182 days
50 mm prism
POTENTIALLY IMPORTANT PROPERTIES - TESTING FOR LIMITED MIXTURES
Segregation and No standard, see test Every 15 min. for first hour,
150 mm x 150 mm
bleeding procedure then hourly
Subsidence No standard 610 mm x 100 mm After hardening & 48 hours
Triaxial shear USACE EM 1110- 73 mm diameter, 150 3 samples each at 7, 28, and
strength 2-1906 mm height 91 days
California Bearing 152 mm diameter, 178
AASHTO T 193 1 sample at 28 days
Ratio (CBR) mm height
73 mm diameter, 150
Resilient modulus AASHTO T 292 1 sample at 28 days
mm height
73 mm diameter, 150 2 samples each at 28 and 91
Water permeability ASTM D 5084
mm height days
87.5 mm × 26.3 mm × 1 sample each daily for 7 days
Drying shrinkage No standard
1000 mm angle & weekly
450 mm x 450mm x
Excavatability No standard 91 days (and later ages)
300 mm
Chloride diffusion ASTM C 1152 150 mm x 300 mm 30 and 91 days
Freezing and 102 mm diameter, 125
ASTM D 560 28 days
thawing mm height
EPA SW-846,
Leaching/environ-
Method 1311
mental impact
(TCLP)
LESS IMPORTANT PROPERTIES - NO TESTING
136 mm diameter, 20-
Direct shear strength ASTM D 3080 None
24 mm height
Thermal ASTM D 5334 76 mm x 203 mm None
conductivity
Air/gas permeability ASTM D 4525 73 mm x 150 mm None
Consolidation ASTM D 2435 63 mm x 200 mm None

83
Table 4.2: Overview of all properties and tests studied
Initial study (see Table
CLSM property Additional studies
4.1)
Flow ASTM D 6103
Soil pocket penetrometer
Setting/Hardening time ASTM C 403
Pocket torvane
Sample size effect
Small vs large machine
Loading rate
Effect of drying sample for 4
to 8 hours before testing
Compressive strength ASTM D 4832 Alternative capping materials
Effects of drainage
Effects of curing methods
Effects of curing temperature
and relative humidity
CLSM-sand comparative study ASTM G 1
ASTM G 1 & ASTM G
Galvanic cells
109
pH ASTM G 51
Resistivity ASTM G 57
Segregation and Bleeding ASTM C 940
Subsidence No Standard
ASACE EM 1110-2-
Triaxial shear strength
1906
California Bearing Ratio (CBR) AASHTO T 193
Resilient modulus AASHTO T 292
Water permeability ASTM D 5084
Drying shrinkage No standard
Excavatability No standard Splitting tensile strength
Chloride diffusion ASTM C 1152
Freezing and thawing ASTM D 560 Effects on permeability
Direct shear strength None
Thermal conductivity None
Air/gas permeability None
Consolidation None
Chemical and toxicity
Leaching/environmental impact None analyses of constituent
materials

84
4.2 FRESH CLSM TEST METHODS

4.2.1 Mixing Procedure and Measurement of Flow, Air Content, and Unit
Weight

For each of the 38 initial mixtures studied trial mixing was performed to

determine the approximate water demand needed for the target 200 to 250 mm

flow. To measure flow in trial mixtures, as well as actual test mixtures, ASTM D

6103-97 was followed. This method uses a 75-mm x 150-mm plastic cylinder,

with the bottom removed. The cylinder is raised, and the flow diameter is

measured (sometimes referred to as the “pancake” diameter).

After determining the target water content, the actual mixtures were cast

and test samples were prepared. For all of the 38 mixtures, the constituent

materials were batched using a Sartoris 38-kg capacity scale (0.5-gram accuracy).

For the smaller mix volumes, a 0.028-m3 drum mixer was used. For the larger

mix volumes (needed for measuring additional properties on the six selected

mixtures), a high capacity (0.056-0.070-m3 ) laboratory mixer was used. The

mixing procedures were different for air-entrained and non-air-entrained

mixtures, as discussed below.

For non-air-entrained mixtures, the dry materials (e.g., sand, fly ash, and

cement) were first mixed with approximately half of the expected mixing water

(based on trial mixing) for three minutes, followed by a two-minute rest period.

After the rest period, the remainder of the batched water was added, followed by

three or more additional minutes of mixing. Immediately after mixing, flow

measurements were taken. In most cases, as a result of the trial mixing, the target

flow of 200 to 250 mm was obtained. If the flow was less than desired, small
85
amounts of water were added, followed by an additional minute of mixing, to

obtain the target flow. Rarely was the flow greater than the desired range. Care

was taken to avoid excessive bleeding or segregation.

For air-entrained CLSM mixtures, a relatively dry consistency (i.e., zero

slump) was first obtained in the mixer, followed by the addition of the AEA. This

process was necessary because of the potency of the AEA. If the AEA was added

to an already fluid mixture, the flow would far exceed the desired range, and the

mixture would often suffer from excessive bleeding.

Immediately after mixing, the flow was measured according to ASTM D

6103, and the air content was measured using slightly modified ASTM C 231

(pressure method). If the flow and air content met the desired ranges, test samples

were then cast. If the air content was found to be lower than desired, additional

AEA was added (within the specified dosage of the admixture). It can often be

challenging to obtain both the desired flow and air content. When CLSM is

relatively dry, adding AEA will increase both air content and flow. Conversely,

adding water to a mixture that already has the desired air content but has a lower

than desired flow may actually decrease the air content (the flow may actually

increase or decrease, depending on the effects on air content). The experience

gained in trial mixing allowed converging on the target air content and flow in the

test mixtures much more efficiently. After obtaining the desired flow and air

content, the unit weight was measured using ASTM C 231. Upon successfully

meeting the target air content and flow, various test samples were cast and tested.

86
4.2.2 Setting and Hardening

The setting time of CLSM was measured using two methods, needle

penetration (ASTM C 403) and soil penetrometer (or “pocket” penetrometer). A

typical soil penetrometer is shown in Figure 4.1. Fresh CLSM was placed in two

standard containers (150 mm x 150 mm), one container for each test method.

Before each measurement, the bleed water was removed and weighed. The

cumulative weight of the bleed water removed throughout the test was used to

calculate the sample bleeding (bleed water mass divided by total sample mass).

The air temperature during testing was maintained at 21 °C.

It should be noted that the depth of penetration for the needle penetrometer

was approximately 25 mm, compared with only 6.4 mm for the soil penetrometer.

The differences in penetration depth, as well as contact area, may have a

significant impact on the measured values of penetration resistance. In particular,

the relatively high bleeding of CLSM (compared to concrete) may lead to

difficulties in early measurements, especially when using the smaller needle

penetrometer. Modifications were made to the needle penetrometer to minimize

the influence of bleed water and obtain a more accurate estimate of early

stiffness/strength gain.

87
Figure 4.1: A typical soil pocket penetrometer

To estimate the early-age shear strength of CLSM, a pocket vane shear

(VS) tester, as shown in Figure 4.2, was used. This device is often used by

geotechnical engineers to estimate the undrained shear strength of cohesive soils

in the field. It was utilized in this study to assess the setting/hardening of fresh

CLSM mixtures. A typical pocket VS tester is shown in Figure 4.2

88
Figure 4.2: Pocket vane shear tester used in this program

4.2.3 Segregation

The segregation of six CLSM mixtures was measured quantitatively. A

specially designed mold, which consisted of three separate cylindrical sections,

was used for this purpose. Each cylindrical section had a diameter of 100 mm and

a height of 75 mm. The sections were connected together vertically to produce a

sample cylinder with a diameter of 100 mm and a height of approximately 225

mm. After the samples had set, steel separating plates were inserted at the

junctions between the cylinder sections, thus yielding three separate samples

(upper, middle, and lower). Each sample was then wet sieved, using the #4, #8,

#16, #30, #50, #100 and #200 sieves. Each portion retained on these sieves was

89
then dried in the oven at 110 ºC for 24 hours and weighed. Material passing the

#200 sieve was not collected. Using the resultant gradation from each of the three

sections, a “pseudo” FM was calculated, using the same mathematical approach

as typically used for FM. As is the case with the normal treatment of the FM, two

different gradations can yield the same FM. As such, the overall gradation was

also considered when analyzing the results, as described later.

4.2.4 Subsidence

The term “subsidence” used here is defined as the settlement of the top

surface detected within several hours of CLSM mixture placement. To measure

subsidence, three 100-mm x 200-mm cylindrical plastic molds were vertically

aligned and firmly bonded together using high-strength tape. The bottoms of the

upper two molds were removed, allowing for the casting of a 600 mm high CLSM

specimen for subsidence measurement. Extra tall specimens were cast to ensure

that sufficient subsidence would occur and that the reduction in the column height

could be more readily measured. The choice of specimen height was rather

arbitrary. Because the maximum pressure increases with the height of fresh

CLSM mixture and higher pressure will result in denser material, it is anticipated

that the measured subsidence (in percentage) will increase with specimen height.

A 600-mm height was chosen to represent a typical layer thickness. After casting

the specimen, the top edge of the specimen was leveled off and periodic

measurements were made of the subsided surface, using a specially designed

probe, which was lowered from the top of the mold onto the center of the

90
subsided surface. Bleed water usually evaporated before the first measurement.

Measurements were conducted until subsidence ceased, usually within two hours

of casting the specimens. Because the friction from the mold wall restrains the

settlement of CLSM, the results from laboratory tests may under estimate the real

settlement in the field.

4.3 HARDENED PROPERTIES OF CLSM

4.3.1 Unconfined Compressive Strength

Because of the importance of the compressive strength of CLSM in

specifications, design, and construction, considerable emphasis was placed on

developing a test method with improved accuracy and reliability. This section

describes the basic procedures followed to test the unconfined compressive

strength of the initial 38 mixtures, including methods of preparing test cylinders,

curing, capping, and testing. After describing this approach, information is

provided on the various modifications and improvements investigated using the

mixtures previously shown in Tables 3.9 through 3.16. The results of the initial

compression study, as well as the findings of the various follow-up studies, can be

used to develop and recommend an improved unconfined compression test for

CLSM.

4.3.1.1 Specimen Preparation

Because of the relatively low strength of CLSM (compared to concrete), it

is very important to carefully handle test cylinders, especially when stripping the

91
cylinders from the molds. Although steel split molds are a good option for

casting CLSM specimens, because of the large number of specimens cast in this

program, a lower cost method was used. Before casting CLSM cylinders for

compression testing, the plastic cylinder molds were cut lengthwise in half till the

bottom. Electrical tape was then used to bind the mold back to its original shape.

This approach only has negligible impact on the shape and size of CLSM

specimens. Figure 4.3 shows a typical plastic mold in this study. After mixing

the CLSM, the cylinders were filled and tapped lightly on the sides to remove

large entrapped air voids. Plastic lids were then placed firmly on the cylinders,

and the specimens were moved immediately to a moist-curing or “fog” room,

which was maintained at 100% relative humidity (RH) and 23 ºC. After curing in

the molds for seven days, the cylinders were then stripped by simply removing the

electrical tape and removing the CLSM specimens from the cylinders.

Conventional stripping tools were not used because of possible damage to the

specimens. The cylinders were then kept outside of their molds in the fog room

until testing. Some CLSM mixtures tend to leach and soften slightly upon long-

term fog-room exposure. An investigation on this issue is described later in this

chapter, in which the effects of cylinder storage were studied in detail.

Moist curing was selected for this study so that test results can be

compared from one laboratory to another, even though CLSM is rarely, if ever,

moist-cured in the field. The same argument can also be made for concrete

testing. That is, concrete is rarely moist cured for more than seven days (if at all)

92
in field applications, but standard curing in a fog room provides a benchmark for

specification and construction acceptance.

Figure 4.3: Plastic mold were pre-cut and taped before casting specimens

4.3.1.2 Compression Machine

Although there is an existing ASTM method for measuring the unconfined

compressive strength of CLSM (ASTM D 4832), some modifications were made

to the method for this project, as described later. Most of the compression tests

were performed on a relatively low-load capacity and machine with displacement

control (100 kN Instron machine), but some testing was also performed on a

larger capacity machine with load control (1780 kN Tinius Olson machine) to

evaluate the effects of machine capacity. When using the smaller machine,

displacement control was used. Additional testing was also performed to examine
93
the effects of cylinder size (75 x 100 mm, 100 x 200 mm, and 150 x 300 mm),

and, for this study, a constant apparent strain rate, 2,493 microstrain/min, was

used. For this, the crosshead displacement was set at 0.38 mm/min for the 150-

mm-high specimen, 0.51 mm/min for the 200-mm-high specimen, and 0.76

mm/min for the 300-mm-high specimen. The objective was to produce failure in

about the same amount of time for each cylinder size for a given mixture. A

floating, spherical head was used to minimize eccentricities while loading. The

load-displacement curves were automatically logged to data files on a desktop

computer. A typical compression machine suitable for the testing is shown in

Figure 4.4.

Figure 4.4: A typical set-up of compression testing using geotechnical


unconfined compression machine.

94
For the larger capacity compression machine, load-controlled testing was

employed, as is the case for concrete testing. The typical load rates used for

concrete, 138 to 345 kPa/sec, would fail most CLSM specimens in a matter of

seconds. Thus, a lower load rate of 6.9 kPa/sec was selected. This lower load

rate was possible on the machine for this study but may not be available for many

standard concrete compression machines.

Only the peak load was obtained from the large capacity machine, and no

deformation data was obtained, which again is typical of concrete testing. Sulfur

capping was used for almost all of the cylinders, except for some weaker mixtures

at early ages, where it was not possible to use sulfur caps. For these mixtures,

neoprene pads were used for testing at seven days. As previously mentioned,

several variations were investigated for the unconfined compression test,

including cylinder size, machine capacity, capping method, load rate, and curing.

The results of these studies are described later.

4.3.1.3 Effects of Loading Rate on Compressive Strength

ASTM D 4832-95 provides little guidance regarding load rate. The

method states only to “Apply the load at a constant rate such that the cylinder will

fail in not less than 2 min.” Because of the vagueness in defining the load rate,

additional testing was performed to investigate the effects of loading rate on

compressive strength.

Table 3.9 shows the mixture proportions used for the load rate study.

These mixtures (a total of seven) were selected based on previous testing of the 38

95
mixtures in this project and were chosen to provide a wide range of materials and

proportions. Using these seven mixtures, the effects of displacement rate (cross-

head displacement of small load-frame) on compressive strength and deformation

at peak load were studied.

From each of the mixtures in Table 3.9 (A1-A7), standard cylinders (75

mm x 150 mm) were cast and moist-cured until the age of testing. The effects of

loading rate were first assessed for mixtures A-1 to A-4 with a wide range of load

rates (0.13 mm/min, 0.38 mm/min, and 0.89 mm/min). After performing these

tests, a narrower range of load rates (0.25 mm/min, 0.38 mm/min, 0.51 mm/min,

and 0.64 mm/min) were evaluated using mixtures A5-A7. Later in this chapter,

the terms “wide range” and “narrow range” refer to the aforementioned

investigations of load rates. The aim was to determine a suitable load rate range

that produces repeatable compressive strength values and can be performed in a

relatively short time. The latter concern was due to the fact that several mixtures

from early research took a relatively long time (i.e., greater than 10 to 15 minutes)

to fail in compression under displacement control, which would not be ideal for a

testing laboratory that must test many cylinders daily.

4.3.1.4 Cylinder Curing and Conditioning

Another possible source of error and confusion in ASTM D 4832 involves

the curing conditions and the treatment of cylinders before testing. According to

ASTM D 4832, curing of CLSM cylinders is performed with the specimens in the

molds (in the fog room) until the time of testing. This is different than the

96
normal concrete approach to stripping the cylinders from their molds after about

the first day of curing and then curing them in the fog room. ASTM D 4832 also

specifies a drying time of 4 to 8 hours for test cylinders after the moist-curing

period and before they are tested in compression. Concrete cylinders, on the other

hand, are specified to remain moist until the time of testing, with no required

drying time. Research was conducted to investigate the effects of cylinder storage

(i.e., in or out of molds) and specimen conditioning or drying prior to testing.

The mixtures shown in Table 3.10 were used to study the effects of drying

time (0.5, 2, 4, and 8 hours) and cylinder curing (in mold, out of mold, and in

lime-saturated water) and on the compressive strength of CLSM. The curing

regimes studied included:

• Leaving CLSM in the molds and curing in the fog room at 23 o C


• Stripping the cylinders at 3 days and curing in the fog room at 23 o C.
• Stripping the cylinders at 7 days and curing in lime-saturated water at 23 o C

After curing according to the above regimes, the test cylinders were then

capped with sulfur-capping compound and loaded at a standard loading rate of

0.38 mm/min. The cylinders were kept moist until the time of testing, i.e., no

specimen drying was allowed.

Because of the interesting outcome of these tests, as described later,

additional mixtures were subsequently cast to study these issues in more detail.

To study variations in curing regimes, the mixtures in Table 3.15 were cast and

test cylinders were prepared. This study aimed to identify possible differences in

compressive strength when four different curing conditions were utilized, as

97
summarized in Table 4.3. Note that for curing condition D, the cylinders were

placed outside the laboratory and were exposed to the high summer temperature

and dry atmosphere of Austin, Texas. All cylinders were capped with sulfur

capping compound and tested at a loading rate of 0.38 mm/min.

Table 4.3: Four curing conditions used in the study (using G-series mixtures)
Curing Curing regime
condition
A (normal) Keep sample in mold with cap on, for seven days in fog room. Then
strip cylinder and keep cylinders in fog room until time of testing
B (mold) Keep sample in mold, with cap on, for seven days in fog room. Then
remove cap and keep cylinder in mold in fog room until time of testing
C (cap) Keep sample in mold, with cap on, in the fog room until time of testing
D (out) Keep sample in mold, with cap off, outdoors until time of testing

4.3.1.5 Effects of Curing Temperature and Humidity on Compressive Strength

As already addressed, the temperature that CLSM is exposed to during its

strength-gain process may be very important, especially when mixtures containing

certain fly ashes are used. Because CLSM is used in many different environments

in practice, the same mixture proportions could exhibit different strength values.

This study is intended to identify factors affecting strength gain of CLSM

mixtures. This study was a follow-up to earlier testing that suggested that

temperature plays a major role in CLSM strength development.

Three curing temperatures (10 ºC, 21 ºC, and 38 ºC) and six CLSM

mixtures (H-1 to H-6 in Table 3.16) were selected to study the strength gain of

CLSM across a range of practical construction conditions. CLSM was cast into

standard cylinder molds (75 mm x 150 mm) and moved to the appropriate

98
temperature-controlled chamber until the date of testing. The cylinders were

stored in two different manners. Half the cylinders from each mixture were

stripped after three days and returned to the same chamber until the time of testing

(without control over relative humidity in the chamber). This condition is

designated later in this report as “dry” curing. Temperature and humidity were

monitored throughout the test. The other half of the specimens from a given

mixture were kept inside the molds with the caps firmly placed on top until the

day of testing (designated as “wet” curing). These cylinders were placed directly

next to the cylinders that had already been stripped.

Cylinders were tested for compressive strength at 7, 28, and 91 days.

After compression testing, the moisture contents of the specimens were measured

to assess the effects of curing conditions on the moisture content (or evaporable

water content) and strength of CLSM.

4.3.1.6 Effects of Drainage Conditions on Compressive Strength

Unlike conventional concrete, CLSM is rarely, if ever cured in the field. It

is used as a geotechnical material and placed without concern for curing. During

the strength-gaining process of CLSM, much of the CLSM is continuously

exposed to air and in contact with the surrounding soil and/or structures.

Different environments may significantly affect the final strength of CLSM as the

water cement ratio may be affected by the seepage of water into surrounding

materials or the loss of water through evaporation of bleed water.

99
The effects of seepage and evaporation were investigated in a study using

the mixtures detailed in Table 3.14 (F-1 to F-8). This study also investigated the

effects of temperature on strength gain, using the fog room as a control and

ambient conditions (hot Texas summer weather) as a test condition. As described

later in this report, the findings of this temperature effects study were quite

interesting, and subsequent testing was performed using controlled-temperature

environments to further elucidate the influence of temperature on CLSM strength,

especially for mixtures containing high volumes of fly ash.

To simulate the field conditions, CLSM mixtures were cast in plastic

molds buried in loose sand. Figure 4.5 shows the test set-up. To simulate the

condition of no water loss, CLSM mixtures were cast in plastic molds without

holes and tight lids were put on (condition “Cap”) immediately after casting. To

simulate conditions that only surface water evaporation is possible, mixtures were

cast in plastic molds without holes and lids (condition “No cap”). To simulate

moderate water seepage, mixtures were cast into molds with seven uniformly

distributed 3.6 mm diameter holes on the bottom (condition “Bottom holes”). No

lid was put on the cylinders. To simulate severe drainage condition, mixtures

were cast into molds with holes not only on the bottom but also on the side

(condition “Side holes”). There were 36 holes on the wall and seven holes on the

bottom per mold. All holes were of 3.6 mm diameter. No lid was put on the

cylinders. To reduce local variations, CLSM specimens from a given mixture

were randomly placed throughout the test box

100
Loose
Mold sand

450 mm 300 mm

Compacted
450 mm sand

450 mm

Figure 4.5: Drainage condition simulation

4.3.1.7 Alternative Capping Materials for Compression Testing

In preliminary laboratory trials and throughout the whole laboratory

program, sulfur capping was found to be an effective method of obtaining

repeatable compressive strength data. However, for early age samples and/or for

particularly low strength cylinders, it may not be possible to cap cylinders with

sulfur because of the risk of damage to the sample. In these instances, neoprene

pads were used. In addition, some limited comparative studies were performed

early in the laboratory program in which sulfur capping was compared to

neoprene pads. However, because only limited testing was performed, it was

decided to significantly expand the scope of the original work to investigate not

only sulfur and neoprene, but also high-strength gypsum capping compound, for a

range of CLSM mixtures. Other motivations for studying alternatives to sulfur

101
are the potential health concerns over the fumes generated from sulfur capping

stations and the length of time needed to cap cylinders with sulfur.

The use of neoprene pads has gained popularity in recent years for

personnel testing concrete. In addition to avoiding the fumes associated with

sulfur, neoprene pads are easier and faster to use than having to cap with sulfur. It

is well known that higher strength concrete requires higher neoprene durometer

values, and vice-versa. Thus, for CLSM, it is expected that softer neoprene pads

(much softer than those used for concrete) may be needed. Another possible

alternative to sulfur capping compound is high strength gypsum, which is

sometimes used for concrete. Because of the high flowability of the gypsum

paste, it could be helpful in capping CLSM cylinders that are prone to suffering

edge damage (e.g., weak cylinders). The set-up of gypsum capping is shown in

Figure 4.6. In addition, gypsum is non-hazardous, without the odor and health

concerns associated with sulfur.

To address the capping-related issues, a comprehensive investigation of

alternative capping materials was launched. Included in this study were sulfur

caps, gypsum (or hydrostone) caps, and neoprene pads with durometer values of

20, 40, 50, 60, and 70. Eight mixtures (E-series) were used in this study, as

previously described in Table 3.13.

102
Figure 4.6: Method of capping cylinders with gypsum compound

4.3.2 Triaxial Shear Stre ngth

The conventional soil triaxial test (USACE EM 1110-2-1906) was

followed for testing six CLSM mixtures, as I-series shown in Table 3.17. The

samples were cast in Shelby tubes of approximately 70 mm diameter and were

stripped after seven days. The testing was performed under consolidated and

drained conditions. The pore water pressure was maintained at 34.5 kPa, and the

confining pressures were 69.0 kPa, 103.5 kPa, and 172.5 kPa, respectively.

Theses values resulted in effective confining stress of 34.5, 69.0, and 138.0 kPa,

respectively. The loading rate was 0.38 mm/min, the same loading rate used for

most of the unconfined compression tests. The tests were terminated when the

residual strength was reached or the stress-strain curve became essentially flat.

The approach of stress paths was used to interpret the data. By curve fitting, the

effective internal friction angle, φ´, and effective cohesion intercept, c´, were

found.

103
4.3.3 Water Permeability of CLSM

The water permeability of six CLSM mixtures (I-series) was measured

using ASTM D 5084 -90, “Standard Test Method for Measurement of Hydraulic

Conductivity of Saturated Porous Materials Using a Flexible Wall Permeameter”.

The cylinders were cured in the fog room for 28 days. A back pressure of 69 kPa

was applied and maintained until no additional water entered the sample

(approximately 30 minutes). This was assumed to represent a saturated condition.

Because the samples were moist-cured prior to testing, the samples were already

essentially saturated prior to sample conditioning. CLSM specimens were

relatively incompressible, thus, the requirement that B ≥ 0.95 was lifted. Rather,

the change of the “B” value with respect to the change of confining pressure was

monitored similar to the testing of rock. Constant-head test was used.

Eleven mixtures, designated as D-series in Table 3.12, were utilized to

study the freezing and thawing effects on permeability. The CLSM cylinders

were first moisture cured for 28 days and then were exposed to 12 freezing-and-

thawing cycles. Because CLSM specimens without air entrainment usually break

down in the testing process, polyvinyl hear-shrink plastic film was used to hold

the specimens together. The specimens were 100 mm × 125 mm. Porous stones

were wrapped on both ends of the specimens. Figure 4.7 shows the specimen set-

up. The falling-head method was used for these mixtures.

104
Figure 4.7: Specimen set-up for testing of permeability of CLSM mixture
after freezing-and-thawing cycles

4.3.4 Drying Shrinkage of CLSM Mixtures

No standard test methods exist for the measurement of the drying

shrinkage of CLSM. A method commonly used (but not a standard method) in

Germany for self-leveling floor screeds was modified and used in this study.

CLSM was cast into an 87.5 mm × 26.3 mm × 1000 mm steel “C” channel. This

angle had one fixed end plate with an anchor and one movable end plate with an

anchor. Prior to casting, wax paper was placed on the inside of the angle to

reduce friction. CLSM was then placed in the fabricated forms. The amount of

shrinkage was measured by an linear voltage displacement transducer (LVDT) of

105
an accuracy of 0.01 mm, which measured the displacement of the movable end

plate. Shrinkage measurements were taken daily for the first week and once a

week thereafter.

4.3.5 California Bearing Ratio

No significant modifications were made to AASHTO T 193-99, “Standard

Method of Test for The California Bearing Ratio,” for testing CLSM in this study.

The only difference was that the material (CLSM) was cast or poured into the

molds, without the need for compaction, as is needed for testing soils. After

seven days of curing, the collar was removed, and the surface was trimmed level

using a straight edge. The specimens were tested at the age of 28 days. Six

selected CLSM mixtures were tested.

4.3.6 Resilient Modulus

The resilient modulus of six selected CLSM mixtures was determined in

accordance with AASHTO T 292–91, “Standard Method of Test for Resilient

Modulus of Subgrade Soils and Untreated Base/Subbase Materials”. In trial

testing, it was found that the deviator stresses listed in Table 4 of AASHTO T

292-91 were not high enough to introduce sufficient deformations. Thus, the

selection of deviator stresses was based on experience performed at Texas A&M

University. Load conditioning of 41 kPa was used for the 1000 repetitions.

106
4.3.7 Excavation Study of CLSM Mixtures

The excavatability of CLSM was assessed for six of the original 38 CLSM

mixtures included in this study to gain an “order of magnitude feel” for the

relative ease of excavating various CLSM mixtures. CLSM was cast into 450-

mm x 450-mm x 300-mm plywood boxes and allowed to harden. Attempts were

made to correlate “walkability” with soil penetrometer values as the CLSM

gained strength in the first few hours. Walkability is the time at which an average

person can walk on the material. Long-term excavatability was assessed at an age

of approximately nine months using typical hand tools, including a shovel and a

pick for six selected mixtures. The compressive strength of lab-cured cylinders

was also measured. In addition, a relatively new instrument, the Humboldt

GeoGauge® (stiffness gauge), was used at the time of excavation to attempt to

correlate excavatability with the stiffness of CLSM, as measured by the

GeoGauge®.

After the initial excavation study, a more comprehensive study on long-

term strength gain and excavatability was conducted. Nine CLSM mixtures (C-1

to C-9 in Table 3.11) were included in the study. A field penetrometer (field

version of ASTM C 403) was used to evaluate the strength gain of CLSM

mixtures in the boxes. The dynamic cone penetrometer (DCP) was also utilized

to estimate the excavatability of CLSM mixtures. The DCP is a modified and

simplified version of the penetrometer used by the Country Roads Board,

Victoria, Australia. It is used by geotechnical engineers to obtain an index of in-

situ CBR and to estimate the strength of soil as a function of depth. The testing

107
consists of dropping a hammer (8 kg in mass) from a height of 575 mm, which

forces a steel rod with a conical head into the CLSM or soil. The penetration

depth per blow is then recorded. The corresponding DCP index value was used to

estimate a soil strength value. This value is often correlated with CBR.

Another approach to estimate the excavatibility of CLSM is the

calculation of its removability modulus. In its CLSM specification, Hamilton

County, Ohio, defines a Removability Modulus (RE), as reproduced below in

Equation 1.a.

W 1 .5 × 104 × C 0 .5 (4.1)
RE =
106
Where, W – In situ unit weight (pcf)

C – Thirty-day unconfined compressive strength (psi)

When SI units are used, as required by AASHTO, the equation is rewritten

as Equation 1.b below:

W 1. 5 × 0.619 × C 0 .5 (4.2)
RE =
106
Where, W – In situ unit weight (kg/m3 )

C – 30-day unconfined compressive strength (kPa)

Engineers in Hamilton County, Ohio and the city of Cincinnati have found

this methodology to be effective in limiting long-term strength gain and ensuring

future excavatability. The same approach was used to calculate RE values for the

C-series mixtures and comparing the results to other direct or indirect indices of

CLSM strength gain. If a CLSM mixture demonstrates a RE value higher than or

equal to 1.0, it is deemed not excavatible in the future.

108
In preliminary investigations, it was found that the splitting tensile

strength of CLSM to be a very simple property to measure (without the need to

cap the cylinders), and the test results may be helpful in predicting excavatability.

In addition, the stress conditions of CLSM specimens under splitting tensile

testing may be quite similar to the stress conditions associated with digging the

material with a shovel or backhoe. Eight mixtures (E-series mixtures shown in

Table 3.13) were tested for tensile strength, and comparisons were made with

other CLSM properties, such as compressive strength.

4.4 DURABILITY OF CLSM

4.4.1 Freezing-and-Thawing Resistance

ASTM D 560, “Standard Test Methods for Freezing and Thawing

Compacted Soil-Cement Mixtures,” a method designed to measure the freeze-

thaw resistance of soil-cement mixtures, was used with one minor modification.

The modification was that the application of wire scratch brush on thawed

specimens was not required because of the low strength of CLSM. One freeze-

thaw cycle involved cycling between -18 °C (a freezer) and 23 °C (the fog room).

Samples were exposed to 12 cycles unless they suffered severe damage at an

earlier time. Mass loss was used as the index of damage. For the initial study,

three cylinders (100 mm x 125 mm) were moist-cured for seven days and three

other cylinders were moist-cured for 28 days prior to exposure to freeze-thaw

cycles. To study the effects of freeze-thaw damage on permeability the

specimens were moist-cured for 28 days prior to freezing-thawing exposure. It

109
should be noted that tests typically used for concrete, such as ASTM C 666, were

found to be too severe in preliminary trials, and the modified soil cement method

was found to be a more suitable approach.

4.4.2 Wetting-and-Drying Resistance of CLSM

A preliminary study on the wetting-and-drying resistance of CLSM

mixtures was performed. After 7 and 28 days of moist curing, the wetting-and

drying cycles were initiated. The specimens were soaked in water at 23 ºC for 24

hours and then dried for 24 hours at room temperature (relative humidity

approximately 50%). The mass of CLSM specimens was recorded before and

after soaking. The testing was stopped after 12 cycles. Six selected mixtures (#4,

#6, #23, #24, #22r, and #26) were investigated.

4.5 CORROSION OF DUCTILE IRON IN CLSM

4.5.1 Research Approach

Corrosion of underground utilities can cause premature failures and can

result in the need for costly repairs. Concerns have been expressed regarding the

corrosion performance of ductile iron pipe in soil and CLSM. Limited work has

been performed on corrosion of ductile iron pipe embedded in CLSM and work is

needed to either substantiate or dismiss these concerns. In general, several factors

can influence the corrosion performance of utilities embedded in both soils and

CLSM. These factors can include84 :

• the presence of aggressive chemicals or ions

110
• Acidity or alkalinity (pH of the solution at the pipe interface)

• electrical conductivity or resistivity of the soil or CLSM

• porosity (aeration) or drainage conditions near the utility

• moisture

As pitting depths depend on sizes of test specimens, it is difficult, if not

impossible, to simulate field performance of iron pipelines. Thus, in the testing

program, uniform corrosion will be assumed such that mass loss can be measured.

Because CLSM is quite different from native soils from an electrochemical

perspective, it is possible that a galvanic cell is formed by a pipe contacting

CLSM and soil at different parts of one cross-section or at different sections along

the pipe. For the first scenario, when a small portion of pipe contacts native soil

(forming an anode of a small area) and the large part contacts CLSM (forming a

cathode of a large area), the worst case of galvanic corrosion is set up. Mass loss

testing was performed in both uncoupled and coupled conditions to assess the

inherent corrodibility of metals placed solely in CLSM and to determine the

impact of galvanic cells on the corrosion process.

4.5.2 Mass Loss Test

4.5.2.1 Uncoupled Condition

To evaluate the actual corrosion performance of ductile iron pipe (and

other metallic materials) embedded in CLSM, samples can be embedded in

CLSM and evaluated for mass loss in laboratory conditions. Analyses of the

111
corrosion activity of coupons embedded in CLSM mixtures can be compared with

the corrosion activity of coupons embedded in a standard ASTM Ottawa sand.

For comparative testing, Ottawa sand was selected as a control material due to its

minimal variance in gradation and composition. However, any soil type could be

used in such a test. The experimental test set-up is shown in Figure 4.8. All of

the initial 38 CLSM mixtures were evaluated in this study. Specimens were cast

and moist cured for 28 days, then immersed in a 3.5% sodium chloride (NaCl)

solution to accelerate corrosion. Mass loss measurements were taken after 18

months of exposure. Preparing, cleaning, and weighing of specimens were

performed in accordance with ASTM G 1-90, “Standard Practice for Preparing,

Cleaning, and Evaluating Corrosion Test Specimens.” Mass loss provides a

measurement of average corrosion rate during the exposure period and thus can be

used as a parameter for evaluating the corrosion susceptibility of coupons

embedded in different materials and exposed to different environments.

75 x 150mm
cylinder

3.5% NaCl
solution
Pipe
coupon 100 mm

CLSM
or sand

Figure 4.8: Corrosion test set-up for comparing corrosion performance of


coupons in CLSM and sand (uncoupled).

112
4.5.2.2 Coupled Condition (Galvanic)

Figure 4.9 shows the laboratory test set-up for evaluating samples that

may experience galvanic coupling. Samples were moist cured for 28 days before

being exposed to a 3.5% NaCl solution and being electrically connected by a 10-

Ω resistor. Two coupons with essentially equal areas were used in this set-up.

All of the 38 original CLSM mixtures were evaluated in this study. Preparing,

cleaning, and weighing of specimens was performed in accordance with ASTM G

1–90.

10 Ω
resistor

100 x 200 mm
cylinder
3.5% NaCl solution

Coupons CLSM Sand 100 mm

Figure 4.9: Corrosion test set-up for comparing corrosion performance of


galvanically coupled coupons in CLSM and sand.

4.5.3 pH of Solutions Extruded from CLSM

CLSM sample solutions were extruded from 50 mm × 100 mm cylinder

specimens. The pH values were measured with a pocket pH-meter according to

113
ASTM G 51-95, “Standard Test Method for Measuring pH of Soil for Corrosion

Testing.” The pH values were recorded when the data were stabilized, usually

two or three minutes after inserting the meter into the solution.

4.5.4 The Resistivity of CLSM

This test followed the California DOT Test 532, “Method for Estimating

the Time to Corrosion of Reinforced Concrete Structures.” The set-up for

measuring the resistivity of CLSM is shown in Figure 4.10. The set-up is

composed of plastic plates for the frame and steel plates as electrodes for the

current. The box size measures 100 mm ×150 mm × 50 mm. The resistance

values of 38 CLSM mixtures were measured at an age of 182 days.

Steel Screws
Plates

Figure 4.10: Set-up for resistance measurement

114
4.5.5 Chloride Diffusion

For underground iron pipes, the rate of corrosion can be accelerated by the

presence of aggressive ions. In predicting the susceptibility of pipes placed in

engineered backfill materials, it may be beneficial to establish the quantity of

aggressive ions at the pipe location. In concrete applications, chloride contents as

a function of depth can be determined using specific ion probes to determine how

chlorides migrate through CLSM.

Five CLSM mixtures (#4, #6, #22r, #24, and #26) were evaluated for

chloride ingress rate, as shown in Figure 4.11. From each mixture, two CLSM

cylinders (150 mm x 300 mm) were ponded for 22 weeks with a 3.5% NaCl

solution. Samples were then tested for chloride contents as a function of cylinder

depth.

Chloride
Solution

150 x 300 mm
CLSM cylinder

Figure 4.11: Set-up for chloride ion diffusion

115
4.6 LEACHING AND ENVIRONMENTAL IMPACT

Recent concerns raised by the Environmental Protection Agency (EPA)

regarding the use of coal combustion products in construction have rekindled the

interest in studying the environmental implications of using CLSM containing

coal combustion products and other waste materials. Although the EPA did not

ban or curtail the use of coal combustion products, as originally expected, the

topic will likely resurface in the future. Because the use of by-product and waste

materials in CLSM is so common, some research was performed to assess the

potential toxicity of materials used in this study and to develop a protocol for

assessing materials being considered for use in construction.

The laboratory testing included a full chemical analysis of the by-product

and recycled materials included in this project, including fly ash (class F, Class C,

and high-carbon), bottom ash, and foundry sand. The chemical analysis was

conducted using inductively coupled plasma emission spectroscopy (ICP), gas

chromatography-mass spectrometry (GC-MS), total organic carbon (TOC), and

atomic absorption. Information such as the calcium oxide (CaO) content of fly

ashes can be very helpful in assessing the potential reactivity and effects on long-

term strength gain.

The assessment of the toxicity of the by-product materials used in this

study focused on the identification and quantification of heavy metals. Generally,

heavy metals are more of a concern than organics in fly ash, so the limited

laboratory study only tested for heavy metals. Various techniques are available

116
for the measurement of organics in constituent materials and CLSM mixtures, but

they were not included in this study.

For each of the by-products included in this study (three fly ashes, one

bottom ash, and one foundry sand), the total heavy metal concentration was first

determined in accordance with EPA Method 610, where nitric acid and hydrogen

peroxide were used to digest the materials. The eight elements analyzed included

arsenic, barium, cadmium, chromium, lead, mercury, selenium, and silver.

Because this testing determines the total amount of heavy metals, and not the

leachable amount, the extraction values are often assumed to be 20 times that of

the TCLP limits. If any of the by-product materials yielded values in excess of

the toxicity limits (20 times the TCLP limit), TCLP was then conducted to assess

the type and amount of heavy metals that are actually leachable from the

materials. Method 40CFR 261.24 was used to extract the samples for TCLP

testing, and the same eight heavy metals as previously listed were measured. The

concentrations of the extracts were compared to the TCLP limits which follow:

Element TCLP Limits (ppm)


Arsenic 5.0
Barium 100.0
Cadmium 1.0
Chromium 5.0
Lead 5.0
Mercury 0.2
Selenium 1.0
Silver 5.0

Testing the constituent materials directly, rather than testing the materials

encapsulated in CLSM, represents the worst-case scenario for the purpose of

117
toxicity screening. If the constituent materials test below the TCLP toxicity limits

under this worst-case scenario, the materials can be classified as non-toxic, and no

additional tests need to be performed. None of the materials tested in this study

exceeded the TCLP limits, and thus tests on actual CLSM specimens were not

deemed necessary. However, if any of the constituent materials had exceeded the

TCLP limits when tested by themselves, CLSM mixtures would have been cast

for subsequent leaching tests. The overall approach to the toxicity and leaching

tests performed in this study is shown in Figure 4.12.

118
STEP 1:
Chemical analysis of all raw materials

STEP 2:
Heavy metal test of all raw materials
using EPA 610

Yes
Less than
TCLP limits ?

No

STEP 3:
TCLP test (Method 40CFR 261.24)
on potentially toxic raw materials
from Step 2

Yes Raw materials deemed non-toxic.


Less than Accepted to be used in CLSM
TCLP limits ? mixtures.

No

STEP 4:
TCLP test (Method 40 CFR 261.24)
on CLSM mixture containing
potentially toxic raw materials from
Step 3

Yes
Less than
TCLP limits ?

No

Raw materials deemed toxic.


Rejected to be used in CLSM
mixtures.

Figure 4.12: Flow chart to study toxicity (heavy metals) of CLSM


constituent materials

119
Chapter 5: Results and Discussions

5.1 INTRODUCTION

In this section, results from the testing program described in Chapter 4 are

provided. Specifically, the results of tests on the fresh and hardened properties,

durability characteristics, and environmental leaching aspects of CLSM mixtures

are presented. Detailed discussions on the test results are included, along with

some discussions on related mechanisms. Focus is placed on the applicability of

test methods for field applications and on better understanding factors affecting the

short- and long-term behavior of CLSM, with specific focus on factors influencing

strength.

5.2 CLSM MIXTURE PROPORTIONS AND FRESH PROPERTIES

5.2.1 Water Demand/Flow Behavior

An important aspect of this study was the assessment of the fresh or plastic

properties of CLSM, both in terms of evaluating candidate test methods and

determining the relationship between constituent materials, mixture proportions,

and fresh CLSM properties. Table 5.1 summarizes some of the important

parameters, including water demand (to obtain the target flow), air content, flow,

unit weight, and bleeding (%) for the initial 38 mixtures. Note that mixtures ending

in “r”, such as mixture 2r, denote mixtures that were repeated for statistical

120
Table 5.1: CLSM mixture proportions and fresh properties

Mixture Cement Fly ash type Fly ash Fine aggregate Water flow Total Air Fresh unit
content content type demand bleeding content weight
(kg/m3) (kg/m3) (kg/m3) (cm) (%) (%) (kg/m3)
1 30 Class C 180 Concrete sand 211 20 NA 0.9 1965
2 60 Class C 180 Concrete sand 206 20 2.45 0.95 2108
1r 30 Class C 180 Concrete sand 206 21 2.08 0.9 1974
15 30 Class C 360 Foundry sand 486 20 0.13 2.75 1741
3 60 Class C 360 Bottom ash 577 18 4.32 1.65 1754
8 60 High carbon 180 Foundry sand 532 24 1.04 3.3 1647
10 30 High carbon 180 Bottom ash 628 14 4.81 2.0 1681
9 60 Class F 360 Foundry sand 520 20 0.54 2.5 1684
5 60 Class F 180 Bottom ash 600 18 5.84 2.5 1739
12 30 Class C 360 Bottom ash 572 22 3.64 2.7 1774
4 30 Class F 360 Concrete sand 220 20 0.39 2.2 2199
7 30 Class F 180 Foundry sand 501 20 0.57 2.1 1817
3r 60 Class C 360 Bottom ash 541 20 2.58 2.1 1997
4r 30 Class F 360 Concrete sand 220 22 2.92 1.8 2211
24 60 Class F 1200 None 486 24 2.25 2.8 1635
23 60 None 0 Bottom ash 454 14 1.30 28.5 1382
18 60 None 0 Concrete sand 200 22 0.70 16.5 1836
14 60 Class F 360 Concrete sand 216 22 1.00 1.3 2174
2r 60 Class C 180 Concrete sand 206 25 0.21 0.5 2291
29 60 Foundry sand 0 None 373 23 0.28 2.6 1812
30 30 Foundry sand 0 None 414 20 0.40 2.0 1789
17 30 None 0 Bottom ash 582 13 4.35 20.0 1447
11 60 High carbon 360 Bottom ash 573 23 6.42 1.7 1743
6 30 High carbon 360 Concrete sand 315 20 2.26 1.3 2103
16 30 None 0 Concrete sand 295 20 2.33 16.0 1922
21 30 None 0 Concrete sand 170 18 0.62 25.5 1789
22 60 None 0 Concrete sand 131 20 0.05 26.5 1748
22r 60 None 0 Concrete sand 136 18 0.43 25.5 1802
5r 60 Class F 180 Bottom ash 600 16 7.20 1.4 1887
26 60 None 0 Concrete sand 136 17 0 25.5 1802
16r 30 None 0 Concrete sand 295 19 2.35 15.5 1874
13 60 Class C 360 Foundry sand 499 20 0 1.8 1902
25 60 High carbon 1200 None 853 24 7.38 1.3 1322
19 30 None 0 Bottom ash 492 13 1.08 25.0 1385
20 60 None 0 Bottom ash 525 13 3.41 18.5 1485
27 60 Class F 1200 None 486 23 1.28 0.7 1638
20r 60 None 0 Bottom ash 525 13 1.44 15.5 1511
28 60 Class F 180 Concrete sand 220 20 1.33 1.4 2182

121
purposes. The results of the mixtures are presented in the order in which they were

mixed and are not sorted by mixture type. The fresh properties of the A- to I-series

mixtures can be found in Table 3.9 through 3.17.

The water demand to obtain the target flow was found to be one of the most

important parameters affecting CLSM. Certain constituent materials, such as high

carbon fly ash, bottom ash, and foundry sand, were found to particularly increase

the water demand of CLSM. By increasing the water content of CLSM, most of

the important properties, such as unconfined compressive strength, were impacted.

ASTM D 6103 was found to be a suitable method for measuring the flow of

CLSM. For mixtures with low density (high air content), it was found that the 200

mm flow was sometimes difficult to obtain, and for these mixtures, a flow of 175

mm was found to be satisfactory. The bottom ash used in this study lacked

sufficient fines for good workability and was prone to bleeding. However, when

used with high contents of fly ash, the mixtures were more workable and exhibited

reduced bleeding.

The effects of material source and quantity on the water demand of CLSM

were analyzed using a statistical program (ECHIP) for the non-air-entrained

mixtures. Figure 5.1 shows the results. This graph illustrates the statistically

significant variables that affect water demand. In this graph, the effect is the

difference between the specified variable level and the mean water demand of

446.3 kg/m3 . The reference variable levels are concrete (river) sand, Class C fly

ash, 180 kg/m3 fly ash content, and 30 kg/m3 cement. The effect of reference level

is the opposite of the sum of remaining levels. The plus sign (+) in the key after the

122
variable description indicates that the difference was positive and dash (-) indicates

that the difference was negative. The negative sign means that the variable level

will reduce that amount of water from the mean. The effect of a reference level is

the opposite of the combination effects of all other variable levels. From the figure

it can be seen that the fine aggregate source was the most significant factor

affecting the water demand of mixtures. The use high-carbon fly ash also increased

the water demand. There is no significant difference between the use of the Class

C and Class F fly ash. The effect of high content of fly ash (360 kg/m3 ) was

negligible.

10

Bottom ash (+)


Foundry sand (+)
High carbon (+)
60kg cement*high carbon (-)
60kg cement*class F (-)
60kg cement (-)
Class F (-)
60kg cement*360kg fly ash (-)
360kg fly ash

0
0 20 40 60 80 100 120 140 160
Effect on water demand (kg/m3 )

Figure 5.1: Effects of variables on water demand of non-air-entrained


CLSM mixtures

123
Table 5.2 shows the results from the analysis of variance (ANOVA)

calculations on water demand. ANOVA is a statistical method for comparing two

or more populations or treatment. The Mean Squares column shows the summary

statistic for the source and the degree of freedom (DF) column shows the number

of model terms summarized by the mean square. The P column gives the

significance. Values less than 0.05 are usually considered statistically significant

and thus affecting the water demand. Thus, this table identifies the significant

variables as fly ash type, fine aggregate source, and the interactions between

cement content and fly ash type.

Table 5.2: Analysis of variance (ANOVA) for water demand

Source of Variation Mean Square Degree of freedom (DF) Significance, P


Cement content 894.919 1 0.0595
Fly ash type 3986.39 2 0.0003
Fly ash content 1.6386 1 0.9293
Sand type 214854 2 0
Cement content * fly ash type 2082.62 2 0.0035
Cement content * fly ash content 681.909 1 0.0932
Error 198.063 10
Replicate error 132.1 5

5.2.2 Air Contents of Fresh CLSM Mixtures

The air contents of CLSM mixtures were measured by the approach

presented in Chapter 4. This approach was found to be easy and fast. The results

are shown in Table 5.1. The range of entrapped air contents was similar to those of

conventional concrete. It was extremely difficult to entrain air in mixtures

containing foundry sand. This is most likely attributed to the presence of clay, coal

fines and other organic materials in the foundry sand. It is believed that the

124
foundry sand prevented the encapsulation or formation of entrapped tiny air voids,

thus making the AEA none-effective. Air-entraining agent dosages were increased

up to 1300% of the manufacturer’s recommended dosage, but it was still not

possible to entrain considerable air.

5.2.3 Unit Weight of Fresh CLSM Mixtures

The results of unit weights of fresh CLSM mixtures are shown in Table

5.1. The unit weights of CLSM mixtures in this study ranged from 1322 to 2291

kg/m3 . The unit weights of CLSM mixtures with entrapped air were only slightly

lower than ordinary Portland cement concrete because of the higher water contents

in CLSM mixtures. Mixtures without fine aggregates only exhibited lower

densities as a result of the higher water demand and the lower specific gravity of fly

ash.

5.2.4 Setting/Hardening

The setting/hardening behavior of CLSM is important for many

applications, especially those where early strengths are required to satisfy

construction demands (i.e., timing between lifts or early opening to traffic). Test

methods are needed to more easily assess the setting of CLSM, both in the

laboratory and in the field. This section discusses some of the preliminary findings

regarding the setting time of CLSM using the following methods:

• Needle penetrometer

125
• Soil pocket penetrometer

• Pocket vane shear tester

Because CLSM exhibits substantially lower strength than normal concrete

mixtures, it is not possible to assign initial and final CLSM setting times as is

typically associated with concrete. In this study, however, penetration resistance

(ASTM C 403) was correlated with soil penetrometer values for all mixtures, and

hardening time related to “walkability time” was investigated for select mixtures.

Appendix B shows the raw data from the soil penetrometer (SP) and needle

penetrometer (NP) tests for the Stage One 38 mixtures, as well as the eight

mixtures from the I-series. The walkability time was assessed by preparing large

“CLSM boxes” that were walked on at various ages. It was found that the soil

penetrometer values were in the range of 4.3 kPa to 7.4 kPa (average of 6.1 kPa)

when CLSM mixtures were able to support the weight of an average person with an

approximate 6 mm indentation.

When measuring the needle penetration of CLSM, it should be noted that a

certain minimum strength of CLSM is required to obtain meaningful test results.

Thus, it is often not feasible to compare the setting time of CLSM mixtures to each

other at predefined time increments, but rather, the timing of measurements was a

function of constituent materials and mixture proportions. Also, because a needle

penetrometer penetrates deeper into mixtures than a soil penetrometer, it is less

subject to bleed water effects. Despite these differences (25 mm versus 6.4 mm) in

penetration depth, there was a limited correlation between soil penetrometer and

126
needle penetrometer values for the 38 mixtures, as shown in Figure 5.2 with a R2

(the square of the correlation coefficient) of approximately 0.75 for all the CLSM

penetration data combined.

0.8

0.7
0.6 R2 =0.75

0.5
0.4
0.3

0.2
0.1
0
0 2 4 6 8 10 12 14

ASTM C 403 penetration (MPa)

Figure 5.2: Correlation between ASTM C 403 and soil penetrometer values

One potential problem with using the soil pocket penetrometer is the effects

of bleeding on the measurements. The accumulation of significant bleed water

(and fines) is common in CLSM construction and can make surface measurements

or assessments challenging. It is difficulty to locate the “true” top surface of fresh

CLSM and being able to quantify the penetration resistance of the top layer of

CLSM. To address this potential problem, a stand was designed, as shown in

Figure 5.3, to minimize the effects of bleeding on field measurements. When using
127
the pokcet penetrometer, the stand was first placed on the mixture surface. The top

of the penetrometer was then inserted into the hole on the top the stand, where its

dead weight pushed the tip of the penetrometer through the soft fines and bleed

water and came to a rest on the dense surface that marked the “true” top. The 6.4

mm penetration depth was then marked on the tip, based on the measurement taken

from the top surface of the stand. Finally, the needle was pushed down and the

reading was recorded. This modified stand was used for the E-series mixtures.

Figure 5.3: Stand for the pocket soil penetrometer

The soil pocket penetrometer gives an estimation of the unconfined

compressive strength of cohesive soil by penetration resistance, while the needle

penetrometer gives the bearing pressure. This test is based on Terzaghi’s bearing

capacity equation for circular footings, and the soil friction angle is assumed to be

zero. Thus, the bearing capacity/resistance is only attributed to internal cohesion.

The unconfined compressive strength is two times the cohesion, according to the

theoretical total stress failure envelope. For the pocket soil penetrometer, the

penetration is required to be 6.4 mm and the soil underneath is assumed to have

failed at this depth. According to Bhat, Bucchi suggested this method for the rapid

determination of soil strength components.10 Using Terzaghi’s equation, which

assumes that only cohesion contributes to the bearing strength, the following

formula can be derived:

128
1.3 × 5.7q u
qf = 1.3 Nc Cu = = 3.71 qu (5. 1)
2

Where qf = ultimate bearing capacity

Cu= undrained cohesion

qu = unconfined compressive strength

The pocket penetrometer is made to give readings directly in qu. By

multiplying the reading from the soil penetrometer by 3.71, the penetration

resistance (qf) can be back calculated. This is used in Figure 5.4 to compare the

penetration resistance measured with the pocket penetrometer and needle

penetration tests.

129
3.00
SP

BSP
2.50
Series3
back calculated soil penetrometer resistance

a
Soil penetrometer reading (SP) and

2.00
(BSP) (MPa)

1.50

1.00

0.50

0.00

0.00 0.50 1.00 1.50 2.00 2.50 3.00

Needle penetration (MPa)

Figure 5.4: Comparison between needle penetration (NP) and soil


penetrometer reading (SP) values and the back calculated
resistance from soil penetrometer (BSP).

From Figure 5.4, it was observed that at the early stage of setting, the soil

penetrometer may yield higher resistance values than those measured by needle

penetration. This may be because the soil penetrometer is more sensitive at low

values than the needle penetrometer that measures resistance up to 890 N. As

hydration proceeds and CLSM gains stiffness, the needle penetrometer clearly

exhibited higher values than those of the soil penetrometer. When using the needle

penetrometer for a penetration depth of 25.4 mm, it was observed that that the load

per unit area increased to a peak and then dropped off considerably. This

phenomenon was not observed for the soil penetrometer, which only requires a

130
penetration depth of 6.4 mm. Because CLSM mixtures are often plastic for several

hours after mixing, the soil penetrometer may not generate a bearing failure.

Figure 5.5 shows a needle (ASTM C 403) after the penetration testing.

Note the small wedge at the tip of the needle. This confirmed that a general failure

has been reached and is in accordance with Terzaghi’s assumption on ultimate

bearing capacity. Cracks caused by the needle penetration were observed on

specimen surfaces. Clearly, the penetration of 25.4 mm caused large plastic

deformations, thus affecting the bearing capacity. This also offers a possible

explanation as to why penetration values sometimes decreases, even though the

mixture is still gaining strength.

The pocket VS tester was used to test the shear strength of CLSM mixtures

at early ages. There was some limited correlation between the VS values and the

soil penetrometer values, as shown in Figure 5.6 (R2 =0.45). The VS values

represent an estimate of cohesion and unconfined compressive strength based on

some assumptions on failure plane, whereas soil penetrometer values are measured

indirectly.

131
Wedge

Figure 5.5: Setting time needle and the wedge at the tip of the needle

(following ASTM C 403).

132
0.50

0.40
R2 = 0.45
0.30

0.20

0.10

0.00
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40

Soil penetrometer (MPa)

Figure 5.6: Comparison between soil penetrometer and vane shear values

5.2.5 Bleeding of Fresh CLSM Mixtures

Bleeding was found to be common for CLSM mixtures. Even for CLSM

mixtures with high air content, a very small amount of bleeding water was

observed on the mixture surface. Flash fill (mixture C-2) showed little bleeding

because of its fast setting nature. The bottom ash used in this study lacked

sufficient fines for good workability and was prone to bleeding. However, when

used with high contents of fly ash, the mixtures were workable with reduced

bleeding.

Factors affecting the bleeding of CLSM mixtures with moderate air

contents were also analyzed using the statistical software program, ECHIP. The

results from the analysis are shown in Figure 5.7. The mean bleeding is 2.52%.

The foundry sand used in this study reduced the bleeding significantly, while the

133
bottom ash significantly increased the bleeding. The fly ash type had a minimal

effect on the bleeding of CLSM mixtures.

10
Bottom ash (+)

Foundry sand (-)


XXXXXX High carbon FA

XXXX 60 kg cement * high carbon FA

XXXX 360 kg fly ash

XXX Class F FA

XX 60 kg cement

XX 60 kg cement * Class F
XX 60 kg cement * 360 kg FA

0
0 1 2 3 4
Bleeding (% of weight)

Figure 5.7: Effects of variables on bleeding of non-air-entrained CLSM


mixtures. “+” means increasing the bleeding, “-“ means
decreasing, and “X” means the effect is not significant compared
with variations.

5.2.6 Segregation of Fresh CLSM Mixtures

Using the method for assessing segregation described earlier in Chapter 4,

the results from the test shown in Table 5.3 were obtained. These data represent

the gradation of aggregates sieved from different depths within a CLSM cylinder.

A “pseudo-fineness modulus” was calculated using the same approach used to

calculate the FM for sand used in concrete. Based on the data shown in Table 5.3,
134
very little segregation occurred in any of the mixtures. However, the test method

developed to assess segregation showed promise as a useful tool to quantify

segregation. This method will be applied in the proposed field-testing program,

with the intention of correlating the laboratory results with actual field

performance.

Table 5.3: Segregation data for selected CLSM mixtures


Cumulative Cumulative Cumulative Cumulative Cumulative Cumulative
retained retained retained retained retained retained
(%) (%) (%) (%) (%) (%)
Sieve # Mix # 4 Mix # 23
4 4.55 3.30 6.83 10.82 21.10 16.63
8 23.70 24.01 27.60 33.41 41.05 35.13
16 46.97 45.68 50.08 52.88 54.88 50.98
30 70.29 67.40 71.41 66.83 68.26 63.89
50 86.90 82.68 87.69 79.49 82.27 78.57
100 95.94 92.11 96.09 92.31 92.55 91.00
200 100 100 100 100 100 100
Pseudo-
Fineness 3.28 3.15 3.40 3.36 3.6 3.36
modulus
Sieve # Mix # 6 Mix # 22r
4 5.32 3.43 3.70 4.44 3.47 3.53
8 14.72 23.27 24.26 24.38 25.94 24.75
16 38.70 45.23 43.94 45.56 45.93 44.76
30 60.30 63.81 65.90 65.34 68.39 66.08
50 83.78 84.69 85.22 88.03 88.59 89.72
100 95.30 95.97 95.31 98.09 98.04 98.29
200 100.00 100.00 100.00 100 100 100
Pseudo-
Fineness 2.98 3.16 3.18 3.26 3.30 3.27
modulus
Sieve # Mix # 26
4 3.75 5.44 3.74
8 27.06 29.06 25.87
16 47.03 47.67 46.72
30 67.41 72.09 69.52
50 88.79 90.12 89.65
100 98.23 98.07 98.26
200 100 100 100
Pseudo-
Fineness 3.32 3.42 3.34
modulus
*Note: Because mixture 24 did not contain aggregates, segregation was not measured

135
5.2.7 Subsidence

The results of subsidence testing for six CLSM mixtures are shown in Table

5.4. All of the CLSM mixtures exhibited appreciable subsidence, with the

exception of mixture #23, which had relatively poor flowability. With the

exception of mixture #23, there was a limited correlation between subsidence and

bleeding, as shown in Figure 5.8. The use of an accelerating admixture reduced

bleeding and slightly reduced subsidence, as shown by comparing mixtures #26

and #22r (identical mixtures but mixture #26 contained an accelerator).

Table 5.4: Subsidence results of six selected CLSM mixtures

Mix No. 4 Mix No. 24


Time (hrs) Subsidence (mm) Time (hrs) Subsidence (mm)
1.25 2.95 3 5.3
2.3 3.43 5.5 6.8
3.57 3.41 6 6.8
4.57 3.5
5.17 3.41
Mix No. 23 Mix No. 6
Time (hrs) Subsidence (mm) Time (hrs) Subsidence (mm)
3 0.3 1 9.4
4 0.3 2.17 10.5
3.33 12.65
4.33 14.85
6.5 15.85
Mix No. 22r Mix No. 26
Time (hrs) Subsidence (mm) Time (hrs) Subsidence (mm)
3.17 1.22 3.67 1.5
5.17 2.15 5.67 1.5
6.33 2.15
Note: The total specimen height was 600 mm

136
3

2.5

1.5

0.5

0
0 0.5 1 1.5 2 2.5

Bleeding ( weight %)

Figure 5.8: Correlation between bleeding (weight percent) and subsidence

5.3 HARDENED PROPERTIES OF CLSM

5.3.1 Unconfined Compressive Strength

5.3.1.1 Compressive Strength Development

A great deal of emphasis was placed in Stage One on assessing the

unconfined compressive strength of CLSM. This section first summarizes the

findings from the initial mixtures (38 in all), in which several aspects of

compression testing were examined, including the effects of cylinder size, capping

material, and load rate. Based on these original mixtures, some useful predictive

models were developed to predict the strength gain (short- and long-term) of

CLSM. The original study led to several follow-up studies, each of which focused

137
in more detail on issues involving testing parameters. Detailed investigations on

load rate, curing and conditioning of cylinders, effects of drainage on strength, and

the use of alternative capping materials are included in following sections. The

findings of the initial, broad study and the later, detailed studies could be used to

refine and improve existing methods of measuring the unconfined compressive

strength of CLSM.

The unconfined compressive strengths of the originally proposed mixtures

at 3, 7, 28, and 91 days are shown in Table 5.5. The results were found to be

repeatable, with quite low values of coefficient of variation. The data shown in this

table were for 75 mm × 150 mm small, sulfur-capped cylinders tested on a smaller

capacity machine, as previously described.

An interesting observation that illustrates the uniqueness of CLSM is that

most mixtures show a drastic change in the load-deflection curve as the curing time

is increased. Figure 5.9 illustrates this behavior for Mixture #10, which was typical

of most CLSM mixtures. At early ages, CLSM acts more like a soil, with more

ductile behavior. But as time progresses, CLSM begins to act more like concrete,

with higher strength and lower ductility.

As stated earlier, various modifications to the unconfined compression test

were studied in the initial investigation, some of which were later addressed in

more detailed research. Table 5.6 shows a comparison between sulfur-capped and

neoprene-capped cylinders for the selected six CLSM mixtures. In general, sulfur-

capped cylinders yielded higher strengths than cylinders using neoprene pads. The

138
Table 5.5: Unconfined compressive strength of original 38 CLSM mixtures
3-day f’c C.O.V. 7-day f’c C.O.V. 28-day C.O.V. 91-day C.O.V.
Mixture
(MPa) (%) (MPa) (%) f’c (MPa) (%) f’c (MPa) (%)
1 0.12 8.17 0.21 6.81 1.09 4.91 1.87 2.8
2 0.29 2.42 1.76 10.11 3.69 4.03 6.26 13.5
1r 0.13 14.36 0.24 1.18 1.35 7.69 2.34 0.3
15 0.07 8.22 0.11 5.66 0.18 3.97 0.25 1.4
3 0.33 10.50 0.57 2.42 1.36 6.83 2.02 2.9
8 0.09 9.75 0.11 8.27 0.25 4.69 0.33 10.6
10 0.12 2.27 0.16 15.12 0.22 7.93 0.26 1.3
9 0.09 9.66 0.13 12.40 0.22 4.99 0.25 2.9
5 0.14 13.70 0.18 8.15 0.46 16.11 0.57 5.3
12 0.30 16.17 0.27 6.16 0.57 4.70 0.86 3.4
4 0.34 4.94 0.48 6.18 0.79 11.04 1.08 13.6
7 0.09 3.22 0.11 3.40 0.12 9.70 0.16 5.8
3r 0.46 13.08 0.58 4.36 1.49 5.78 1.97 8.3
4r 0.41 13.61 0.57 5.80 0.94 3.92 1.03 6.9
24 0.34 4.77 0.22 1.40 0.44 0.13 0.58 4.6
23* - - 0.04 6.42 0.14 9.52 0.18 7.6
18* - - 0.33 6.75 0.70 1.08 0.79 4.0
14 0.58 6.61 1.07 13.46 2.15 8.14 3.49 16.8
2r 0.42 9.30 1.58 2.86 4.90 2.44 6.87 0.3
29 0.18 0.57 0.31 0.22 0.63 1.9 0.98 6.1
30 0.09 7.08 0.14 3.10 0.26 8.55 0.28 2.8
17* - - 0.01 31.79 0.07 18.90 0.13 16.8
11 0.33 1.66 0.42 4.29 0.75 3.48 0.94 4.7
6 0.40 10.20 0.47 0.74 0.83 4.88 1.09 4.7
16* - - 0.06 11.90 0.13 12.03 0.16 8.5
21* - - 0.09 10.59 0.16 11.84 0.18 11.7
22* - - 0.43 8.97 0.73 4.33 1.01 4.8
22r 0.32 4.19 0.50 9.68 0.96 17.53 0.93 7.0
5r 0.17 12.50 0.28 10.33 0.55 10.08 0.78 16.3
26 0.43 7.08 0.76 8.23 1.14 15.75 1.53 2.6
16r* - - 0.07 9.74 0.15 23.57 0.17 8.3
13 0.28 0.45 0.35 3.33 0.74 3.2 1.12 4.4
25 0.17 4.44 0.30 5.72 0.40 30.90 0.50 9.0
19* - - 0.02 0.71 0.06 45.02 0.06 13.2
20* - - 0.04 36.42 0.21 1.01 0.29 26.0
27 0.22 4.61 0.29 1.72 0.36 3.83 0.55 6.6
20r* - - 0.04 49.90 0.15 32.77 0.24 10.4
28 0.28 3.02 0.47 0.86 0.70 1.95 0.94 0.2
* Mixtures were too weak to be tested at three days.

139
3000
28 days

2500

2000
7 days
Load (N)

1500

1000
3 days
500

0
0 1 2 3 4 5

Deformation (mm)

Figure 5.9: Load-deformation response of Mixture #12 at 3, 7, and 28 days

Table 5.6: Comparison of capping materials and methods


Mixture 3 days 7 days 28 days
Sulfur Neoprene Sulfur Neoprene Sulfur Neoprene
(C.O.V.) (C.O.V.) (C.O.V.) (C.O.V.) (C.O.V.) (C.O.V.)
MPa (%) MPa (%) MPa (%) MPa (%) MPa (%) MPa (%)
4 0.34 (4.9) 0.25(15.1) 0.48 (6.2) 0.34 (6.0) 0.79(11.0) 0.55(14.4)
6 0.45 (6.8) 0.25(14.7) 0.47 (0.7) 0.36 (6.3) 0.81 (4.9) 0.61 (3.2)
23 - - 0.04 (6.42) 0.03 (12.92) 0.14 (9.52) 0.12 (3.14)
24 0.34 (4.8) 0.15 (0.3) 0.22 (1.4) 0.19 (20.5) 0.44 (0.1) 0.30(18.8)
22r 0.32 (4.2) 0.22 (8.1) 0.50 (9.7) 0.37 (10.6) 0.93(17.5) 0.54 (4.4)
26 0.43 (7.1) 0.29 (9.1) 0.76 (8.2) 0.42 (11.2) 1.11(15.8) 0.73 (7.2)

neoprene pads used in this study had a durometer value of 50 (an index of

hardness), which is a typical durometer value for conventional concrete cylinder

testing. A more comprehensive study on capping materials, including neoprene

140
pads with varying durometer values, was subsequently performed, as described

later in this chapter.

Efforts were made in this study to develop predictive models for the

compressive strength of CLSM. Various models and statistical approaches were

considered. No single model was found to work well for the entire range of

materials and mixture proportions; however, predictive models for subsets of the

mixtures were found to be quite accurate. For instance, separate models were

developed for air-entrained (both for “high” and “moderate” air contents), and non-

air-entrained CLSM.

In developing the predictive model for air-entrained CLSM (up to 91 days),

the mixtures containing bottom ash as the fine aggregate, because of the

aforementioned problems with bleeding, were not included in the model. By trial-

and-error, the formula predicting the compressive strength of air-entrained CLSM

mixtures (not including bottom ash mixtures) is shown below:

f c' = a ⋅ e b ( w / c ) ( 5 .2 )
a = 0 . 3074 ⋅ ln( t ) + 0 . 22
b = 0 .0086 ⋅ ln( t ) − 0 . 272

Where, f c' is the compressive strength in MPa, w/c is the water-cement ratio, and t

is the age in days.

The measured and predicted compressive strengths for air-entrained

mixtures are plotted in Figure 5.10. There was very good correlation, with a R2

value of 0.97. This formula was also found to be effective in predicting long-term

141
strength gain (i.e., beyond 91 days). For example, cylinders from mixture 22r were

tested for compressive strength after 256 days, with an average strength of 1.0

MPa, compared to the predicted value of 1.1 MPa.

1.4

1.2
R2 =0.97
1
Predicted strength (MPa)

0.8

0.6

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4
Measured strength (MPa)

Figure 5. 10: Comparison of measured and predicted strengths of air-


entrained CLSM mixtures at the ages of 7, 28, and 91 days.

A predictive formula was also developed for non-air-entrained CLSM

mixtures. An initial approach, similar to that used by Bhat10 , was used in which the

water-cement ratio was the only variable used to predict the compressive strength.

This approach yielded an R2 value of 0.8. To improve this accuracy, a formula was

developed that included the water-cement ratio, aggregate type, fly ash type, and

the fly ash content as strength-predicting variables. A critical aspect to this

approach was to assign numerical values to the non-numerical variables used in the

142
formula. Through a trial and error process, the following constants (k) were

selected for the materials used in this investigation. Concrete or river sand (kriver

sand) was assigned a value of 1.0, foundry sand (kfoundry sand) a value of 0.2, bottom

ash (kbottom ash ) a value of 1.0, Class C fly ash (kC ash ) a value of 2.2, Class F fly ash

(kF ash ) a value of 1.0, and high-carbon fly ash (kHC ash ) a value of 0.75. The

equation for predicting the compressive strength, S(t), is shown below, and a

comparison between predicted and actual compressive strengths is shown in Figure

5.11.

S(t) = b 0 (t) ⋅ (k agg.type ) b 1 (t) ⋅ (k fly ash type ) b 2 (t) ⋅ (w/c) b 3 (t) ⋅ (k fly ash content ) b 4 (t) (5. 2)
where,

b0 (t)= -0.00071•t2 + 0.131•t -0.755

b1 (t)= -0.000118•t2 +0.0134•t+0.417

b2 (t)= -0.0000842• t2 +0.0154• t+0.0941

b3 (t)= -0.0034•t-1.03

b4 (t)= 0.745-0.0178•t when t 30 days

b4 (t)=0.217 when t > 30 days

Where, S(t) is the compressive strength in MPa and t is the age in days.

For field applications, the empirical equations presented in this report can

be used as a guide for developing mixture proportions to ensure future

excavatability of CLSM or to ensure that a minimum strength is obtained to meet a

design specification. It should be noted that the derived equations have limitations.

As is the case for experimental studies, the validity of the equations is only for the
143
specific materials used in this study. It is recommended that those interested in

predicting CLSM strength perform a series of mixtures using local materials to

generate a database that can serve as the basis for predictive models.

6 R2 =0.97
5

0
0 1 2 3 4 5 6 7

Predicted strength (MPa)

Figure 5.11: Comparison of measured and predicted strengths of


non-air-entrained CLSM mixtures

Although the predictive equations presented in this report were developed

for the specific materials used in this study, a reasonable correlation was obtained

between the equations and test data from a different series of mixtures using

completely different materials sources. Table 5.7 shows the mixture proportions

for this study, along with a comparison between the measured and predicted 7- and

144
28-day compressive strength values. Although there was a reasonable correlation

between predicted and actual strengths, this correlation could have been improved

by using the specific material types and behaviors in developing materials-specific

parameters and constants (k values).

Table 5.7: Comparison between predicted compressive strengths (using


equations shown in this section) and actual strengths from previous
study (using different constituent materials)

Mixture* Cement Fly ash Concrete Water Air Flow Fresh unit
(kg/m3 ) (kg/m3 ) sand (kg/m3 ) content (mm) weight
(kg/m3 ) (%) (kg/m3 )
I-1 30 180 1500 283 1.30 200 2036
I-2 60 180 1500 327 1.25 210 2077
I-3 120 180 1500 335 0.65 220 2087
Mixture* 7 days (MPa) 28 days (MPa)
Measured Predicted Measured Predicted
I-1 0.202 0.314 0.368 0.704
I-2 0.572 0.536 1.062 1.277
I-3 1.832 1.081 2.728 2.696
* Mixture codes are different from those used in present study. These mixtures from a previous study contained high carbon
fly ash in combination with three dosages of Type I portland cement

Some interesting information on how constituent materials affect CLSM

properties was obtained from this study. Most of the significant effects were

related to the influence of water demand on compressive strength. Materials that

tended to increase water demand (such as foundry sand and high-carbon fly ash)

generally yielded lower compressive strengths than materials with lower inherent

water demands. The chemical reactivity of fly ashes was found to be critical, as the

strength of CLSM containing Class C fly ash was higher than similar mixtures

containing Class F or high-carbon fly ash. Class C fly ashes, by definition, have

higher calcium (CaO) contents than Class F fly ashes (and the high-carbon fly ash
145
used in this study) and are more effective in increasing strength, even as early as

seven days.

5.3.1.2 Effects of Cylinder Curing on Compressive Strength

This section describes the findings of investigations on the effect of cylinder

storage conditions on the unconfined compressive strength of CLSM. An initial,

smaller investigation on the effects of cylinder curing (i.e., in mold, out of mold, in

lime water) is described first (using the B-series mixtures), followed by a more

comprehensive study on cylinder curing (using the G-series mixtures). The overall

objective of these studies was to determine the most efficient and accurate

method(s) of curing CLSM test cylinders.

Table 5.8 shows the effects of curing method on mixture B-1. The findings

show that the compressive strength of CLSM can vary widely, depending on

whether the cylinders are cured in or out of their molds, or whether the cylinders

are cured out of their molds in lime-saturated water. This initial study, using only

one mixture, illustrated the importance of curing parameters, especially when

considering long-term strength values (i.e., at 91 days). The highest 91-day

strength was achieved when the cylinders were left continuously in their molds,

suggesting that cylinders stored out of their molds or in lime water may be subject

to substantial leaching. The concern over cylinder storage conditions, in particular

the potential effects of specimen leaching and loss of mass, led to further

investigations using a range of materials and mixture proportions, as described later

in this section.

146
Table 5.8: Effects of curing methods on compressive strength (Mixture B-1)

Curing 7-day strength C.O.V. 28-day strength C.O.V. 91-day strength C.O.V.
method* (kPa) (%) (kPa) (%) (kPa) (%)
Cylinders kept
in molds, 313.6 2.9 1360.9 2.0 3811.5 9.4
stored in fog
room
Cylinders were
stripped at 3 362.2 1.2 1394.7 3.8 2396.5 4.4
days and cured
in fog room
Cylinders were
stripped at 7 352.6 4.6 1157.2 4.8 2670.8 3.8
days and cured
in lime-
saturated water
* Refers to curing regimes previously described in Chapter 4. All curing was performed at 23 °C.

To further study the effects of curing regime on compressive strength, a

more comprehensive investigation was performed using ten CLSM mixtures (G-

series mixtures). A range of constituent materials and mixture proportions was

used to better understand the potential effects on strength values for different curing

methods. The curing conditions investigated included the following:

• Curing Condition A (“normal”) – Cylinder kept in mold (with cap on) for
seven days in fog room. Cylinder then stripped and kept in fog room until time
of testing. This is referred to as “normal” because most cylinders tested in this
project followed this procedure.
• Curing Condition B (“mold”) – Cylinder kept in mold (with cap on) for seven
days in fog room. Cap is then removed and cylinder is kept in fog room until
time of testing.
• Curing Condition C (“cap”) – Cylinder kept in mold (with cap on) until time of
testing.
• Curing Condition D (“out”) – Cylinder kept in mold (with cap off) outdoors
until time of testing.

147
After subjecting cylinders from each of the G-series mixtures to the curing
regimes just described, all cylinders were capped with sulfur capping compound
and tested at a loading rate of 0.38 mm/min. The compressive strengths measured
at 28 and 91 days are shown in Tables 5.9 and 5.10, respectively. The remainder of
this section discusses these results and also describes some of the nuances observed
when testing different materials and mixture proportions. Some interesting
observations were made that illustrate that the strength of CLSM is significantly
affected by variations in curing regime, and further, that these variations are a
function of specific materials and mixture proportions. Some of the more
interesting results are illustrated in Figures 5.12 to 5.15, as described later in this
section.

Table 5.9: Compressive strength at 28 days using different curing methods

Mixture Curing condition A Curing condition B Curing condition C Curing condition D


(normal) (mold) (cap) (outside)
Average C.O.V Average C.O.V Average C.O.V Average C.O.V
(kPa) (%) (kPa (%) (kPa) (%) (kPa) (%)
G-1 559.1 3.2 365.1 2.3 344.3 6.0 536.0 5.9
G-2 246.8 16.5 267.9 4.7 269.2 14.3 380.1 5.2
G-3 89.5 24.1 93.6 24.6 59.4 6.5 too weak -
G-4 247.8 11.6 167.0 16.1 164.9 4.8 - -
G-5 893.5 1.7 877.0 3.3 991.3 4.4 1259.0 4.8
G-6 369.5 7.8 326.1 10.2 306.6 7.3 295.5 8.2
G-7 150.8 11.5 145.4 4.8 161.5 3.7 280.0 8.3
G-8 170.3 9.2 137.3 6.4 160.9 16.1 600.1 8.3
G-9 317.9 12.9 328.6 13.3 279.0 18.7 496.0 14.5
G-10 486.3 3.6 412.2 9.6 411.6 6.1 986.9 4.4

In general, there were substantial differences between the strength of CLSM


cylinders stored outdoors (in hot Austin, TX weather) and cylinders cured in the
fog room. However, there was not a consistent trend for all the mixtures studied,
illustrating that temperature and curing conditions are sensitive to constituent
148
material type and proportions. Some mixtures lost strength when stored outdoors,
whereas others showed significant increases in strength when stored outdoors.

Table 5.10: Compressive strength at 91 days using different curing methods

Mixture Curing condition A Curing condition B Curing condition C Curing condition D


(normal) (mold) (cap) (outside)
Average C.O.V Average C.O.V Average C.O.V Average C.O.V
(%) (%) (%) (%)
G-1 754.6 7.7 500.7 5.0 479.4 6.0 808.8 2.3
G-2 346.1 7.8 308.6 11.1 266.5 15.3 422.7 4.5
G-3 101.9 7.3 97.4 2.0 96.3 15.5 49.6 6.1
G-4 305.8 11.4 272.9 7.9 239.0 8.1 408.5 7.1
G-5 1248.8 19.2 1342.3 14.2 1244.3 7.9 - -
G-6 378.0 13.9 367.7 3.7 366.6 7.1 378.0 13.9
G-7 175.8 10.7 179.3 4.1 199.6 7.6 271.9 8.3
G-8 218.0 2.0 210.5 8.8 201.5 12.6 943.0 2.7
G-9 408.8 6.3 347.3 14.0 353.4 24.1 501.1 13.7
G-10 785.9 17.8 688.0 6.2 617.7 5.4 1080.1 6.3

The largest increase in strength was observed for mixtures G-2, G-8, and G-

10, which contained Class C fly ash, where the high temperatures apparently

helped to activate the fly ash. Mixtures G-2, a typical “flash fill” mixture with 275

kg/m3 of Class C fly ash and no portland cement, showed a 40 percent increase in

strength when stored outside rather than in the fog room.

A similar trend is observed when CLSM containing Class C fly ash is

compared directly to Class F fly ash for mixtures containing foundry sand (Figure

5.12) and bottom ash (Figure 5.13). Mixtures containing Class F or Class C fly

ashes exhibited higher strengths for cylinders cured outdoors in a hot climate, but

the differences were more profound for Class C fly ash. In fact, the difference

149
between fog-room cured and outdoor-cured cylinders were as high as 250 percent

for the mixtures containing Class C fly ash. Thus, using laboratory-cured cylinders

to assess the field performance of CLSM containing fly ash (especially high-

calcium fly ash) in hot environments must be done with caution. The effects of

temperature on CLSM hydration are especially important when large amounts of

fly ash are used, and when the fly ash to cement ratio is high. Another impact of

higher temperature and outdoor curing (with the caps removed from the cylinders)

is the reduction in water content caused by evaporation, which lowers the effective

water-binder ratio.

Figure 5.12: Effects of curing conditions on compressive strength for Mixtures


G-7 and G-8

For high air content mixtures, G-3 and G-6, the effects of curing methods

varied with cement content (and strength levels), as shown in Figure 5.14. Mixture

G-3 contained 30 kg/m3 of cement and exhibited relatively low strengths. In fact,

150
mixture G-3 suffered such a reduction in strength when stored outdoors (compared

to fog-room curing) that cylinders could not be tested at an age of 28 days. At 91

days, cylinders could be tested, but the resultant strength was significantly lower

than fog-room cured cylinders. Mixture G-6, which contained 45 kg/m3 of cement,

exhibited less difference in strength (comparing outdoor-curing to fog-room curing)

than the lower cement content mixture (G-3). For mixture G-6, there was a

reduction in strength for outdoor-stored cylinders compared to fog-room cured

cylinders at 28 days, but this difference became negligible at 91 days. These

findings suggest that high-air content CLSM mixtures (without fly ash) benefit

more from moist-curing than they do from high-temperature exposures.

Figure 5.13: Effects of curing conditions on compressive strength for


Mixtures G-9 and G-10

151
Figure 5.14: Curing condition effects on compressive strength of Mixture G-3
and G-6

The influence of cement content on CLSM strength is also illustrated in

Figure 5.15 for mixtures containing low (15 kg/m3 ) and intermediate (30 kg/m3 )

cement contents and Class F fly ash. Mixture G-4, the low cement mixture, yielded

quite low strength values, and curing in the mold showed negative effects on the

strength when compared to stripped samples, especially after 28 days. Mixture G-

5, with the intermediate cement content, the strengths of outside cured specimens

showed 1.4 times those of specimens cured in the fog room. This further confirms

that both material type and mixture proportions affect the compressive strength of

CLSM, and that different combinations of these parameters will be affected by

methods of curing and testing. Thus, the development and recommendation of a

152
standard test method is essential in generating reproducible test results that can be

applied to specifications, design criteria, and construction guidelines.

Figure 5.15: Curing condition effects on compressive strength of Mixture G-4


and G-5

The findings obtained from the G-series mixtures regarding curing

temperature and cylinder storage led to a final investigation on the effects of curing

temperature and humidity on compressive strength, as discussed before. This

additional investigation was deemed important to better quantify the effects of

temperature on CLSM hydration and strength gain, especially when CLSM

incorporates fly ashes of varying reactivities.

153
5.3.1.3 Effects of Drainage Conditions on Compressive Strength

As described in the Chapter 4 of this dissertation, a comprehensive study

was performed to assess the effects of drainage conditions on CLSM strength. A

specially designed “curing box” was constructed to simulate the placement of

CLSM in native soil, and various drainage conditions were assessed. This section

presents the results of compression testing for cylinders stored in the following

manners:

• “Normal” – standard fog room curing (the same curing regime used throughout
most of this study)
• “No Cap” – cylinders were placed in the curing box with the cylinder caps
removed to simulate surface evaporation.
• “Cap” – cylinders were placed in the curing box with the cylinder caps left
firmly on top to simulate a situation without water loss.
• “Bottom holes” - cylinders were placed in the curing box with the cylinder caps
removed and with holes drilled in the bottom of the cylinders to simulate both
surface evaporation and bottom drainage.
• “Side holes” - cylinders were placed in curing box with the cylinder caps
removed and with holes drilled in the bottom and sides of the cylinders to
severe drainage (from top, bottom, and sides).

Additional information on this test set-up, including curing conditions, was


provided in Chapter 4. It should be noted that the curing boxes were kept in a hot
environment (in a non-air-conditioned room in the laboratory in mid-summer),
whereas the “normal” curing was performed at 23 °C in the fog room. Thus, the
effects of both temperature and drainage could be observed. The “cap” curing, as
previously described, was a useful reference in that these cylinders followed the

154
same time-temperature history as the other cylinders stored in the curing box, but
the cylinders were not subjected to drainage or evaporation.

The compressive strength test results at 7 and 28 days are shown in Table

5.11 and Table 5.12, respectively. Note that mixture F-2, which was a Class C fly

ash mixture without cement, was not tested at 28 days, due to the rapid setting time

of this mixture, which made it difficult to cast sufficient specimens.

In general, there were not substantial differences in compressive strengths

for the four curing methods that utilized the curing box, suggesting that the

influence of drainage and evaporation may not be critically important, at least

based on this laboratory simulation study. If differences were observed for a given

mixture, the difference was more apparent at 7 days than at 28 days. Once again,

the effects of temperature were evident as cylinders cured in the plywood box set-

up (ambient hot weather) tended to exhibit higher strengths than companion

cylinders cured in the fog room. This was particularly evident for mixtures

containing fly ash. The only fly ash mixture that did not show higher strengths in

the curing box, compared to the fog room, was mixture F-2, which was a so-called

“flash fill” mixture. This mixture set extremely fast, with little or no water

available for seepage or evaporation. Also, because the mixture set so fast and

hydrated so rapidly within the first few hours, the effects of long-term temperature

differences were probably not as important.

Another interesting observation, which confirmed some earlier findings,

was that the effects of curing temperature on strength may be quite complex

because it is not just a function of constituent material type but also of material

155
quantity. For instance, when low cement contents (15 kg/m3 ) were used in

combination with fly ash, such as in mixtures F-4, F-6, and F-8, there were only

small differences in strength as curing temperature was increased. However when

the cement content was increased to 30 kg/m3 (with the fly ash content remaining

constant), the temperature effects were significantly more pronounced, such as in

the case of mixtures F-5 and F-7. Thus, one must consider the combined effects of

material type and quantity when assessing the long-term strength gain of CLSM,

especially when using fly ash.

Table 5.11: Compressive strength at 7 days under different drainage


conditions
Mixture Normal No cap Cap Bottom holes Side holes
Average C.O.V. Average C.O.V. Average C.O.V. Average C.O.V. Average C.O.V.
(%) (%) (%) (%) (%)
F-1 308.8 1.5 496.1 6.2 377.3 22.4 469.1 10.1 559.9 1.9
F-2 105.3 5.8 101.8 - 110.0 0.4 110.1 3.3 116.7 -
F-3 84.1 5.0 86.6 22.4 81.7 19.8 97.6 2.6 99.4 15.2
F-4 127.8 7.1 107.2 16.2 105.8 16.7 93.2 11.6 116.8 16.0
F-5 303.8 1.9 435.7 4.0 463.6 5.1 485.6 7.7 523.2 1.7
F-6 132.8 4.1 170.3 4.0 141.7 4.3 196.4 3.9 192.8 1.1
F-7 252.4 10.4 360.7 6.1 388.5 1.9 411.7 7.6 407.8 8.0
F-8 162.5 8.0 187.3 4.0 201.6 6.7 192.2 11.6 212.6 7.5

156
Table 5.12: Compressive strength at 28 days under different drainage
conditions
Mixture Normal No cap Cap Bottom holes Side holes
Average C.O.V. Average C.O.V. Average C.O.V. Average C.O.V. Average C.O.V.
(%) (%) (%) (%) (%)
F-1 532.3 16.3 615.1 14.4 511.5 29.7 554.2 19.4 658.7 9.0
F-2* - - - - - - - - - -
F-3 144.1 14.2 138.4 14.2 132.0 17.7 153.6 25.6 157.3 30.0
F-4 291.9 5.7 309.0 7.5 253.1 16.3 250.7 6.8 283.6 10.2
F-5 842.1 9.0 1497.2 3.1 1329.3 3.6 1576.2 5.0 1586.0 3.6
F-6 255.1 2.7 334.3 18.6 297.4 14.1 394.8 8.6 412.3 9.2
F-7 779.5 14.1 1215.4 8.1 1412.6 2.2 1217.2 4.0 1278.3 1.7
F-8 493.1 6.9 499.8 5.7 462.5 2.4 528.7 3.8 560.3 4.0
* Cylinders from mixture F-2 were not tested at 28 days.

In summary, this investigation confirmed that temperature plays a key role

in many CLSM mixtures and suggests that drainage and evaporation may not be as

critical as temperature-induced effects. Additional work is needed to investigate

CLSM placed in actual applications to gain insight into real drainage conditions, as

opposed to the simulated conditions used in this study.

5.3.1.4 Specimen Preparation Before Testing

This section describes the findings on the effect of cylinder conditioning on

the unconfined compressive strength of CLSM. A preliminary study on cylinder

conditioning (i.e., drying specimens after curing but before testing) is described.

As previously mentioned, ASTM D 4832 specifies that cylinders be allowed

to air-dry for 4 to 8 hours before capping. It was observed that this long waiting

period may be inconvenient and impractical in a testing laboratory and may also

lead to variations in test results. To investigate this, cylinders from mixture B-2

157
were moist-cured for various ages, removed from the fog room, then subjected to

the following conditions:

• Cylinders air-dried for 0 to 0.5 hour before being sulfur-capped and


tested
• Cylinders air-dried for 2 hours before being sulfur-capped and tested
• Cylinders air-dried for 4 hours before being sulfur-capped and tested
• Cylinders air-dried for 8 hours before being sulfur-capped and tested

The results of testing after 7, 28, and 91 days of curing are shown are shown

in Table 5.13. Although there were some variations in strength results at each age

when the specimens were allowed to dry for different periods of time, the

variations were not appreciable in most cases. This suggests that it is not necessary

to air dry specimens for 4 to 8 hours before testing, as is specified in the ASTM

standard. This important finding might be incorporated into a proposed

compression test method, in which cylinders are specified to be tested without

undergoing a mandatory drying period. This recommendation is consistent with

the testing of conventional concrete. Also, by eliminating the drying period, it is

expected that more accurate results may be obtained, and laboratory productivity

will increase (i.e., more cylinders can be tested in a single day).

Table 5.13: Influence of specimen air-drying on compressive strength


(Mixture B-2)
Test age 0-0.5 hour air dry 2 hours air dry 4 hours air dry 8 hours air dry
Strength C.O. V. Strength C.O. V. Strength C.O. V. Strength C.O. V.
(kPa) (%) (kPa) (%) (kPa) (%) (kPa) (%)
7 days 310.1 5.8 329.2 11.3 357.5 2.7 363.2 2.0
28 days 1575.1 1.9 1536.5 7.8 1447.2 5.7 1649.8 3.3
91 days 3289.4 2.0 3226.9 4.6 3430.4 3.9 3536.1 4.6

158
5.3.1.5 Alternative Capping Materials for Compression Testing

A comprehensive study was performed on the use of alternative capping

materials for compression testing, as already described in Chapter 4. The materials

included in this investigation were neoprene pads with Shore “A” durometer values

of 20, 40, 50, 60, and 70, as well as sulfur-capping compound and gypsum (or

“hydrostone”). After curing the CLSM cylinders for various ages (7, 28, and 91

days), the cylinders were capped and tested using a similar approach which

followed throughout most of this project. A load rate of 0.38 mm/min was used for

all cylinders, and load and deformation values were recorded by a data acquisition

system.

Initial trials using gypsum led to the development of a more user-friendly

method of using this capping material. When using the gypsum, a water-gypsum

ratio of 0.3 was used, and a split 75-mm x 150-mm plastic mold was used as a

“dam” to retain the fluid gypsum mixture on top of the cylinder end. Sealing

grease was used to prevent the gypsum mixture from leaking through the gap

between the cylinder surface and plastic mold. Because of the self-leveling

property of gypsum mixtures, cylinders with damaged ends could be easily capped.

The long setting time of gypsum mixtures, up to 40 minutes, made the capping

process time consuming, and thus, gypsum was only used for testing at 28 days.

Table 5.14 shows the compressive strength values measured for the

different mixtures and capping materials at 7, 28, and 91 days. Table 5.15 shows

the corresponding coefficients of variation of measured compressive strength using

the various capping methods. The sulfur capping method provided the highest

159
strength values for all the eight mixtures. Also, in general, sulfur capping

generated the lowest variations compared to the other capping methods. Note that

Mixture E-2, a “flash fill” mixture, set so quickly that it was not possible to

produce enough cylinders for testing at all ages. Thus, testing was only performed

at an age of 7 days.

Table 5.14: Compressive strength results using different capping materials


Mixture E-1 E-2 E-3 E-4 E-5 E-6 E-7 E-8
Sulfur 7 day 0.31 0.11 0.08 0.13 0.20 0.34 0.43 0.19
28 day 0.53 0.14 0.29 0.33 1.03 1.19 0.66
91 day 0.85 0.14 0.45 0.48 1.58 1.71 1.24
Gypsum 28 day 0.52 0.11 0.24 0.36 0.95 1.11 0.59
D70 7 day 0.28 0.09 0.06 0.12 0.20 0.28 0.36 0.17
28 day 0.42 0.11 0.24 0.34 0.92 1.05 0.42
91 day 0.10 0.34 0.46 1.12 1.39 0.91
D60 7 day 0.29 0.09 0.05 0.11 0.16 0.33 0.36 0.18
28 day 0.57 0.12 0.26 0.27 0.66 1.02 0.42
91 day 0.10 0.37 0.31 1.29 1.50 0.89
D50 7 day 0.26 0.08 0.05 0.13 0.21 0.34 0.40 0.19
28 day 0.51 0.10 0.25 0.28 0.87 0.96 0.47
91 day 0.10 0.35 0.32 1.23 1.44 1.08
D40 7 day 0.27 0.07 0.04 0.13 0.22 0.33 0.34 0.21
28 day 0.51 0.10 0.22 0.25 0.91 1.01 0.50
91 day 0.79 0.14 0.33 0.45 1.36 1.37 0.80
D20 7 day 0.25 0.09 0.04 0.11 0.20 0.32 0.40 0.17
28 day 0.71 0.09 0.24 0.31 0.90 0.96 0.57
91 day 0.13 0.40 0.36 1.30 1.43 1.05

160
Table 5.15: Coefficients of variation for compressive strengths using different
capping materials
Mixture E-1 E-2 E-3 E-4 E-5 E-6 E-7 E-8
Sulfur 7 day 1.6 5.8 5.0 7.1 15.7 17.6 8.4 6.9
28 day 16.3 14.2 5.7 20.7 13.3 4.0 10.3
91 day 3.6 15.3 10.6 10.5 6.9 5.3 8.1
Gypsum 28 day 18.0 7.7 6.1 5.5 5.3 7.8 14.9
D70 7 day 5.3 12.4 20.3 5.9 3.7 12.0 9.7 9.5
28 day 45.9 19.3 22.0 44.5 16.1 14.9 20.8
91 day 22.8 6.8 24.9 8.3 3.0 3.7
D60 7 day 18.0 6.5 15.5 7.2 17.3 12.0 13.2 22.1
28 day 12.9 2.9 2.5 4.0 23.9 7.6 22.0
91 day 30.4 12.9 16.4 12.5 6.6 21.4
D50 7 day 17.5 5.7 28.7 15.9 7.4 0.7 4.6 30.0
28 day 18.1 15.5 5.0 20.9 15.1 6.4 6.5
91 day 17.0 22.2 12.6 4.2 4.0 11.5
D40 7 day 17.9 27.6 23.3 21.4 10.1 13.0 6.8 15.2
28 day 15.4 14.0 9.5 20.5 19.5 10.8 11.2
91 day 12.7 2.7 4.6 27.8 6.6 14.6 9.4
D20 7 day 16.5 1.9 20.5 15.8 10.3 7.3 9.6 16.6
28 day 7.1 1.1 14.5 10.1 19.2 6.8 5.9
91 day 6.1 7.4 22.2 9.1 6.0 23.5

Figures 5.16 through 5.18 compare the coefficients of variation of strength


values obtained by using different capping methods at 7, 28, and 91 days,
respectively. Neoprene pads resulted in as high as 45% variation in the measured
compressive strength values, with the hardest pads resulting in the largest
variations.

161
Figure 5.16: Effects of capping methods on coefficient of variation (7-day
compressive strength)

162
Figure 5.17: Effects of capping methods on coefficient of variation (28-day
compressive strength)

Figure 5.18: Effects of capping methods on coefficient of variation (91-day


compressive strength)

163
The process for the qualification of unbonded capping systems described in

ASTM C 1231 was followed in analyzing the test data in this study. Under close

examination, the relationship expressed as equation (3) in the test method was

found to be only approximate. Specifically, it involves a “student t” testing with

the null hypothesis that xp –0.98xs ≥ 0. Therefore, the calculation of standard

deviation of the difference should be based on xp and 0.98 xs. When testing

concrete according to ASTM C 1231, the strength from different capping methods

is assumed to be essentially equal, and the error using the proposed process can

typically be neglected. However, it has been shown that different capping materials

yield different strength values for CLSM and as such, the validity of the specified

ASTM C 1231 analysis process was examined in more detail. The theoretically

correct process should calculate the difference in strength for each pair of cylinders

using the following equation:

d i = x pi − ax si (5. 3)

Where:

di = Difference in strength of a cylinder with unbonded capping compared

to a percent of strength of another cylinder with sulfur capping,

a = Percent. Strength measured with unbonded capping should be larger

than “a” percent of strength measured with sulfur capping at a 95% confidence.

The initial estimate of a0 was selected to be the value obtained through the

ASTM C 1231 process satisfying equation (3) in ASTM standard. Adhering to the

164
process outlined in ASTM C 1231-99 (using the proposed equation 5.4 above, a

value of b0 could be obtained by satisfying:

x p = b0 a0 xs + (tsd ) /( n)1 / 2 (5. 4)

If b0 is equal to or very close to 1.0, the initial estimate could be deemed

correct. Otherwise, an iterative process should be followed, letting a1 =a0 × b0 , and

repeating until converging on an accurate value. Usually, only one or two cycles

were needed for convergence. Table 5.16 illustrates the qualification of unbonded

caps using the unmodified ASTM C 1231 process and the proposed, modified

method. For each durometer hardness and gypsum cement, more than 40 pairs of

specimens were included in the statistical analysis. The differences between the

two processes were found to be within six percent of each other, a relatively small

difference. However, it was believed that it was worthy of consideration, given the

differences in strengths observed using the different capping methods.

Table 5.16: Comparison of “a” values calculated by two methods


Capping system “a” value calculated by “a” value calculated by
ASTM C 1231 process modified ASTM C 1231
Gypsum 0.9203 0.9205
D 20 0.7990 0.8126
D 40 0.7316 0.7914
D 50 0.8032 0.8081
D 60 0.7629 0.7705
D 70 0.7512 0.7379

165
ASTM D 4832 states that the alternative capping systems are acceptable

when the average strength obtained is not less than 80% of the average strength of

companion cylinders capped with sulfur capping compound. According to this

criterion, the use of gypsum and neoprene pads with durometer values of 20 and 50

could be qualified when using the ASTM C 1231 process. By using the modified

method (as shown in Table 5.16), gypsum and neoprene pads with durometer

values of 20, 40, and 50 could be qualified.

As previously described, full load-deformation curves were recorded for

each of the compression tests. Rather than providing all of this data in this

dissertation, some representative results are presented next. Figure 5.19 shows a

typical load deformation curves for compressive cylinders of mixture E-1 at 28

days. Because the deformation-measuring device records the total movement

between loading platens, the compression of neoprene pads is also included the

deformation value. As a result, higher deformation values and longer loading times

were recorded in the testing. Thus, care should be taken in performing tests and

analyzing data when different capping materials are used because both the

measured deflection and testing time are directly influenced. Unlike concrete

cylinders, where deflections can be measured from surface mounted gages, it is

difficult to measure deflections or strains from the cylinder surface because of the

inherent low strength of CLSM. For CLSM, the most relevant data can be obtained

from using rigid caps (gypsum or sulfur) because the measured deflections (based

on relative platen movement) are less affected by deflections in the capping

material.

166
4
D60
3.5

3 D20
D20
2.5 D40
Load (kN)

D70 D50
2
Sulfur D40 D60
1.5
D70
1 Gypsum Gypsum
D50
0.5 Sulfur

0
0 100 200 300 400 500 600 700
Displacement (mm)

Figure 5.19: Load-deflection curves for mixture E-1 (compression


testing at 28 days)

In summary, the main findings from this capping material study include the

following:

1. Sulfur-capped compression cylinders generally produced the highest

strength values with the lowest variations.

2. Gypsum-capped specimens generally yielded higher strength values

than those using neoprene pads. However, gypsum takes a

relatively long time to set and may not be as user-friendly in a

testing laboratory environment.

3. There was not a significant difference between using different

durometer pads, especially when considering the variations in

167
results. However, when applying the acceptance criteria described

in ASTM C 1231, only neoprene pads with durometer values of 50

or less were found to be acceptable for use with CLSM.

5.3.1.6 Effects of Load Rate on Compressive Strength

Because of the general lack of guidance provided in ASTM D 4832

regarding the loading rate, additional emphasis was placed on assessing the effects

of loading rate on the compressive strength of CLSM in this study. This section

describes the results of two studies, the first of which focused on a relatively wide

range of loading rates (0.085 to 0.58 %/min), and the second of which covered a

narrower range of loading rates (0.16 to 0.42 %/min). A deflection-controlled

machine, often used to test soils, was used for this investigation. The results of

these two studies, coupled with the findings from the initial investigation, could be

used to recommend loading rates for a refined method to measure the unconfined

compressive strength of CLSM.

The loading rate is an important parameter when considering compression

testing. The testing of CLSM cylinders should be accurate. If strength values are

found to be influenced by loading rate, then a load rate range must be defined and

required for accurate testing. The loading rate also determines the length of time

needed to test a given cylinder. This required length of time must be sufficient to

ensure accuracy (i.e., not a sudden cylinder failure) and to complete a given test in

a reasonable amount of time. In a laboratory or testing facility, the time required to

test cylinders may be critical, especially if many tests are performed in a single day.

168
The compressive strength results for the two loading rate range studied,

referred to as “wide range” and “narrow range” studies, shown in Tables 5.17 and

5.18. The cylinder measures 75 mm x 150 mm were used in this study. The full

load-deflection curve was also obtained for each cylinder tested, and the

deformation at peak load for each test specimen was recorded, as shown in Tables

5.19 and 5.20. The deformation at peak load was used to determine if changes in

loading rate resulted in changes in modes of failure, such as a change from

relatively ductile to brittle failure. Discussion follows on how loading rates

affected both the compressive strength of CLSM and the deformation at peak load.

Table 5.17: Effects of loading rates (wide range) on compressive strength

Age Mixture 0.085 %/min 0.25 %/min 0.58 %/min


(days) Average C.O.V. Average C.O.V. Average C.O.V.
(kPa) (%) (kPa) (%) (kPa) (%)
A-1 291.5 4.0 344.9 5.5 373.6 7.0
7 A-2 126.0 6.0 108.6 11.4 120.1 2.7
A-3 222.5 5.2 245.7 7.6 269.2 5.7
A-4 128.6 6.1 132.4 6.1 132.8 8.9
A-1 867.8 5.7 914.5 11.7 962.5 4.4
28 A-2 227.1 4.4 238.9 2.3 222.9 5.2
A-3 675.1 2.9 708.1 0.8 682.4 2.0
A-4 587.4 4.4 634.1 3.8 630.3 4.2
A-1 1166.8 6.8 1196.2 2.7 1377.5 0.7
91 A-2 410.9 4.3 429.7 3.1 409.0 2.0
A-3 891.4 12.5 886.7 2.1 987.5 9.6
A-4 1439.7 8.6 1221.5 12.8 1239.5 5.3

169
Table 5.18: Effects of load rates (narrow range) on compressive strength

Age Mixture 0.16 %/min 0.25 %/min 0.33 %/min 0.42 %/min
(days) Average C.O.V. Average C.O.V. Average C.O.V. Average C.O.V.
(kPa) (%) (kPa) (%) (kPa) (%) (kPa) (%)
A-5 533.4 9.1 532.6 4.6 530.8 11.3 565.7 7.3
7 A-6 279.1 4.3 291.3 6.4 288.8 1.9 310.1 2.3
A-7 193.5 8.0 190.5 3.1 188.0 2.2 220.1 8.4
A-5 1050.1 7.1 1090.5 5.0 1048.3 4.1 1066.7 5.0
28 A-6 752.0 9.0 745.3 5.3 760.4 12.1 779.3 10.9
A-7 1039.1 5.4 1022.0 3.5 1085.3 3.6 1078.8 4.0
A-5 1339.5 3.4 1614.4 7.6 1529.3 8.4 1590.2 4.5
91 A-6 829.2 9.0 990.3 5.3 1079.7 12.1 1016.5 10.9
A-7 1871.9 5.4 1784.3 3.5 1863.7 3.6 1983.8 4.0

Table 5.19: Effects of loading rates (wide range) on deformation at peak load
0.085 %/min 0.25 %/min 0.58 %/min
Age Mixture Deformation C.O.V. Deformation C.O.V. Deformation C.O.V.
(days) (mm) (%) (mm) (%) (mm) (%)
A-1 0.711 23.9 0.673 2.7 0.803 20.2
7 A-2 4.252 16.3 3.459 15.9 4.252 2.3
A-3 0.960 23.5 0.892 28.6 0.765 13.4
A-4 1.521 13.4 1.547 19.8 1.402 5.1
A-1 0.572 13.5 0.566 11.8 0.566 6.4
28 A-2 1.466 13.3 1.801 17.2 1.537 11.8
A-3 0.665 2.3 0.861 12.3 0.848 5.1
A-4 0.648 1.4 0.683 20.9 0.691 14.4
A-1 0.716 0.7 0.597 1.7 0.638 1.9
91 A-2 1.920 0.4 2.042 2.7 1.880 3.4
A-3 0.909 1.8 0.925 1.3 0.851 2.6
A-4 0.777 1.7 0.843 4.7 1.031 3.9

170
Table 5.20: Effect of narrow range load rates (narrow range) on deformation
at peak load
0.16 %/min 0.25 %/min 0.33 %/min 0.42 %/min
Age Mixture Deformation C.O.V. Deformation C.O.V. Deformation C.O.V. Deformation C.O.V.(
(days) (mm) (%) (mm) (%) (mm) (%) (mm) %)
A-5 0.577 15.8 0.526 14.5 0.480 34.1 0.434 31.0
7 A-6 0.648 11.6 0.620 12.2 0.627 9.0 0.599 8.1
A-7 1.229 17.9 1.224 31.1 1.166 7.6 1.077 2.9
A-5 0.610 16.0 0.726 20.3 0.650 9.7 0.528 11.6
28 A-6 0.711 7.2 0.640 18.7 0.681 3.0 0.759 5.1
A-7 0.838 12.3 0.861 12.2 0.780 12.8 0.828 23.4
A-5 0.493 4.8 0.732 12.9 0.759 3.2 0.602 6.9
91 A-6 0.866 20.5 0.676 6.7 1.036 17.7 1.214 13.5
A-7 0.775 21.5 0.640 9.6 0.813 20.4 0.808 24.1

The effects of load rate on compressive strength were found to vary,

depending upon the type of mixture and the age of testing. Interestingly, some

mixtures (i.e., mixture A-2) were relatively insensitive to load rate, whereas others

were quite sensitive. There was no consistent trend among the mixtures, which

suggested the compressive strength was either directly or indirectly proportional to

load rate. However, there were some general trends that were insightful, as

described next.

For testing at 7 and 28 days, the range of loading rates from 0.16 to 0.42

%/min was generally found to generate consistent results. The effects of loading

rate on strength were more significant at 91 days than at 7 or 28 days. Mixture A-2

(containing foundry sand) was an interesting case where the compressive strength

varied very little, even across a wide range of loading rates. This is advantageous

from a testing perspective because the deformations at peak load of specimens

containing foundry sand were found to be large, and testing in some cases took

more than 10 minutes per sample when using a loading rate of 0.25 %/min. Higher
171
loading rates could thus be used for these types of mixtures to speed up the testing

without affecting test result accuracy. Because the variations in compressive

strength with loading rate were to some extent mixture or material specific, a load

rate between 0.and 0.42 %/min may be selected for improved compressive strength

testing.

The deformations at peak load provided a useful index of the relative

brittleness or ductility of CLSM. The effects of curing time on CLSM ductility

were profound, where at early ages (7 days), the deformations were somewhat

greater than those at later ages (28 and 91 days), as the specimens were more

ductile and extensible.

In general, when varying the loading rate, the deformations observed at

peak load did not vary as much as the actual compressive strength values. This

suggests that the mode of failure was not greatly affected by loading rate

modifications. Thus, the range of loading rates between 0.16 to 0.42 % per minute

can be recommended for the compression test method, based mainly on selecting a

range that yielded accurate and repeatable strength values in a reasonable amount

of time.

5.3.1.7 Effect of Specimen Size

Table 5.21 shows a comparison between different size cylinders, in which

no discernible size effects were observed. It should be noted that all cylinders were

tested at the same effective strain rate (0.25 %/min), using the same deflection-

172
controlled machine. There is no general trend found for the effects of specimen

size.

Table 5.21: Effects of cylinder size on unconfined compressive strength


Mix 7 days 28 days
75 x 150 mm 100 x 200 mm 75 x 150 mm 100 x 200 mm 150 x 200 mm
MPa (C.O.V.:%) MPa (C.O.V.:%) MPa (C.O.V.:%) MPa (C.O.V.:%) MPa (C.O.V.:%)
4 0.79 (11.0) 1.13 (7.01)
4r 0.58 (6.2) 0.49 (6.3) 0.94 (3.9) 0.84 (6.5)
6 0.47 (0.7) 0.50 (8.8) 0.81 (4.9) 0.87 (2.0) 0.81 (10.9)
23 0.04 (6.4) 0.07 (7.2) 0.14 (9.5) 0.21 (8.3) 0.19 (3.6)
24 0.22 (1.4) 0.30 (11.1) 0.44 (0.1) 0.59 (3.9) 0.52 (1.8)
22r 0.50 (9.7) 0.64 (2.5) 0.93 (17.5) 0.87 (2.5) 0.72 (16.2)
26 0.76 (8.2) 0.70 (7.8) 1.11 (15.8) 1.09 (27.2) 1.08 (2.0)

5.3.1.8 Effects of Different Testing Machine

Most of the research on compression testing performed in this study utilized

a smaller capacity testing machine under deflection control. However, many

laboratories use larger capacity concrete compression machines under load rate

control. To assess the relative difference in strength values, a limited study was

performed to compare the results of a small capacity machine (Instron, 100 kN

capacity) under deflection control with a larger capacity machine (1780 kN Tinius

Olson) under load control. For instance, for specimens from mixture #22 at the age

of 7 days, the small machine yielded a strength of 0.50 MPa and a COV of 1.4%,

and the large machine yielded a compressive strength of 0.24 MPa and a COV of

22.4%. The larger compression machine showed significantly lower strength

values and significantly higher variations than those obtained with the smaller

machine. Thus, caution should be taken when using a large capacity machine

173
(Tinius Olson, 1780 kN capacity) under load control. The results from this

investigation are summarized in Table 5.22. It should be noted that in order to

meet the-time-to-failure limits described in ASTM D 4832 (not less than two

minutes), a load rate of 6.9 kPa/sec was used for the large machine. In general, the

results obtained from the larger compression machines lack the accuracy in the

lower range of load values to assess the strength of CLSM.

Table 5.22: Effects of different testing machines on compressive strength


(coefficient of variation is included in parentheses)
Age 7 days 28 days
Specimen 75 x 150 100 x 200 75 x 150 100 x 200
Size (mm)
Machine Small Large Small Large Small Large Small Large
Strength Strength Strength Strength Strength Strength Strength Strength
MPa (%) MPa (%) MPa (%) MPa (%) MPa (%) MPa (%) MPa (%) MPa (%)
Mix #4 0.48 (6.2) 0.26 (15.8) 0.79(11.0)
Mix #4r 0.49 (6.3) 0.48 (12.2) 0.94 (3.9) 0.84 (6.5) 0.81 (22.8)
Mix #6 0.47 (0.7) 0.35 (7.3) 0.50 (8.8) 0.43 (9.3) 0.81(4.9) 0.66 (20.6) 0.87 (2.0) 0.84 (2.3)
Mix #23 0.04 (6.4) 0.08 (69.6) 0.07 (7.2) 0.10 (33.8) 0.14(9.5) 0.16 (14.2) 0.21 (8.3) 0.15 (33.2)
Mix #24 0.22 (1.4) 0.24 (28.9) 0.30 (11.1) 0.26 (39.3) 0.44(0.1) 0.35 (12.7) 0.59 (3.9) 0.37 (10.1)
Mix #22r 0.50 (9.7) 0.31 (22.4) 0.64 (2.5) 0.48 (19.6) 0.93(17.5) 0.82 (10.3) 0.87 (2.5) 0.70 (0.5)
Mix #26 0.76 (8.2) 0.35 (7.3) 0.70 (7.8) 0.65 (13.9) 1.11 (15.8) 0.96 (2.1)

5.3.1.9 Effects of Curing Temperature and Humidity

The study summarized in this section was a follow-up to previous testing

that suggested that temperature plays a major role in CLSM strength development.

Three curing temperatures (10 ºC, 21 ºC, and 38 ºC) and six CLSM mixtures (H-1

to H-6 in Table 3.16) were selected to study the strength gain of CLSM across a

range of practical construction conditions. After casting the specimens into plastic

174
molds, the cylinders were immediately transported to the appropriate temperature-

controlled chambers. After three days of storage in the chambers, half the cylinders

from each mixture were stripped or removed from the molds and returned to the

same chamber until the time of testing. This curing regime is referred to as “dry”

in subsequent discussions. The other half of the specimens from a given mixture

were kept inside the molds with the caps firmly on the top of the mold until the day

of testing (designated as “wet” curing). These cylinders were placed directly next

to the cylinders that had already been stripped.

The results of compression testing at 7, 28, and 91 days are shown in Table

5.23. Also included in Table 5.23 are the moisture contents of specimens measured

immediately after compression testing to assess the effects of curing conditions on

the moisture content (or evaporable water content) and strength of CLSM. The test

results at 7, 28, and 91 days are given in Table 5.23 and illustrated in Figures 5.20

through 5.25.

The results summarized in Figures 5.20 to 5.25 confirm the earlier

conclusion that temperature and storage conditions play a major role in the

compressive strength of CLSM cylinders. Once again, it was difficult to discern

clear trends in behavior across all material types and mixture proportions, but

several general observations could be made. These include:

• Mixtures containing fly ash exhibited significant strength gains at 38 °C,


compared to lower temperatures. As expected, the increase in strength was
more pronounced for Class C fly ash, compared to Class F fly ash, mainly
due to the difference in reactivity. For example, compared to cured at 21 °C
, mixture H-5, containing Class C fly ash, showed a 160% increase in

175
strength when cured at 38 ºC, whereas mixture H-2, containing Class F fly
ash, showed a 40% increase in strength for the same conditions. The CaO
content of fly ash is generally the most important variable affecting
compressive strength and should be taken into account when using CLSM,
especially when temperature extremes may be encountered.

176
Table 5.23: Compressive strength of specimens cured at different
temperatures at different ages

Age 7 days 28 days 91 days


Mix Temp* Moisture Compressive C.O.V. Moisture Compressive C.O. V. Moisture Compressive C.O.V.
# content strength (%) content strength content strength
(ºC) (%) (kPa) (%) (kPa) (%) (%) (kPa) (%)
10 dry 36.9 328.6 3.5 17.8 548.5 7.4 2.0 258.5 2.7
10 wet 38.0 210.5 6.0 37.1 240.9 7.8 27.0 323.4 5.5
H-1 21 dry 18.1 367.5 2.8 1.6 607.3 6.5 1.1 480.2 6.5
21 wet 38.0 299.2 23.0 37.8 266.2 11.8 26.8 632.3 3.3
38 dry 19.5 828.5 2.2 1.8 722.3 4.4 1.1 756.7 12.2
38 wet 22.6 440.4 13.5 35.3 802.2 2.4 26.0 917.3 0.4
10 dry 7.4 188.3 10.2 1.2 314.1 7.0 2.7 118.9 13.6
10 wet 10.3 151.9 8.6 9.6 194.2 12.3 5.5 260.5 3.6
H-2 21 dry 1.8 214.1 2.3 0.3 171.4 10.1 6.5 142.7 7.5
21 wet 9.8 172.4 2.1 9.1 233.1 6.5 3.3 345.3 10.2
38 dry 0.3 256.6 12.2 0.2 211.9 4.9 12.2 263.6 3.2
38 wet 8.6 220.9 11.7 8.9 458.2 7.3 0.4 634.6 2.9
10 dry 6.0 1384.3 1.5 1.6 2315.4 4.9 1.1 1395.4 13.0
10 wet 8.7 937.7 16.2 9.1 1141.2 8.9 7.9 1367.9 3.2
H-3 21 dry 1.6 1213.4 5.1 0.6 963.6 9.6 0.5 852.9 8.9
21 wet 8.7 695.7 8.8 8.6 786.2 3.2 7.4 919.8 6.7
38 dry 0.3 1222.5 6.8 0.0 944.5 3.8 0.4 1042.0 13.5
38 wet 6.3 864.0 2.1 6.2 3844.8 7.4 2.3 3880.9 20.4
10 dry 7.4 314.0 2.8 1.4 1486.2 2.8 1.1 895.8 14.8
10 wet 10.0 185.9 9.4 37.7 893.2 9.6 8.1 1670.3 4.9
H-4 21 dry 1.2 669.6 3.0 0.3 628.6 9.8 0.4 458.6 9.9
21 wet 8.8 501.7 4.1 7.7 1570.7 7.3 3.8 3743.6 4.9
38 dry 0.4 2615.0 8.5 0.9 2041.2 3.5 0.3 2060.6 5.6
38 wet 5.9 2098.8 9.2 3.7 12116.8 11.1 1.4 11512.6 7.0
10 dry 8.7 273.1 6.9 1.4 711.4 3.2 0.8 421.9 5.8
10 wet 10.0 232.6 4.5 9.7 544.5 9.6 8.6 1362.6 6.3
H-5 21 dry 1.4 420.8 3.9 0.3 411.2 5.3 0.3 330.7 12.1
21 wet 10.7 316.8 6.7 10.2 815.2 2.9 8.8 1497.7 4.6
38 dry 0.3 1524.7 9.1 0.2 1423.7 4.6 0.1 1339.4 11.9
38 wet 7.4 1472.5 7.3 7.7 2282.0 8.6 3.5 2638.2 12.0
10 dry 6.3 281.9 16.7 1.6 740.5 8.7 1.1 669.3 3.4
10 wet 8.7 210.6 6.9 8.1 470.5 1.2 7 922.2 5.6
H-6 21 dry 1.0 480.4 16.9 0.3 434.5 15.0 0.3 372.9 18.7
21 wet 7.5 371.4 15.8 6.2 744.7 7.2 6.2 929.6 9.4
38 dry 0.4 816.8 11.9 0.2 828.0 28.1 0.2 782.0 4.1
38 wet 6.1 562.5 10.7 4.8 786.3 12.6 0.4 991.3 7.7
*: dry = cylinders stripped after 3 days; wet = cylinders kept in mold until time of testing

177
Figure 5.20: Compressive strength of mixture H-1 for different curing
conditions

Figure 5.21: Compressive strength of mixture H-2 for different curing


conditions

178
Figure 5.22: Compressive strength of mixture H-3 for different curing
conditions

Figure 5.23: Compressive strength of mixture H-4 for different curing


conditions

179
Figure 5.24: Compressive strength of mixture H-5 for different curing
conditions

Figure 5.25: Compressive strength of mixture H-6 for different curing


conditions

180
• Drying (from an age of three to seven days) generally increased the 7-day
strength of CLSM, compared to keeping the cylinders in molds
continuously for 7 days. For 91-day testing, the opposite effect was
generally observed. That is, cylinders continuously left in sealed plastic
molds exhibited higher strengths than companion cylinders that were
allowed to dry after the first three days. The results at an intermediate age
of 28 days were not as well defined, and general trends were not evident.
• The measurement of the moisture content of cylinders at the time of testing
may provide useful information on its performance, especially when
subjected to drying conditions. In some cases, the reduction in moisture
content for cylinders cured outside of their molds at higher temperatures
was accompanied by microcracking and subsequent strength reductions.
• The standard curing methods might severely under-estimate CLSM strength
development in the field, especially in humid environments. On the
contrary, for the case of mixture H-4, the dry specimens at 21 ºC only
developed 25% of compressive strength estimated in the lab. When
choosing a CLSM mixture, its proportions should be carefully studied in
relation to its application environments, such as temperature and loss of
moisture.
• In general, mixtures without fly ash were less sensitive to curing
temperature and drying conditions. The predictive models described earlier
in this chapter were generally effective in predicting CLSM without fly ash,
but significant differences were evident when applying the predictive
models to fly ash mixtures. Thus, particular attention should be paid to
assessing the long-term strength gain of CLSM containing fly ash,
especially when high-calcium ashes are used in warmer climates.
• Field experience is essential in understanding the performance of local
materials in specific applications and climates. The experience gained is

181
particularly important in assessing long-term strength gain and its effects on
excavatability.
• An assessment of the on-site strength of CLSM should take into account
laboratory-obtained test results, but it should also take into account climatic
conditions. An understanding of material reactivity is helpful in
extrapolating laboratory results to field performance.

5.3.2 Triaxial Shear Strength

Using the I-series CLSM mixtures, the triaxial shear strength of several

CLSM mixtures was measured. The results, shown in Table 5.24, confirmed the

observation of other researchers that the strength of CLSM is composed of both

chemical bonding and internal frictional resistance. For the mixtures investigated

in the present study, the behavior was found to be a function of specific material

and mixture proportions, and the effects changed with increased curing time.

By using the approach of stress paths, the strength parameters of CLSM

mixtures were calculated. For mixtures I-1, I-2, and I-3, the internal friction angles

and cohesion both increased with time (between 7 and 28 days). Their friction

angles at 28 days were in the range of a very dense granular soil, and the mixtures

behaved like dense sand, with lower residual strengths than ultimate strengths. For

mixture I-4, the strength development was manifested mainly as an increase in

internal friction angle, whereas for mixture I-5, an increase in cohesion was the

dominant factor. These mixtures (I-4 and I-5) had high air contents and exhibited

behavior similar to loose or uncompacted sands. It was interesting to note that for

mixture I-6, the friction angle decreased and cohesion greatly increased with time.
182
Figure 5.26 illustrates some of the more profound differences in CLSM behavior,

contrasting the stress-strain behavior of mixtures I-1 and I-4 samples at 28 days.

Table 5.24: Results of triaxial compression tests


Mixture 7 days 28 day
Friction angle φ′ Cohesion c′ Friction angle φ′ Cohesion c′
(degree) (kPa) (degree) (kPa)
I-1 36.1 31.8 42.8 40.1
I-2 36.1 96.1 38.7 174.7
I-3 39.4 251.7 47.9 346.2
I-4 22.0 43.9 23.9 43.1
I-5 19.5 89.8 18.5 130.0
I-6 37.3 44.4 33.9 93.4

600

500

I-4
400
I-1

300

200

100

0
0 2 4 6 8 10
Strain (%)

Figure 5.26: Stress-strain curves of mixture I-1 and I-4 at 28 days (triaxial
compression testing with an effective confining pressure of 69
kPa)

183
5.3.3 Water Permeability

The water permeability (or hydraulic conductivity) for six CLSM mixtures

(I-series) are shown in Table 5.25. These permeability values (measured after 28

days of moist-curing) were in the range of silts. The results show that water-

cement ratio affected the coefficient of permeability greatly. Generally, the lower

the w-c ratio, the lower the permeability. Interestingly, the high air content of

mixture I-5 did not increase the permeability significantly, indicating that the

entrained air bubbles were not well connected. The testing performed on these six

mixtures indicates that CLSM can be easily assessed for water permeability using

equipment commonly used to characterize soils. Additional research on water

permeability in which the effects of freezing-and-thawing damage on permeability

were also assessed.

Table 5.25: Water permeability (hydraulic conductivity) of selected CLSM


mixtures

Mixture Permeability Water-cement ratio


(cm/sec)
I-1 2.46×10-4 9.4
I-2 5.33×10-5 5.5
I-3 1.45×10-5 2.8
I-4 4.20×10-4 3.5
I-5 6.75×10-4 2.9
I-6 2.89×10-5 7.0

The water permeability of D-series mixtures was measured after twelve

freezing-and-thawing cycles in accordance with the modified ASTM D 560. The

184
results are shown in Table 5.26 and compared with permeability values of CLSM

specimens that were not exposed to freezing-and-thawing cycles.

Table 5.26: Test results of water permeability after freezing-and-thawing


cycles
Mixture # D-1 D-2 D-3 D-4 D-5 D-6 D-7 D-8 D-9 D-10 D-11
28-day
compressive
0.13 1.02 0.79 1.47 0.11 0.12 0.20 0.38 0.25 0.53 0.34
strength
(MPa)
Permeability
before F-T
7.38 0.35 0.21 13.02 6.60 3.87 1.63 1.60 3.08 3.63 8.06
cycles (× 10-2
mm/sec)
Permeability
after 12 F-T
14.21 1.90 1.61 8.94 0.08 0.94 1.72 0.42 0.73 0.36 5.94
cycles (× 10-2
mm/sec)

Mixtures that would likely suffer freezing-and-thawing damage were

selected to measure possible changes in permeability (before and after testing, as

shown in Table 5.26). However, the effects of freezing-and-thawing damage on

water permeability were somewhat inconclusive, with some mixtures showing

increased permeability and others showing decreased permeability. This was most

likely due to the test set-up, which was designed to keep the samples intact, thus

allowing for subsequent permeability testing. However, by keeping the samples

intact (and confined), it may not have allowed for an accurate estimate of in-situ

permeability. Because the samples were confined, the expansion due to freezing

and thawing may have actually compacted the samples, resulting in an apparent

185
reduction in permeability. More work is needed to elucidate the effects of freezing-

and-thawing damage on permeability.

5.3.4 Drying Shrinkage

As stated in Chapter 4, there are no standard methods to evaluate the drying

shrinkage of CLSM, and only limited emphasis was placed in this project on the

topic. A method developed in Germany for flooring applications was used in this

study, with the results shown in Table 5.27. The temperature during the testing

period was 20 o C, and the relative humidity was approximately 60 to 65%. For

mixtures without air entrainment, most of the shrinkage apparently occurred during

the first day. Much of this may be attributable to early bleeding and subsidence,

which may result in a net volume decrease of the solid mixture. This method is

intended to assess concrete mixtures, as opposed to CLSM, and no attempts were

made to modify or optimize the method for CLSM. It may be possible that CLSM

does not exhibit enough strength (or stiffness) to effectively cause meaningful

movement of the end anchor, thereby limiting its subsequent effectiveness in

measuring shrinkage strains. More research is needed to examine drying shrinkage

of CLSM and to assess and recommend suitable test methods. Because the topic of

drying shrinkage was not identified as a critical issue for this study, it is not

possible to offer firm recommendations on its relevance or on appropriate tests.

186
Table 5.27: Drying shrinkage of selected CLSM mixtures
-6
Shrinkage strain (x 10 )
Time Mix #4 Mix #24 Mix #23 Mix #6 Mix #22r Mix #26
1 day 2260 2830 80 1440 90 10
2 days 2280 2850 80 1450 90 30
3 days 2280 2860 80 1450 90 50
4 days 2280 2860 80 1450 100 70
5 days 2300 2860 80 1460 100 100
6 days 2310 2880 80 1480 110 130
7 days 2330 2880 80 1500 120 160
2 weeks 2390 2930 160 1540 150 200
3 weeks 2410 2960 180 1590 150 190
4 weeks 2410 2980 180 1529 160 210
5 weeks 2410 2980 180 1600 160 220
6 weeks 2420 2960 180 1610 160 220
7 weeks 2410 2960 180 1600

5.3.5 California Bearing Ratio

Based on a prioritization of the testing program in Chapter 4, only limited

testing was performed on assessing CBR values of CLSM. Standard crushed rock

has a CBR value of 100. The results of testing performed on six CLSM mixtures

are shown in Table 5.28. Most of the mixtures exhibited quite high CBR values,

with the exception of some fly ash mixtures and high-air mixtures. Based on the

values obtained, it can be concluded that the specific mixtures investigated would

function as suitable base or subbase material. More importantly, the results and

experience confirm that it is feasible to determine CBR values for CLSM using

equipment commonly used to evaluate soils in typical testing laboratories.

Table 5. 28: CBR values for six selected CLSM mixtures


Mixture 4 6 23 24 22r 26
CBR (%) 216 176 20 62 115 150

187
5.3.6 Resilient Modulus

The resilient moduli of the six selected CLSM mixtures were measured and

the results are shown in Table 5.29. These values were magnitude higher than

typical soils. Based on the values obtained, it can be concluded that the specific

mixtures investigated would function as suitable base or subbase material. More

importantly, the results and experience confirm that it is feasible to determine

resilient modulus values for CLSM using equipment commonly used to evaluate

soils in typical testing laboratories.

Table 5.29: Resilient modulus of six selected CLSM mixtures


Mixture #4 Mixture #6
-3.3517 -2.6284
Regression equation Mr=3.00×1010(Sd ) Regression equation Mr=3.00×106(Sd )
Coefficient of 0.9155 Coefficient of 0.8483
determination R2 determination R2
Deviator stress (kPa) Total Mr (GPa) Deviator stress (kPa) Total Mr (GPa)
276 199.45 69 69.28
345 66.23 138 3.01
414 52.52 207 1.94
276 2.06
Mixture #23 Mixture #24
-2.2978 -2.8847
Regression equation Mr=1.46×105 (Sd ) Regression equation M r=1.00×108(Sd )
Coefficient of 0.9155 Coefficient of 0.9492
determination R2 determination R2
Deviator stress (kPa) Total Mr (GPa) Deviator stress (kPa) Total Mr (GPa)
34.5 57.61 138 134.32
69 4.79 207 20.63
103.5 3.49 276 11.19
138 2.36 345 8.47
414 4.74
Mixture #22r Mixture #26
-0.4929 -1.4393
Regression equation Mr=3.06×102 (Sd ) Regression equation Mr=6.57×105 (Sd )
Coefficient of 0.9395 Coefficient of 0.8122
determination R2 determination R2
Deviator stress (kPa) Total Mr (GPa) Deviator stress (kPa) Total Mr (GPa)
207 23.33 414 104.44
276 17.99 483 89.48
345 16.6 690 72.93
414 15.77 828 33.25
483 15.25
552 13.54

188
5.3.7 Excavation Study

This section summarizes the results of tests that are directly or indirectly

related to the excavatability of CLSM. Included are the initial findings from tests

conducted on selected CLSM mixtures (six from the original mixture series) and

the results of subsequent, more comprehensive testing on excavatability and related

indices. The use of the splitting tensile test as a potential index of excavatability

was investigated, and as such, the tensile results are presented in this section.

As previously described in Chapter 4, the excavatability of CLSM was

assessed for six of the original 38 mixtures by casting CLSM into 450-mm x 450-

mm x 300-mm plywood boxes. The early strength or stiffness of CLSM was

assessed using a soil penetrometer, and these values were correlated with the

“walkability” or the time at which an average person can walk on the material with

a slight impression. It was found that the soil penetrometer values in the range of

4.32 kPa to 7.35 kPa correlated with the initial “walkability.” Long-term

excavatability was assessed for the six CLSM mixtures at an age of approximately

nine months using typical hand tools, including a shovel and a pick. Prior to

assessing the excavatability, the “stiffness” of the samples was measured using the

Stiffness Gauge instrument (as described earlier). Compressive strengths of lab-

cured cylinders were also measured at the time of excavation.

As shown in Table 5.30, there was no clear correlation between

compressive strength, excavatability, and stiffness (as measured by the Stiffness

Gauge). For example, the lab-cured compressive strength of mixture #23 was quite

low, but the field-cured excavation box was not excavatable. It has been shown in

189
previous testing that laboratory-cured cylinders may not be accurate indicators of

in-situ strength or stiffness, especially when CLSM is exposed to higher

temperatures in the field (as was the case for these samples). Also, the results

suggest that compressive strength, by itself, may not be a good predictor of

excavatability. Another example of lack of correlation was the fact that mixture

#24 had a higher “stiffness” than mixture #22r, yet it was much easier to excavate.

The findings of this initial study led to more comprehensive research on

excavatability, including the assessment of other test methods and indirect indices,

as described next.

Table 5.30: Results of excavation study


Compressive C.O.V. of Stiffness* C.O.V. of
Mixture # strength strength stiffness Excavatability
(MPa) (%) (MN/m) (%)
24 0.31 4.86 11.62 10.27 Easy
22r 1.01 5.49 9.76 5.40 Difficult
6 0.92 9.66 13.20 6.72 No
4 0.70 5.24 30.72 9.10 No
26 1.61 5.91 34.57 2.09 No
23 0.12 22.31 17.15 7.51 No
* Stiffness was measured using the GeoGauge device

A comprehensive follow-up study to the initial excavatability investigation

was performed. A wide range of CLSM mixtures (C-series) was included in the

investigation, and the following methods or approaches were assessed as possible

indices (direct or indirect):

• Unconfined compressive strength (field-cured cylinders)


• Field penetrometer (field version of ASTM C 403 needle penetrometer)

190
• Dynamic cone penetrometer (DCP)
• CBR (estimated from DCP)
• Stiffness gauge (GeoGauge)
• Removability modulus (RE)
• Splitting tensile strength

The results of field penetrometer and compressive strength testing at


various ages are shown in Table 5.31. In many cases, the mixtures were too strong
to obtain a penetration value. Between 7 and 28 days, all mixtures increased in
compressive strength, but some mixtures (C-1, C-3, and C-9) exhibited reductions
in strength between 28 and 91 days. Some of the excavation box samples also
exhibited reductions in strength or stiffness, perhaps due to the long-term exposure
to the elements. Thus, some samples were not penetrable at earlier ages but
became penetrable at later ages.

Table 5.31: Comparison between field penetrometer values (field-cured


samples) and compressive strength values (field-cured cylinders)
Mix 7 days 28 days 91 day 240 days
# Penetration Strength Penetration Strength Penetration Strength Penetration Strength
(MPa) (MPa) (MPa) (MPa) (MPa) (MPa) (MPa) (MPa)
C-1 5.86 0.44 - 0.62 13.78 0.46 11.9 1.13
C-2 4.27 0.11 - 0.24 - 0.28 - 0.49
C-3 2.27 0.06 2.41 0.13 3.31 0.08 Too soft 0.06
C-4 3.24 0.11 - 0.18 9.92 0.40 14.0 0.45
C-5 - 0.41 - 1.37 - 1.54 - 3.29
C-6 7.65 0.19 - 0.29 - 0.35 - 0.36
C-7 - 0.43 - 1.13 - 1.10 - 1.40
C-8 2.83 0.14 - 0.47 - 0.85 - 1.61
C-9 1.98 0.16 - 0.42 6.06 0.30 6.8 0.20
The dash,“-”, indicates samples that were too strong and penetration values were not btained

Figure 5.27 illustrates the relationship between field penetrometer values

and the compressive strength of field-cured companion cylinders. Although there

191
0.6

0.5
R2 = 0.67
Compressive strength

0.4
(MPa)

0.3

0.2

0.1

0
0 2 4 6 8 10 12 14 16
Needle penetration (ASTM C 403) (MPa)

Figure 5.27: Penetration resistance (ASTM C 403) vs. compressive strength at


different ages (all C-series mixtures)

was a general trend that mixtures with higher compressive strength exhibited higher

penetration values, the correlation was not very strong. A possible explanation for

the lack of fit between the two data sets is that the data include all mixtures from

the C-series, even though some mixture types show different inherent relationships

between strength and penetration resistance. For example, paste mixtures exhibit

relatively high compressive strengths but require substantially less penetration

resistance. Grouping CLSM based on specific materials and mixture proportions

generally yielded a better correlation between strength and penetration resistance.

Thus, it is conceivable that reliable databases linking strength to penetration

resistance can be developed for specific materials and commonly used mixture

192
proportions. However, care should be taken in trying to extrapolate trends

developed for one set of common mixtures (i.e., paste mixtures) to other mixture

types (i.e., high-air mixtures).

Table 5.32 summarizes additional results from the excavatability study,

including DCP values, stiffness values (using GeoGauge), and calculated

Removability Modulus (RE) values. The table also shows the compressive strength

for lab-cured cylinders (at 28 days) and field-cured cylinders, which were cured

adjacent to the excavatability boxes and tested at the time of excavation (240 days).

The densities of the lab-cured and field-cured cylinders were measured before

testing them in compression, and these values were used in RE calculations. The

relative ease of excvation was assessed using a hand shovel.

The DCP index values, which indicate the penetration depth per blow, were

measured for each of the excavation boxes. The minimum value for a recordable

count of blows corresponded to a penetration of at least 25 mm. DCP values were

found to decrease until the specimens ultimately suffered large cracks, as shown in

Figure 5.28. After the large cracks appeared, the DCP values progressively

increased. Thus, the lowest index value was taken for each mixture and used in

Table 5.32 because it represented the most difficult portion to excavate, thus

providing a conservative index.

193
Table 5.32: Results of excavatability study
Mix 28-day Density 240-day RE using RE using DCP CBR Stiffness Relative
# compressive (kg/m3 ) compressive 28-day 240-day index (%) (using ease of
strength strength lab-cured field- (mm) GeoGauge) excavation
(lab-cured) (field-cured) cylinders cured (MN/m) (with
(kPa) (kPa) cylinders shovel)*
C-1 533 1512 1134 0.84 1.22 6.4 37 19.37 3
C-2 807 1946 493 1.51 1.18 6.5 36 40.56 7
C-3 144 1724 63 0.53 0.35 29.0 7 19.03 1
C-4 292 1858 447 0.85 1.05 5.2 46 28.97 6
C-5 1027 2094 3290 1.90 3.39 0.5 100 30.87 9
C-6 332 2252 362 1.20 1.25 5.0 48 18.54 8
C-7 1192 2143 1397 2.12 2.29 1.1 100 11.96 8
C-8 658 2171 1615 1.60 2.51 0.6 100 23.89 10
C-9 417 1660 199 0.85 0.59 10.0 22 23.03 4
*Note: The ease of excavation wa s compared by assigning different values to each mixture, where 1 is easiest and 10 is most
difficult.

The GeoGauge was used to assess the relative “stiffness” of the CLSM

specimens. As CLSM mixtures were quite strong (relative to soil), a thin layer of

wet fine sand was placed on the surface prior to testing, as per the

recommendations of the manufacturer. Three readings were taken for each

mixture. The variations were quite high for the device, with coefficients of

variation as high as 40% for some specimens. A comparison between “stiffness”

values and DCP values is shown in Figure 5.29. There was no clear trend between

stiffness values and DCP, nor was there a clear trend between stiffness values and

actual excavatability (by shovel). In general, the GeoGauge was not found to be an

effective means of assessing the properties of CLSM, both because of poor

reproducibility and inability to predict excavatability.

The correlation between the DCP index and the two sets of calculated RE

values (based on 28-day lab-cured cylinders and 240-day field-cured cylinders) are

shown in Figure 5.30 and 5.31. The RE values calculated using field-cured
194
cylinders, tested at the time of excavation, showed a better correlation with DCP

index than RE values based on 28-day lab-cured cylinders. The RE values based

on 28-day lab-cured cylinders did, however, identify all the mixtures that were not

excavatable (i.e., their RE values were greater than 1). However, using the 28-day

data is quite conservative and may not be suitable for all CLSM mixtures.

Figure 5.28: Cracking of CLSM mixture C-2 during DCP penetration

195
45
40
35
30
25
20
15
10
5
0
0 5 10 15 20 25 30 35

Penetration Index (mm per blow)

Figure 5.29: Correlation between stiffness (from GeoGauge) and DCP index

Using a correlation table provided by the DCP manufacturer, the

corresponding California Bearing Ratio (CBR) values were obtained from the DCP

index. However, according to the manual, the highest possible CBR value (100%)

calculated was reached for some of the CLSM mixtures, making it difficult to use

the value quantitatively. Thus, the DCP index, as previously described, was used

as an index for excavatability, rather than calculated CBR values.

As shown in Figure 5.31, a DCP index of 5 mm per blow correlated well

with an RE value of 1.0. This suggests that the DCP may be an effective, user-

friendly method of assessing excavatability in the field. This approach needs

further investigation in field testing, where excavatability may not be assessed only

196
using hand tools, but also with typical, commercial excavation equipment (i.e.,

backhoe).
35

30

25
y = 5.2798x-2.557
R2 = 0.76
20

15

10

0
0.00 0.50 1.00 1.50 2.00 2.50

RE using 28-day strength

Figure 5.30: Correlation between DCP Index and Removibility Modulus (RE)
calculated using 28-day compressive strength (lab-cured
cylinders)

40
35
30
25 y = 5.4769x-1.8217
20 R 2 = 0.93

15
10
5
0
0.00 1.00 2.00 3.00 4.00

RE using field-cured cylinders

Figure 5.31: Correlation between DCP Index and Removibility Modulus (RE)
calculated using 240-day compressive strength (field-cured
cylinders)

197
Another parameter that potentially might be used as an index for

excavatability is the splitting tensile strength of CLSM. Some preliminary trials

found that tensile strength may, in fact, be more suitable than compressive strength

in assessing excavatability. Although splitting tensile tests were not performed on

the C-series mixtures, some tests were performed on other mixture series. The

results are provided in this section because of the potential of applying tensile data

to excavatability predictions.

The splitting tensile strengths of the E-series mixtures were measured, as

shown in Table 5.33. A split specimen of mixture E-1 is shown in Figure 5.32. For

the E-series CLSM mixtures, the splitting tensile strength to compressive strength

ratio ranged from 9 to 17 percent, which is higher than those typically observed for

conventional concrete. Unlike concrete, this ratio did not substantially decrease

with an increase in compressive strength.

Table 5.33: Compressive and splitting tensile strengths at 7 and 28 days


Mixture 7 days 28 days 91 days
Average C.O.V. fst /fc Average C.O.V. fst /fc Average C.O.V. fst /fc
(kPa) (%) (%) (kPa) (%) (%) (kPa) (%) (%)
E-1 57.0 4.0 18.5 62.5 9.8 11.7 161.5 8.2 19.0
E-2 21.1 5.3 17.5 28.4 19.9 14.5 31.2 19.1 11.9
E-3 14.4 8.2 17.1 20.6 17.5 14.3 27.2 11.7 18.9
E-4 25.0 14.5 19.6 29.5 8.3 10.1 57.6 6.7 12.8
E-5 46.9 12.9 23.6 101.1 23.1 9.8 150.7 23.0 9.5
E-6 34.2 19.3 10.0 57.3 14.7 17.3 61.9 8.3 13.0
E-7 50.5 6.9 11.8 164.2 30.3 13.8 188.5 18.5 11.0
E-8 33.2 16.3 17.3 75.4 8.1 11.5 146.1 1.8 11.8
Note: fst is the splitting tensile strength and fc is the compressive strength.

198
Figure 5.32: A cylinder from mixture E-1, before and after being tested for
splitting tensile strength

Additional splitting tensile tests were performed using the H-series mixtures

to assess the effects of drying on tensile strength and the tensile to compressive

strength ratio. This testing was initiated because drying generally has a more

profound effect on tensile strength than compressive strength, at least ni the case of

conventional concrete. This behavior is generally attributed to the effects of

microcracks. The results, shown in Table 5.34, confirmed that drying had a similar

effect on CLSM, significantly lowering the tensile to compressive strength ratio.

The effects of temperature and drying on the tensile strength of CLSM, especially

as it relates to using tensile strength as an index of excavatability, require additional

research. It is expected that further research will help to classify some of these

effects and to determine the potential for using splitting tensile strength (by itself or

in combination with other parameters) as an indicator of excavatability.

199
Table 5.34: Effects of temperature and drying conditions on splitting tensile
strength of CLSM
Mixture H-1 H-2
Condition Average C.O.V. ft /fc Average C.O.V. ft /fc
(kPa) (%) (%) (kPa) (%) (%)
10 ºC, dry 16.7 23.1 6.5 6.0 20.2 5.1
10 ºC, wet 25.6 19.3 7.9
21 ºC, dry 23.8 8.6 4.9 8.8 39.6 6.1
21 ºC, wet 89.0 27.9 14.1
38 ºC, dry 55.0 18.8 7.3 18.9 11.4 7.2
38 ºC, wet 74.5 6.1 8.1 53.4 15.5 8.4
Mixture H-3 H-4
Condition Average C.O.V. ft /fc Average C.O.V. ft /fc
(kPa) (%) (%) (kPa) (%) (%)
10 ºC, dry 100.4 14.1 7.2 65.5 33.0 7.3
10 ºC, wet 166.3 32.1 12.2 95.8 101.4 5.7
21 ºC, dry 49.2 8.5 5.8 32.2 5.0 7.0
21 ºC, wet 114.5 15.5 12.4 455.9 36.0 12.2
38 ºC, dry 87.8 7.5 8.4 158.6 6.3 7.7
38 ºC, wet 525.4 30.9 13.5 1791.7 12.5 15.6
Mixture H-5 H-6
Condition Average C.O.V. ft /fc Average C.O.V. ft /fc
(kPa) (%) (%) (kPa) (%) (%)
10 ºC, dry 28.8 12.3 6.8 56.1 67.5 8.4
10 ºC, wet 133.0 7.5 9.8 84.5 22.8 9.2
21 ºC, dry 25.6 16.6 7.7 64.1 5.3 17.2
21 ºC, wet 127.8 8.6 8.5 142.8 7.3 15.4
38 ºC, dry 106.8 9.1 8.0 114.1 16.6 14.6
38 ºC, wet 198.2 6.0 7.5 125.3 12.5 12.6

200
5.4 DURABILITY OF CLSM

5.4.1 Freezing-and-Thawing Study

This section discusses the results of two studies on the freezing-and-

thawing resistance of CLSM. The first study involved testing six CLSM mixtures

from the original mixture series using a modified version of ASTM D 560, as

described in Chapter 4. This method was originally developed for the assessment

of soil-cement. The second study used the same method to assess a wider range of

CLSM mixtures (D-series), and also measured the effects of freezing-and-thawing

damage on permeability, which is discussed in Section 5.3.3 in this dissertation.

The results of the initial freezing-and-thawing study are shown in Tables

5.35 and 5.36 for samples cured for 7 days and 28 days, respectively. According to

the “soil-cement laboratory handbook”, the maximum allowable weight loss is 14%

after 12 freezing-and-thawing cycles (for a Group A-1 soil).81 Based on this

criterion, most CLSM mixtures with high air content behaved well, especially after

28 days of moist curing. The possible damage of CLSM due to freezing-and-

thawing may be a concern, especially if it leads to a change in volume of the

material. There has been very little field evidence suggesting that this is a problem

with CLSM, but research was needed, as a minimum, to assess the laboratory

performance of CLSM, and to determine if a suitable method could be

recommended. Based on these preliminary findings, ASTM D 560 (with minor

modifications) was found to be an effective and user-friendly method to assess the

freezing and thawing resistance of CLSM.

201
Table 5.35: Freezing-and-thawing results for selected CLSM mixtures (after 7
days of curing)
Mixture # 4 24 23 6 22r 26
Initial weight (g) 2241 1639 1441 2085 1861 1868
Initial moisture content (%) 10.86 37.78 30.37 14.79 10.38 9.17
1 cycle (% of initial weight retained) 100.1 101.2 102.1 100.1 100.5 100.4
2 cycles (% of initial weight retained) 100.3 102.3 100.6 99.9 100.4 100.2
3 cycles (% of initial weight retained) 100.9 103.3 103.0 97.8 100.2 100.5
4 cycles (% of initial weight retained) 101.2 103.2 100.9 95.4 100.3 100.3
5 cycles (% of initial weight retained) 101.3 103.0 103.3 92.6 100.1 100.4
6 cycles (% of initial weight retained) 101.3 99.2 101.3 91.3 100.1 100.5
7 cycles (% of initial weight retained) 97.7 Damaged Damaged 88.6 100.2 100.5
8 cycles (% of initial weight retained) 92.4 88.7 100.2 100.4
9 cycles (% of initial weight retained) 87.0 88.3 100.2 100.6
10 cycles (% of initial weight retained) 80.2 88.2 100.1 100.4
11 cycles (% of initial weight retained) 76.7 87.7 100.1 100.4
12 cycles (% of initial weight retained) 73.7 87.2 100.0 100.4
Moisture content at end of test (%) 13.35 15.92 11.16 6.97
Mass loss (%) 27.9 13.7 0.7 -2.5

Table 5.36: Freezing-and-thawing results for selected CLSM mixtures (after


28 days of curing)
Mixture # 4 24 23 6 22r 26
Initial weight (g) 2241 1638.5 1441 2084.5 1861 1867.5
Initial moisture content (%) 10.86 37.78 30.37 14.79 10.38 9.17
1 cycle (% of initial weight retained) 100.3 100.5 107.1 98.8 103.1 99.9
2 cycles (% of initial weight retained) 100.3 99.2 105.4 96.6 102.8 99.9
3 cycles (% of initial weight retained) 99.7 98.6 103.7 94.2 102.8 99.8
4 cycles (% of initial weight retained) 96.6 94.4 101.7 91.0 102.8 99.4
5 cycles (% of initial weight retained) 90.3 89.6 100.5 87.5 102.6 99.3
6 cycles (% of initial weight retained) 100.9 Damaged 98.0 84.0 102.6 99.3
7 cycles (% of initial weight retained) 97.3 98.0 82.0 102.7 99.4
8 cycles (% of initial weight retained) - 98.2 81.3 102.6 99.3
9 cycles (% of initial weight retained) 88.7 97.6 79.8 102.5 99.1
10 cycles (% of initial weight retained) 86.1 97.4 77.1 102.5 99.2
11 cycles (% of initial weight retained) 86.1 96.9 75.2 102.3 99.0
12 cycles (% of initial weight retained) 85.1 96.6 66.1 102.3 99.0
Moisture content at end of test (%) 14.16 27.32 16.91 9.87 4.99
Mass loss (%) 17.4 1.1 35.1 -2.7 -2.9

202
The freezing-and-thawing test results of the D-series CLSM mixtures are

shown in Table 5.37. It is interesting to note that D-1 mixture (high-air mixture

with 30 kg/m3 of cement) did not pass the twelve cycles, whereas a similar mixture

with 45 kg/m3 of cement (D-11) did pass the entire 12 cycles, suggesting that both

air-void system and strength contribute to freezing-and-thawing resistance.

Mixture D-10 survived all 12 cycles, most likely due to its higher strength

(contributed from the Class C fly ash). The remaining mixtures (D-2 through D-9)

did not survive all 12 cycles. Mixtures containing foundry sand (D5 to D7) were

damaged after only one or two cycles. Figure 5.33 shows damaged CLSM samples

containing foundry sand. It was observed that specimens demonstrated volume

expansion and cracking due to the formation of ice lenses.

Table 5.37: Freezing-and-thawing test results of D-series mixtures


Mixture # D-1 D-2 D-3 D-4 D-5 D-6 D-7 D-8 D-9 D-10 D-11
Original
1.38 1.77 1.70 1.86 1.53 1.47 1.46 1.50 1.55 1.56 1.33
weight (kg)
Moisture
9.8 10.5 13.6 10.1 30.7 35.5 32.4 27.4 28 25.1 7.3
content (%)
28-day
strength 0.13 1.02 0.79 1.47 0.11 0.12 0.20 0.38 0.25 0.53 0.34
(MPa)
1 cycle (%) 105.2 99.0 100.1 99.9 Damaged 92.8 93.0 98.9 98.1 99.2 103.5
2 cycle (%) 102.3 95.3 97.3 99.8 Damaged Damaged 94.9 94.8 98.1 105.9
3 cycle (%) 84.4 85.4 78.4 99.1 91.5 87.1 98.0 108.1
4 cycle (%) 74.1 75.7 73.2 96.4 88.0 83.6 98.1 110.3
5 cycle (%) 66.4 72.4 71.1 88.4 84.6 79.1 97.7 110.3
6 cycle (%) 59.6 55.4 63.8 61.0 81.4 73.5 97.3 110.3
7 cycle (%) 51.5 54.9 52.1 59.7 76.9 69.2 97.1 108.8
8 cycle (%) 45.8 46.3 54.3 57.9 73.7 62.1 95.8 108.2
9 cycle (%) 37.3 30.1 48.6 43.8 68.1 56.3 94.4 104.6
10 cycle (%) 33.6 Damaged Damaged 40.3 56.1 Damaged 92.2 103.6
11 cycle (%) 25.8 Damaged 49.8 89.4 100.6
12 cycle (%) Damaged Damaged 85.1 97.5
Final
moisture 28.7 21.3
content (%)
Dry weight
11.0 11.9
loss (%)

203
Figure 5.33: Specimens after freezing-and-thawing cycles (CLSM mixtures
containing foundry sand)

5.4.2 Wetting-and-Drying Study

In this section, the effects of wetting-and-drying cycles were discussed.

Specimens from six selected CLSM mixtures were exposed to wetting-and-drying

cycles at the ages of 7 and 28 days respectively. In this study, the samples were

dried at room temperature, approximately 20 °C. Table 5.38 and 5.39 show the

results. In these tables, from 1 to 11 cycles, the weights of dry samples before

being submerged in water were expressed as a percentage of the weights before the

204
wetting-and-drying cycles. After 12 cycles, the weight loss was expressed in terms

of calculated original dry weights.

Table 5.38: Wetting-and-drying results for selected CLSM mixtures (after 7


days of curing)
Mixture # 4 24 23 6 22r 26
Initial weight (g) 2206 1640.5 1416.5 2102.5 1805.1 1868.9
Initial moisture content (%) 10.9 37.8 30.4 14.8 10.4 9.2
1 cycle (% of initial weight retained) 98.4 98.1 95.9 97.0 97.8 95.8
2 cycles (% of initial weight retained) 95.0 96.5 96.4 95.1 95.6 98.8
3 cycles (% of initial weight retained) 98.0 96.5 94.8 94.7 96.9 98.2
4 cycles (% of initial weight retained) 98.3 96.4 93.9 94.2 94.2 98.8
5 cycles (% of initial weight retained) 97.5 95.0 93.0 84.4 98.3 92.9
6 cycles (% of initial weight retained) 97.2 93.6 92.3 95.4 92.0 97.5
7 cycles (% of initial weight retained) 97.2 93.1 91.7 92.9 97.5 96.8
8 cycles (% of initial weight retained) 96.3 92.9 90.7 96.6 95.8 97.1
9 cycles (% of initial weight retained) 95.3 92.6 103.4 87.1 97.4 97.2
10 cycles (% of initial weight retained) 95.7 94.3 87.8 94.9 96.8 96.3
11 cycles (% of initial weight retained) 95.4 91.2 92.0 94.6 94.2 94.7
12 cycles (Mass loss, %) 0.6 1.5 11.8 2.0 -0.6 -0.3
Moisture content at end of test (%) 0.099 0.359 0.385 0.117 0.067 0.059

As shown in Table 5.38, CLSM mixtures, with the exception of mixture

#23, cured for 7 and 28 days, performed satisfactorily throughout the wetting-and-

drying cycles. Mixtures 22r and 26 gained weight when exposed to wetting-and-

drying cycles at 7 days. This was probably due to the continuing hydration of

portland cement. The mass loss of mixture #23 was significantly reduced when it

was tested after 28 days of curing when compared to 7 days of curing. This

suggests that strength is an important factor in resisting wetting-and-drying

damage.

205
Table 5.39: Wetting and drying results for selected CLSM mixtures (after 28
days of curing)
Mixture # 4 24 23 6 22r 26
Initial weight (g) 2261.1 1646.9 1472.3 2134 1790.1 1860.5
Initial moisture content (%) 0.1072 0.3709 0.2875 0.1538 0.09788 0.08642
1 cycle (% of initial weight retained) 96.4 97.4 97.4 99.8 99.0 98.6
2 cycles (% of initial weight retained) 95.4 94.8 99.5 88.6 96.5 96.7
3 cycles (% of initial weight retained) 95.5 95.3 100.0 96.6 97.9 98.0
4 cycles (% of initial weight retained) 95.2 92.3 95.7 95.3 97.5 97.3
5 cycles (% of initial weight retained) 97.6 36.0 99.5 96.4 95.8 95.9
6 cycles (% of initial weight retained) 96.2 75.7 79.2 96.5 95.3 95.7
7 cycles (% of initial weight retained) 94.0 93.4 98.2 95.3 98.3 98.2
8 cycles (% of initial weight retained) 97.6 92.1 96.5 95.6 97.8 97.7
9 cycles (% of initial weight retained) 91.2 93.2 98.5 96.2 97.7 97.8
10 cycles (% of initial weight retained) 96.9 93.1 97.8 96.1 97.2 97.0
11 cycles (% of initial weight retained) 94.8 92.7 94.9 95.8 97.3 97.9
12 cycles (Mass loss, %) 0.2 3.6 5.8 0.0 0.0 0.0
Moisture content at end of test (%) 7.6 33.2 33.8 11.0 7.1 6.1

5.5 CORROSION OF DUCTILE IRON IN CLSM

5.5.1 Introduction

In this section, the results of corrosion of ductile iron under accelerated

laboratory testing are presented. As mentioned in Chapter 4, the pH and resistivity

values of CLSM mixtures were measured. The chloride penetration was illustrated

in Figure 5.34. Uncoupled and coupled ductile-iron specimens were exposed to the

3.5% NaCl solution after 28 days of curing. The mass loss of steel specimens was

measured after being exposed to the NaCl solution. Because in cases where the

depths of embedment are high, the corrosion of steel is controlled by diffusion of

dissolved oxygen in the water entrapped in the soil, the composition of carbon steel

206
has little effect on corrosion resistance in a given soil. Thus, the composition of the

ductile iron is not included here.

5.5.2 pH Values of Extruded Solutions from CLSM Mixtures

The measured pH values are shown in Table 5.40. The pH values of

extruded solutions were found to vary with time, with no general trend. A

saturated calcium hydroxide solution has a pH value of 12.582 . Thus, CLSM

mixtures whose pore solutions had pH values less than 12.5 likely had a lower

concentration of Na+, K+, and OH- dissolved or available in solution. Factors

influencing the pH value of the pore solution include the amount of calcium

hydroxide provided through the hydration of portland cement or dissolution from

Class C fly ash, alkalis from raw materials, the consumption of calcium hydroxide

by pozzolanic reaction, the leaching of calcium hydroxide and alkalis to the

environments, and other mechanisms. The difference in pH values between

mixture #2 and #2r could possibly be attributed to the leaching of calcium

hydroxide. It is important to measure pH values of in situ solution CLSM mixtures

to evaluate possible corrosion tendency.

5.5.3 Resistivity of CLSM Mixtures

The most corrosive soils are those that contain large concentrations of

soluble salts. Because of the presence of salts, such soils have relatively high

electrical conductivities (or low electrical resistivities). As a result, resistivity is

often used to approximate the aggressiveness of a soil. The resistivity values of

207
CLSM mixtures were measured using the approach specified in Chapter 4. The

182-day results are shown in Table 5.40. The moisture content of a soil also affects

its resistivity. In this aspect, the saturated condition of CLSM mixtures, as required

by the test method in this study, represents the worst-case scenario when no

aggressive ions are present.

Table 5.40: pH and resistivity values of CLSM mixtures


Mix # 7-day pH 28-day pH 182-days pH 182-day resistivity (Ω-cm)
1 12.05 11.61 11.4 2593
2 11.84 11.67 9.8 2855
1r 11.88 11.42 10.33 3484
15 11.43 10.9 8.63 2122
3 11.99 11.91 11.65 2698
8 11.8 11.9 11.18 3222
10 11.84 11.74 11.28 1781
9 11.8 11.63 11.1 1441
5 11.97 11.72 11.24. --
12 - 11.86 11.58 4086
4 11.89 11.92 11.12 --
7 11.76 11.68 10.8 3196
3r 12.06 11.92 11.58 3248
4r 12.01 11.68 11.2 8487
24 12.1 11.95 11.53 1582
23 12.12 11.17 11.4 3536
18 12.18 12.05 11.93 4128
14 12.09 12.12 11.4 12049
2r 12.16 11.89 11.9 3882
29 11.96 12.06 11.28 3143
30 11.95 12.06 11.2 1624
17 11.96 12.05 11.01 2310
11 12.01 11.93 11.3 3468
6 11.66 11.31 8.7 10478
16 12.24 11.85 11.32 1834
2130 12.23 10.37 12.53 7387
22 12.21 12.24 12.3 8120
22r 12.04 12.24 12.3 3824
5r 12.03 11.94 12.25 12573
26 11.827 11.64 9.07 3091
16r 12.28 12.24 9.8 2881
13 12 11.91 10.1 5605
25 12 12.03 12.93 1048
19 12.27 12.14 12.64 7072
20 12.14 11.97 12.89 5291
27 12.24 12.29 13.03 3065
20r 12.02 12.24 12.97 8644
28 12.08 12.03 12.65 6287

208
But during testing, chloride ions were present and, as such, it may be more

appropriate to measure the resistivity values of CLSM specimens exposed to the

same solution as that used for mass loss testing. However, resistivity alone is not a

sufficient index of susceptibility of steel embedded in CLSM to corrosion because

other corrosion characteristics, such as pH, passivity of steel, and chloride

concentration are also important.

5.5.4 Chloride Concentration and Profiles

For underground iron pipes, the rate of corrosion can be accelerated by the

presence of aggressive ions. In predicting the susceptibility of pipes placed in

engineered backfill materials, it may be beneficial to establish the quantity of

aggressive ions at the pipe location. This study evaluated five CLSM mixtures for

chloride ingress rate. CLSM cylinders (150 mm x 300 mm) were ponded for 22

weeks with a 3.5% chloride solution. Samples were then tested for chloride

contents as a function of cylinder depth. Figure 5.34 shows results from the testing

program. These results indicate that chloride ions easily migrate through CLSM,

and corrosive conditions can develop at relatively early ages.

209
Chloride Profiles for CLSM Mixtures Ponded with 3.5% Chloride Solution
0

50
Depth from top of cylinder (mm)

Mix 4
Mix 6
100 Mix 24
Mix 22R
Mix 26
150

200

250

300
0 10 20 30 40 50 60 70
3
Total chloride content (kg/m )

Figure 5.34: Chloride profiles for select CLSM mixtures after 22 weeks of
exposure.

Visual observations during the testing program indicated that the ponded

solution was migrating through the CLSM at relatively high rates. It was observed

that approximately 650 mL was permeating through the sample from mixture #23

within approximately twenty-four hours. Because of the relatively fast migration

of solution through the sample, ponding was stopped and mixture #23 was not

evaluated for chloride profiles.

It was observed that all other samples exhibited wetting of the cylinder

sides as a result of ponding the chloride solution on the top of the samples. It is

likely that permeation (i.e. flow as a function of a pressure head) of the chloride

210
solution into the CLSM samples was the predominate mode of chloride ingress.

Diffusion of chlorides (i.e., the ingress of chlorides due to differences in chloride

concentration) was most likely a minor contributor.

5.5.5 Mass Loss of Ductile-Iron Specimens

5.5.5.1 Mass Loss of Uncoupled Coupons

This section discusses the corrosion of uncoupled coupons under laboratory

conditions. The test set-up was explained in Chapter 4. As is the case for most

laboratory corrosion tests, the results should be interpreted with caution. This is

especially true for evaluating corrosion of metals in soils, where there are inherent

variations in location, seasons, and other environmental issues. Local pollution

may also influence the corrosion characteristics of steel pipelines. Thus, the results

from laboratory testing may, at best, be used as a potential index of corrosion.

Actual field tests should be conducted to evaluate corrosion characteristics of

pipelines. The corresponding parameters characterizing soil/CLSM conditions

should also be evaluated to assess corrosion rates and provide guidelines for future

applications. The drawback of field testing is the length of time required for

corrosion initiation. It is important to find a balance between accurate, but time-

consuming, field tests and accelerated, but preliminary, laboratory tests. Thus, the

results presented next should be interpreted cautiously.

The ductile-iron specimens were cleaned and weighed in accordance with

ASTM G 51, “Standard Practice for Preparing, Cleaning, and Evaluating Corrosion

Test Specimens.” The results of measured percent mass losses of samples

211
embedded in CLSM only are shown in Table 5.41. Significant variations in mass

loss existed between different mixtures, repeated mixtures, and samples within one

mixture.

Several factors might have contributed to the variations in test results. First,

the variations were possibly due to the inherent nature of CLSM. Because of the

lack of coarse aggregates, it is difficult to break cement lumps and disperse the

material uniformly. This phenomenon is not obvious when ordinary CLSM

specimens, such as 75-mm x 150-mm cylinders for compression, were tested. For

corrosion coupons, due to their small size and corresponding small surface areas,

slight changes of local CLSM conditions may significantly affect the corrosion

results. Local areas containing higher cement and fly ash contents may show

higher local pH values and suffer less corrosion. Also, the properties of the

transition zone between CLSM and coupons may affect the corrosion behavior.

Unlike the case of submerging coupons in liquid corrosive solution directly, a

transition zone exists and may play an important role. For example, the existence

of entrapped air may affect the corrosion behavior by increasing the available

chloride ions and dissolved oxygen. Another potential factor leading to variations

is the actual measurement itself. Because of the very small values of percentage of

mass loss, a very small error may significantly influence the results. Although steel

specimen surfaces are prepared to meet certain requirements, variations are still

expected. The different surface conditions of steel specimens could also affect the

corrosion rate significantly.

212
Table 5.41: Percent mass loss for samples embedded in CLSM only
Sample ID Specimen No. Average
1 2 3
1 0.05 0.05 0.05
1R 0.04 0.05 0.05 0.04
2 0.02 0.02 0.05 0.03
2R 0.03 0.01 0.02 0.02
3 0.01 0.01 0.02 0.01
3R 0.01 0.02 0.06 0.03
4 0.03 0.02 0.02 0.02
4R 0.01 0.01 0.01 0.01
5 0.03 0.05 0.05 0.04
5R 0.04 0.03 0.02 0.03
6 0.01 0.02 0.03 0.02
7 0.03 0.02 0.02 0.02
8 0.02 0.00 0.01
9 0.03 0.02 0.02 0.03
11 0.05 0.01 0.01 0.02
12 0.03 0.04 0.04
13 0.02 0.02 0.13 0.06
14 0.02 0.04 0.02 0.02
15 0.02 0.03 0.02 0.02
16 0.19 0.22 0.15 0.19
16R 0.10 0.12 0.09 0.10
17 0.28 0.72 0.50
18 0.13 0.37 0.25
19 0.62 0.35 0.48 0.48
29 0.03 0.03 0.02 0.02
30 0.03 0.01 0.31 0.12
22 0.37 1.02 0.17 0.52
22R 0.08 0.09 0.11 0.09
23 2.10 5.80 4.60 4.17
24 0.04 1.30 0.67
25 0.03 0.03 0.94 0.33
26 0.11 0.07 0.09 0.09
27 0.06 0.12 0.09
28
20 0.03 0.17 0.03 0.08
21 5.85 1.95 2.85 3.55

213
Because of the variations within a given mixture, the results from a repeated

mixture may be quite different or close to the original mixture. This is because

means from two different populations are compared, which are random variables.

As such, the statistical approach provided below is utilized:83

Step 1: Formulate hypotheses. If the means of two populations are

denoted as µ1 and µ2 , the null hypothesis for a test on two independent means

would be:

H0 : µ1 = µ2

The alternative should be chosen as:

HA: µ1 ≠ µ2

Step 2: Select the appropriate model. For the case of two independent

samples, the hypotheses of step 1 can be tested using the following test statistic:

X1 − X 2
t= (5. 6)
1 1
s ( + ) 0. 5
n1 n 2

in which X1 and X 2 are the means of the samples from populations 1 and 2,

respectively; n1 and n2 are the samples sizes for the samples from populations 1 and

2, respectively; t is the value of a random variable having a t distribution with

degrees of freedom of (n1 +n2 -2); and s is the square root of the pooled variance

given by

(n1 − 1) s12 + ( n 2 − 1) s 22
s2 = (5. 7)
n1 + n2 − 2

214
where s12 and s22 are the variances of the samples from populations 1 and 2,

respectively. This test statistic assumes that the variances of the two populations

are equal, but unknown.

α ). Values of either 1 or 5 percent


Step 3: Select the level of significance (α

are used most frequently.

Step 4: Compute estimate of test statistic. Samples are drawn from the

two populations and the sample means and variances are drawn. In this study,

three samples are drawn per population. The statistic is then calculated to test the

null hypothesis, which assumes the two means are equal.

Step 5: Define the region of rejection. The region of rejection is a

function of the degrees of freedom (n1 + n2 –2), the level of significance, and the

statement of the alternative hypothesis. In this case, the region of rejection is:

t< -tα/2 or t > tα/2

Step 6: Select the appropriate hypothesis. If the sample value lies in the

region of rejection, the null hypothesis should be rejected.

Through this approach, the comparisons between repeated mixtures are

shown in Table 5.42. It is estimated that four out of seven repeated mixtures did

not yield the same corrosion mass losses as the original mixtures. There were two

possible causes for the differences between repeated and control mixtures. One

cause was that the mass loss was very sensitive to small changes of CLSM

ingredient materials or/and mixing conditions, which is different for every mixture.

The other was due to the small sample sizes. As already pointed out, the within-

batch variations were found to be significant. Thus, larger sample sizes (> 3) are

215
required to represent the whole population in corrosion test. Further study is

needed to determine which factor is more dominant or whether both are equally

important.

Table 5.42: Statistical comparison of repeated CLSM mixtures (uncoupled


corrosion)
Population Population Estimated Critical Accepted
# #r X1 X2 S2 statistic (t) value (t), hypothesis
α=0.05
1 1r 0.050 0.047 2.222E-05 1.018 3.182 Null
2 2r 0.030 0.020 2.000E-04 1.591 2.776 Null
3 3r 0.013 0.030 3.667E-04 -1.958 2.776 Null
4 4r 0.023 0.010 1.667E-05 7.348 2.776 Alternative
5 5r 0.043 0.030 1.167E-04 2.777 2.776 Alternative
16 16r 0.187 0.103 7.333E-04 6.924 2.776 Alternative
22 22r 0.520 0.093 9.887E-02 3.053 2.776 Alternative

The mass loss of samples in different CLSM mixtures was found to vary

significantly. Many factors may affect the corrosion behavior of metal in different

CLSM mixtures. These factors include constituent materials, permeability,

chloride ion diffusion coefficients, subsidence, segregation and bleeding, pH of

pore solutions, and resistivity of the CLSM mixtures.

Ions dissolved in water can enter CLSM mixtures by capillary suction,

permeation, and diffusion. In general, samples embedded in CLSM mixtures with

entrained air contents from 15 to 30% showed much higher corrosion rates than

those in mixtures with entrapped air only.

216
Constituent materials of CLSM mixtures may also influence the corrosion

behavior of embedded pipe specimens. Statistical software (SAS) was used to

analyze the significance of potentially important factors. The General Linear

Models (GLM) Procedure was used because of its ability to handle unbalanced

experimental design with repeated measurements. Four types of F-tests were

utilized to evaluate the significance of factors. A p-value was calculated for each

F-statistic and the factor was determined significant if the p-value was less than a

0.05 significance level. This procedure can also cope with cases of missing data.

As mentioned above, the use of fly ash may significantly change the corrosion

behavior of samples in CLSM mixtures. Thus, the analyses were performed

separately for mixtures with and without fly ash. Factors analyzed include cement

content (two levels), fly ash type (three levels), fly ash content (two levels),

aggregate type (three levels), and water-cement ratio (continuous variable). The

results are shown in Tables 5.43 and 5.44.

Results in Table 5.43 indicate that there was no single factor significantly

affecting the corrosion of samples in CLSM mixtures with fly ash. However, the

interaction between cement content and fly ash type did seem to play an important

role. For instance, samples in mixture #11 (60 kg/m3 cement and 360 kg/m3 high-

carbon fly ash) corroded less than those in mixture #1 (30 kg/m3 cement and 180

kg/m3 Class C fly ash). Thus, when choosing CLSM mixture proportions,

precautions should be taken to evaluate combinations of cement content and fly ash

type.

217
For CLSM mixtures with entrained air (without fly ash), air content and

aggregate type were found by one test to significantly affect the corrosion rate of

embedded samples. It is worth noting that air content may influence corrosion

rates. One test in Table 5.44 indicates that the air content could affect corrosion

rate. More study is required to further assess these factors. The use of different

aggregates (concrete sand and bottom ash) may also affect the ingress of chloride

ions by permeability.

Table 5.43: Statistical results of mass-loss test for uncoupled samples (CLSM
mixtures containing fly ash)
Source DF Type I SS Mean Square F Value Pr > F
Cement content 1 0.00021188 0.00021188 0.62 0.4335
Fly ash type 2 0.00109877 0.00054939 1.62 0.2091
Fly ash content 1 0.00045676 0.00045676 1.35 0.2519
Aggregate 2 0.00134288 0.00067144 1.98 0.1497
Water-cement ratio 1 0.00003464 0.00003464 0.10 0.7508
Cement * fly ash type 2 0.00224639 0.00112319 3.31 0.0453
Source DF Type II SS Mean Square F Value Pr > F
Cement content 1 0.00022344 0.00022344 0.66 0.4213
Fly ash type 2 0.00115231 0.00057616 1.70 0.1942
Fly ash content 1 0.00022754 0.00022754 0.67 0.4171
Aggregate 2 0.0097992 0.00048996 1.44 0.2464
Water-cement ratio 1 0.00000012 0.00000012 0.00 0.9852
Cement * fly ash type 2 0.00224639 0.00112319 3.31 0.0453
Source DF Type III SS Mean Square F Value Pr > F
Cement content 1 0.00001570 0.00001570 0.05 0.8306
Fly ash type 2 0.00141477 0.00070739 2.08 0.1358
Fly ash content 1 0.00022754 0.00022754 0.67 0.4171
Aggregate 2 0.0097992 0.00048996 1.44 0.2464
Water-cement ratio 1 0.00000012 0.00000012 0.00 0.9852
Cement * fly ash type 2 0.00224639 0.00112319 3.31 0.0453
Source DF Type IV SS Mean Square F Value Pr > F
Cement content 1 0.00001570 0.00001570 0.05 0.8306
Fly ash type 2 0.00141477 0.00070739 2.08 0.1358
Fly ash content 1 0.00022754 0.00022754 0.67 0.4171
Aggregate 2 0.0097992 0.00048996 1.44 0.2464
Water-cement ratio 1 0.00000012 0.00000012 0.00 0.9852
Cement * fly ash type 2 0.00224639 0.00112319 3.31 0.0453
Note that “*” means interaction.

218
Table 5.44: Statistical results of mass-loss test for uncoupled samples (CLSM
mixtures containing no fly ash)
Source DF Type I SS Mean Square F Value Pr > F
Cement content 1 0.17285714 0.17285714 0.10 0.7584
Air content 1 20.88580011 20.88580011 11.72 0.0024
Aggregate 1 18.73837547 18.73837547 10.52 0.0037
Water-cement ratio 1 0.01619013 0.01619013 0.01 0.9249
Cement * aggregate 1 0.71627884 0.71627884 0.40 0.5326
Source DF Type II SS Mean Square F Value Pr > F
Cement content 1 0.09839648 0.09839648 0.06 0.8164
Air content 6.74734560 6.74734560 3.79 0.0645
Aggregate 1 1.75261123 1.75261123 0.98 0.3321
Water-cement ratio 1 0.60023984 0.60023984 0.34 0.5675
Cement * aggregate 1 0.71627884 0.71627884 0.40 0.5326
Source DF Type III SS Mean Square F Value Pr > F
Cement content 1 0.27679692 0.27679692 0.16 0.6973
Air content 1 6.74734560 6.74734560 3.79 0.0645
Aggregate 1 0.00236750 0.00236750 0.00 0.9713
Water-cement ratio 1 0.60023984 0.60023984 0.34 0.5675
Cement * aggregate 1 0.71627884 0.71627884 0.40 0.5326
Source DF Type IV SS Mean Square F Value Pr > F
Cement content 1 0.27679692 0.27679692 0.16 0.6973
Air content 1 6.74734560 6.74734560 3.79 0.0645
Aggregate 1 0.00236750 0.00236750 0.00 0.9713
Water-cement ratio 1 0.60023984 0.60023984 0.34 0.5675
Cement * aggregate 1 0.71627884 0.71627884 0.40 0.5326

Comparing mixture 22r and 26, 24 and 27 suggests that the addition of an

accelerating admixture did not change the corrosion performance of pipe coupons

in CLSM. Because only two pairs of mixtures were tested, this finding is only

preliminary.

5.5.5.2 Mass Loss of Coupled Coupons

This section provides results of mass loss of coupled coupons, as shown in

Table 5.45. In this testing, a galvanic cell was formed as a result of pipe coupons

219
being embedded in different environments, which were electrically connected to

complete the circuit for corrosion current. A 10-Ω resister was used to obtain

current flow between coupons.

The procedure listed in Section 5.5.5.1 was used to evaluate the

repeatability of CLSM mixtures with regard to corrosion. The results are shown in

Table 5.46. Five of six mixtures were found to be statistically repeatable. The

repeatability of mass loss of coupled samples was much higher than those of

uncoupled specimens. This was probably due to the different corrosion

mechanisms involved with galvanic coupling. The coupon embedded in CLSM

functioned as a cathode and the coupon in sand as anode. The lower variation was

evident for the within-batch and between-batch variations of results provided here.

A t-test indicated that the means of mass loss for coupled samples

embedded in sand were not significantly different for CLSM mixtures with and

without fly ash content. This was probably because current flow resulting from the

differences in potentials of the samples was governing the corrosion rates.

Similar to Section 5.5.5.1, factors possibly affecting sample corrosion were

analyzed separately for CLSM with and without fly ash. The results are shown in

Tables 5.47 and 5.48.

220
Table 5.45: Percent mass loss for coupled samples embedded in sand
Sample ID Specimen No. Average
1 2 3
1R -- -- 1.68 1.68
2 2.42 1.48 1.80 1.90
2R 2.48 1.52 1.80 1.93
3 3.70 -- 5.70 4.70
3R 1.90 1.25 1.55 1.57
4 2.35 1.45 1.85 1.88
4R 3.50 2.20 1.50 2.40
5 1.50 2.65 2.45 2.20
5R 2.65 2.00 3.58 2.74
6 1.75 2.85 4.65 3.08
7 4.60 5.95 1.50 4.02
8 2.80 -- 3.10 2.95
9 1.42 0.18 -- 0.80
11 2.70 7.60 1.60 3.97
12 1.97 1.78 2.35 2.03
13 1.25 1.75 0.10 1.03
14 2.25 1.84 1.60 1.90
15 4.20 3.60 4.00 3.93
16 2.80 8.30 12.00 7.70
16R 3.00 6.25 1.30 3.52
17 3.90 3.50 4.95 4.12
18 2.80 2.95 1.25 2.33
19 5.60 4.70 -- 5.15
29 2.10 1.60 3.44 2.38
30 1.80 1.44 1.61 1.62
22 2.90 2.70 4.25 3.28
22R 1.20 2.35 2.30 1.95
23 2.80 2.60 -- 2.70
24 11.60 2.80 5.30 6.57
25 4.00 4.65 2.00 3.55
26 3.25 2.55 2.55 2.78
27 6.00 2.65 -- 4.33
28 5.50 5.40 0.40 3.77
20 2.90 4.00 5.80 4.23
21 2.30 1.58 1.30 1.73

221
Table 5.46: Statistical comparison of repeated CLSM mixtures on corrosion
of coupled coupons embedded in sand
Population Population Estimated Critical value Accepted
# #r X1 X2 S2 statistic (t) (t), hypothesis
α=0.05
2 2r 1.90 1.93 0.24 -0.084 2.776 Null
3 3r 4.70 1.57 0.74 3.998 3.182 Alternative
4 4r 1.88 2.40 0.62 -0.806 2.776 Null
5 5r 2.20 2.74 0.50 -0.937 2.776 Null
16 16r 7.70 3.52 13.88 1.375 2.776 Null
22 22r 3.28 1.95 0.57 2.169 2.776 Null

For CLSM mixtures with fly ash, all four statistical tests indicated that

water-cement ratio had a significant effect on corrosion rates in coupled cases.

Three out of four tests showed that cement contents and aggregate type were also

important factors.

For CLSM mixtures without fly ash, no single factor or interaction was

found to significantly influence the corrosion rate of coupled specimens. Unlike

the case of uncoupled corrosion, air content did not play a key role here. Only one

test indicated that cement content affects the corrosion behavior. Further study is

needed to identify the important parameters affecting the corrosion process.

222
Table 5.47: Statistical results of mass-loss test for coupled samples (CLSM
mixtures containing fly ash)
Source DF Type I SS Mean Square F Value Pr > F
Cement content 1 1.83194087 1.83194087 0.11 0.7402
Fly ash type 2 22.93176577 11.46588288 0.70 0.5036
Fly ash content 1 23.10524882 23.10524882 1.41 0.2429
Aggregate 2 62.42531783 31.21265891 1.90 0.1633
Water-cement ratio 1 99.80592246 99.80592246 6.08 0.0183
Source DF Type II SS Mean Square F Value Pr > F
Cement content 1 84.9477888 84.9477888 5.17 0.0287
Fly ash type 2 0.6473494 0.3236747 0.02 0.9805
Fly ash content 1 21.3668191 21.3668191 1.30 0.2611
Aggregate 2 162.0849627 81.0424813 4.94 0.0124
Water-cement ratio 1 99.80592246 99.80592246 6.08 0.0183
Source DF Type III SS Mean Square F Value Pr > F
Cement content 1 84.9477888 84.9477888 5.17 0.0287
Fly ash type 2 0.6473494 0.3236747 0.02 0.9805
Fly ash content 1 21.3668191 21.3668191 1.30 0.2611
Aggregate 2 162.0849627 81.0424813 4.94 0.0124
Water-cement ratio 1 99.80592246 99.80592246 6.08 0.0183
Source DF Type IV SS Mean Square F Value Pr > F
Cement content 1 84.9477888 84.9477888 5.17 0.0287
Fly ash type 2 0.6473494 0.3236747 0.02 0.9805
Fly ash content 1 21.3668191 21.3668191 1.30 0.2611
Aggregate 2 162.0849627 81.0424813 4.94 0.0124
Water-cement ratio 1 99.80592246 99.80592246 6.08 0.0183

A comparison between mass loss of samples embedded in sand for Mixture

#29 and #30 showed that higher cement content dramatically increased the

corrosion rate of the anode in this study. There was not enough information to

study the effect of accelerating agent on corrosion rate in coupled cases.

The investigation of the influence of embedding pipe coupons in different

environments, where one environment is CLSM and the other is sand, indicates that

placing pipe in these different environments while coupled, could lead to increased

corrosion.

223
Table 5.48: Statistical results of mass-loss test for coupled samples (CLSM
mixtures without fly ash)

Source DF Type I SS Mean Square F Value Pr > F


Cement content 1 24.52736310 24.52736310 4.38 0.0486
Air 1 6.03657813 6.03657813 1.08 0.3108
Aggregate 1 0.17048459 0.17048459 0.03 0.8631
Water-cement ratio 1 0.72230698 0.72230698 0.13 0.7230
Cement * aggregate 1 1.30641681 1.30641681 0.23 0.6340
W-C ratio * cement 1 2.43267052 2.43267052 0.43 0.5169
Source DF Type II SS Mean Square F Value Pr > F
Cement content 1 5.09210317 5.09210317 0.91 0.3510
Air 1 4.02414115 4.02414115 0.72 0.4060
Aggregate 1 0.45015898 0.45015898 0.08 0.7795
Water-cement ratio 1 0.39793855 0.39793855 0.07 0.7923
Cement * aggregate 1 3.71715967 3.71715967 0.66 0.4242
W-C ratio * cement 1 2.43267052 2.43267052 0.43 0.5169
Source DF Type III SS Mean Square F Value Pr > F
Cement content 1 6.48091742 6.48091742 1.16 0.2941
Air 1 4.02414115 4.02414115 0.72 0.4060
Aggregate 1 2.62623812 2.62623812 0.47 0.5008
Water-cement ratio 1 2.21736283 2.21736283 0.40 0.5358
Cement * aggregate 1 3.71715967 3.71715967 0.66 0.4242
W-C ratio * cement 1 2.43267052 2.43267052 0.43 0.5169
Source DF Type IV SS Mean Square F Value Pr > F
Cement content 1 6.48091742 6.48091742 1.16 0.2941
Air 1 4.02414115 4.02414115 0.72 0.4060
Aggregate 1 2.62623812 2.62623812 0.47 0.5008
Water-cement ratio 1 2.21736283 2.21736283 0.40 0.5358
Cement * aggregate 1 3.71715967 3.71715967 0.66 0.4242
W-C ratio * cement 1 2.43267052 2.43267052 0.43 0.5169

5.6 LEACHING AND ENVIRONMENTAL IMPACT

The potential leaching of constituent materials, specifically heavy metals,

from CLSM was investigated. Interest in leaching and environmental issues

regarding CLSM has increased mainly because of concerns raised by the

Environmental Protection Agency (EPA) regarding the use of fly ash in

construction.26 Emphasis in this study was placed on characterizing the potential

224
toxicity of the by-product materials used in this study, including bottom ash,

foundry sand, and fly ash (Class C, Class F, and high carbon).

The chemical analysis, heavy metal test, and TCLP test on raw materials

were performed by Wyoming Analytical Laboratories, INC., Golden, Colorado. As

previously discussed in the Test Methods and Procedures section of this

dissertation, a systematic approach (shown earlier in Figure 4.13) was taken to first

characterize the chemical composition of the by-product materials used in the

study, then to quantify the total heavy metals present in these materials, and lastly

to measure the heavy metals that actually are able to leach from these materials.

The chemical compositions of the fly ashes and foundry sand were already reported

in Table 3.3 and Table 3.4. This information is particularly important for fly ashes,

where the chemical composition (especially CaO and SiO 2 content) can be used to

estimate pozzolanic and cementitious reactivity, especially as it relates to long-term

strength gain and excavatability.

Table 5.49 summarizes the total concentration of heavy metals present in

the raw by-product materials used in this study. These results represent the total

concentration of the eight key heavy metals. A “rule of thumb” that some

practitioners use is that the concentration of total heavy metals can be up to 20

times the standard TCLP limits.26 According to this guideline, bottom ash, Class C

fly ash, and Class F fly ash exceeded the “rule of thumb” value for arsenic. Thus,

additional testing was performed (using the TCLP method) to determine the actual

amount of heavy metals that are available to leach from these materials. Because

the foundry sand and high-carbon fly ash did not have significant amounts of total

225
heavy metals, the materials were classified as non-toxic, and no subsequent

leaching tests were performed.

Table 5.49: Analysis of heavy metal concentration of extract from raw


materials (mg/L)

Element TCLP 20 times Bottom Foundry Class C Class F High –carbon


limits TCLP limits ash sand fly ash fly ash fly ash
Arsenic 5.0 100 170* 7.7 280* 160* 58
Barium 100.0 2000 2000 240 1300 320 1200
Cadmium 1.0 20 0.23 0.28 1.55 2.1 0.51
Chromium 5.0 100 10 18 87 96 16
Lead 5.0 100 <0.2 18 <0.2 37 <0.2
Mercury 0.2 4 <0.2 <0.2 <0.2 <0.2 <0.2
Selenium 1.0 20 <0.2 <0.2 <0.2 2.4 <0.2
Silver 5.0 100 <0.2 <0.2 <0.2 <0.2 <0.2
* concentration exceeded the “rule of thumb” value of 20X the TCLP limits

The TCLP results for Class C fly ash, Class F fly ash, and bottom ash (raw
materials) are shown in Table 5.50. The concentration of heavy metals that leached
from each material was well below the EPA-recommended TCLP limits, and as
such, the materials were classified as non-toxic and suitable for use in CLSM. If
any of the by-product materials had exhibited significant leaching of heavy metals
(above the TCLP limits), the last step would have been to assess the actual leaching
of heavy metals from CLSM containing the material(s). This systematic approach
can be followed for any material being considered for use in CLSM. Although all
the materials used in this study were deemed non-toxic, it may be possible that
certain materials considered for a given CLSM application may be more of an
environmental concern.

226
Table 5. 50: TCLP test results for Class C fly ash, Class F fly ash, and bottom
ash (mg/L)

Element TCLP limits Bottom ash Class C fly ash Class F fly ash
Arsenic 5.0 0.12 0.074 0.37
Barium 100.0 3.61 0.30 0.17
Cadmium 1.0 0.001 0.004 0.024
Chromium 5.0 0.01 0.29 0.11
Lead 5.0 <0.01 <0.01 <0.01
Mercury 0.2 <0.01 <0.01 0.11
Selenium 1.0 <0.01 0.37 0.02
Silver 5.0 <0.01 <0.01 <0.01

227
Chapter 6: Conclusions and Future Work

6.1 CONCLUSIONS

This study focused on fresh, hardened, and durability characteristics of

CLSM mixtures. A wide range of CLSM mixtures were included in this research

to address the variability of raw materials incorporated in CLSM. Waste

materials, such as fly ash, bottom ash, and foundry sand, were used. Suitable

existing test methods were examined and modified in this study to compensate for

the lack of CLSM test methods. Existing geotechnical test methods, such as

triaxial shear and resilient modulus, were included in the research to assess their

applicatability to CLSM. Based on this comprehensive study, the following

conclusions can be drawn.

Fresh properties of CLSM

1. Source of aggregates i.e. foundry sand, bottom ash, and river sand is

the most significant factor affecting bleeding and water demand of

CLSM mixtures to obtain a target flow;

2. The foundry sand and bottom ash used in this study led to difficulties

in entraining large volumes of air in CLSM;

3. With slight modifications described in Chapter 4, ASTM C 231 is a

good approach to measure the air contents and unit weights of fresh

CLSM mixtures;

228
4. Both the needle (ASTM C 403) and pocket soil penetrometer can be

utilized to characterize the setting and hardening of CLSM mixtures.

Precautions should be taken to interpret the results from the two

methods because the penetration depths and underlying mechanisms

are different;

5. The initial and final setting concepts, used in concrete technology, may

not be appropriate for CLSM.

Hardened properties of CLSM

6. For CLSM mixtures with high air content (without fly ash), the water-

cement ratio is the main factor affecting compressive strength;

7. For CLSM mixtures containing fly ash, a predictive model was

developed to estimate the development of compressive strength. Not

only the water-cement ratio, but also the fly ash type, fly ash content,

aggregate type influences the CLSM strength;

8. Variations in load rate can affect the measured values of compressive

strength of CLSM specimens. A loading rate between 0.042 to 0.16 %

per minute is recommended for compression tests to achieve adequate

accuracy and to accomplish testing in a reasonable time;

9. Testing machines with smaller, more accurate load cells yield more

repeatable compression test results than conventional concrete

compression machines with larger capacities;

229
10. The air drying of CLSM specimens for 4 to 8 hours, as specified by

ASTM D 4832, was found to be unnecessary in this study. Thus, this

waiting period is not recommended as part of standard compression

testing;

11. Leaching of hydration products from CLSM specimens (i.e., during

wet-curing) may lead to variations in strength. Direct contact between

standing water and CLSM should be avoided, especially at early ages.

Dripping water could damage CLSM specimens;

12. Curing temperature and humidity can significantly affect the strength

development of CLSM mixtures with fly ash. The compressive

strength values can be an order of magnitude different with different

curing conditions;

13. The inclusion of Class C fly ash in CLSM mixtures typically yields

higher compressive strengths. If future excavation is specified, the use

of Class C fly ash should be investigated prior to use. Local

experience serves as an excellent tool for mixture proportioning and

obtaining the desired short- and long-term strength;

14. CLSM curing appears to be significantly influenced by surrounding

temperature and humidity rather than drainage conditions. This

provided very critical information to people specifying compressive

strength of CLSM mixtures based on a performance-based

specification. The curing temperature and humidity are so important,

230
that appropriate conditions should be clearly specified for the mixtures

proposed by the contractors;

15. High-strength gypsum is an appropriate alternative capping material to

sulfur. However, the capping procedure may take more than half an

hour per sample;

16. Neoprene pads with hardness durometer less than 50 were found

suitable for testing the compressive strength of CLSM, provided that

the strength reduction (compared to sulfur capping) less than 20% is

accepted as specified in ASTM D 4832;

17. Removability Modulus can be utilized to predict the ease of future

excavation of CLSM mixtures in the field. This approach was found

to be conservative in this study;

18. The dynamic cone penetrometer (DCP) is a viable method to evaluate

the in-situ strength of CLSM mixtures. The DCP index correlates well

with the Removability Modulus (RE) if the strength of field-cured

samples are utilized to calculate RE;

19. No acceptable correlations between compressive strength and needle

penetration resistance (ASTM C 403) were found;

20. Stiffness (measured by Stiffness Gauge) of CLSM mixtures did not

correlate well with penetration index from DCP;

21. Measuring the California Bearing Ratio (CBR) and Resilient Modulus

of CLSM mixtures was found to be feasible;

231
22. The water permeability of CLSM mixtures were found to be in the

range of silts;

23. When testing water permeability according to ASTM D 5084-90, to

check the saturation of samples, the requirement that B > 0.95 can be

waived. Instead, the change of B values (approaching a constant) with

change of confining pressure should be checked;

24. It is feasible to measure the triaxial properties of CLSM mixtures

using conventional geotechnical equipment;

25. Beyond the first day, the drying shrinkage of CLSM mixtures may be

very small (according to the limited testing performed in this study).

Durability Characteristics of CLSM

26. For CLSM specimens cured in the fog room in this study, the pH

values of extruded solutions change with time and mixture

proportions. The pH values ranged from 8.63 to 13.03 in this study;

27. The resistivity of CLSM mixtures after 182 days ranged from 1781 to

12573 Ω-cm for CLSM specimens cured in the fog room;

28. Mass loss of ductile-iron specimens embedded in CLSM in an

uncoupled conditions demonstrated large variations within mixtures

and between mixtures;

29. For the uncoupled conditions, the corrosion rates of specimens in

CLSM mixtures with fly ash might be affected by the interaction

between cement content and fly ash type;

232
30. For the uncoupled conditions, no single factor was found to

significantly influence the corrosion rates of specimens in CLSM

mixtures without fly ash;

31. The mass loss of ductile-iron specimens embedded in sand and in

CLSM in a coupled condition showed less variation than those in an

uncoupled condition;

32. For the coupled conditions, corrosion rates of specimens in sand with

CLSM containing fly ash were affected by the water-cement ratio and

possibly the aggregate source;

33. For the coupled conditions, no factor was found to significantly

influence the corrosion rates of specimens in sand when CLSM

mixtures did not contain fly ash;

34. To survive twelve freezing-and-thawing cycles, CLSM mixtures

required both high air contents and adequate strengths;

35. CLSM mixtures can typically survive wetting-and-drying cycles, even

if cured only for seven days, provided that the samples are allowed to

air dry;

6.2 FUTURE WORK

Although this research focused on many important issues, additional

research is needed to further advance the state-of-the-art regarding CLSM. Some

issues that deserve further exploration include:

233
1. Evaluate the feasibility of ASTM D 6023-96, “Standard Test Method

for Unit Weight, Yield, Cement Content, and Air Content

(Gravimetric) of Controlled Low-Strength Material (CLSM)” ;

2. Optimize the mixture proportions of CLSM containing bottom ash as

aggregate to obtain desired flowability without excessive bleeding;

3. Identify a suitable method to characterize aggregate properties that

may affect the flowability of CLSM;

4. Study the hydration products formed in CLSM mixtures because their

high water-cement ratios significantly exceed normal ranges for

conventional concrete. Also, the ratio of fly ash to portland cement is

much higher than that for concrete.

5. Research is needed to study the hydration of fly ash, especially Class

C fly ash. Focus should be placed on the effects of temperature and

humidity;

6. More research is needed to evaluate the use of CLSM as a pavement

subbase and base;

7. Additional work is required to evaluate the excavatability of CLSM in

field applications and examine the effects of surrounding environments

and local climate conditions;

8. The mechanisms of corrosion of metals in CLSM need to be studied in

more detail. Potential influencing factors should be identified. If

possible, corrosion prediction criteria similar to those for soil should

be established to assist in design and construction;

234
9. More severe wetting-and-drying tests should be performed on CLSM

to evaluate its durability.

235
Appendix A: Referenced Standard Test Methods

ASTM Standard Test Methods:

ASTM C 33-1999, “Standard Specification for Concrete Aggregate”

ASTM C 39-1996, “Standard Test Method for Compressive Strength of


Cylindrical Concrete Specimens”

ASTM C 128-1997, “Standard Test Method for Specific Gravity and Absorption
of Fine Aggregate”

ASTM C 136-1996a, “Standard Test Method for Sieve Analysis of Fine and
Coarse Aggregates”

ASTM C 143-1998, “Standard Test Method for Slump of Hydraulic-Cement


Concrete”

ASTM C 231-1997, “Standard Test Method for Air Concrete of Freshly Mixed
Concrete by the Pressure Method”

ASTM C 311-1998b, “Standard Test Method for Sampling and Testing Fly Ash
or Natural Pozzolans for Use as a Mineral Admixture in Portland-Cement
Concrete”

ASTM C 403-1997, “Standard Test Method for Time of Setting of Concrete


Mixtures by Penetration Resistance”

ASTM C 494-1998a, “Standard Specification for Chemical Admixtures for


Concrete”

ASTM C 496-1996, “Standard Test Method for Splitting Tensile Strength of


Cylindrical Concrete Specimens”

ASTM C 618-1998, “Standard Specification for Coal Fly Ash and Raw or
Calcined Natural Pozzolan for Use as a Mineral Admixture in Concrete”

236
ASTM C 939-1997, “Flow of Grout for Preplaced-Aggregate Concrete (Flow
Cone Method)”

ASTM C 940-1998a, “Standard Test Method for Expansion and Bleeding of


Freshly Mixed Grouts for Pre- Placed Aggregate Concrete in the Laboratory”

ASTM C 1231-1999, “Standard Practice for Use of Unbonded Caps in


Determination of Compressive Strength of Hardened Concrete Cylinders”

ASTM D 559-1996, “Standard Test Methods for Wetting and Drying Compacted
Soil-Cement Mixtures”

ASTM D 560-1996, “Standard Test Methods for Freezing and Thawing of


Compacted Soil-Cement Mixtures”

ASTM D 2435-1996, “Standard Test Method for One-Dimensional Consolidation


Properties of Soil”

ASTM D 4429-1993, “CBR (California Bearing Ratio) of Soils in Place”

ASTM D 4525-1990, “Standard Test Method for Permeability of Rocks by


Flowing Air”

ASTM D 4612-1986, “Standard Practice for Calculating Thermal Diffusivity of


Rocks”

ASTM D 4832-1995, “Preparation and Testing of Controlled Low-Strength


Material (CLSM) Test Cylinders”

ASTM D 5084-1990, “Measurement of Hydraulic Conductivity of Saturated


Porous Materials using a Flexible Wall Permeameter”

ASTM D 5334-1992, “Standard Test Method for Determination of Thermal


Conductivity of Soil and Soft Rock by Thermal Needle Probe Procedure”

ASTM D 5335-1992, “Standard Test Method for Linear Coefficient of Thermal


Expansion of Rock Using Bonded Electrical Resistance Strain Gages”

ASTM D 5971-1996, “Standard Practice for Sampling Freshly Mixed Controlled


Low-Strength Material”

237
ASTM D 6023-1996, “Standard Test Method for Unit Weight, Yield, and Air
Content (Gravimetric) of Controlled Low Strength Material”

ASTM D 6024-1996, “Test Method for Ball Drop on Controlled Low Strength
Material to Determine Suitability for Load Application”

ASTM D 6103-1997, “Standard Test Method for Flow Consistency of Controlled


Low-Strength Material”

ASTM G 1-1995, “Standard Practice for Preparing, Cleaning, and Evaluating


Corrosion Test Specimens”

ASTM G 51-1999, “Standard Test Method for Measuring pH of Soil for Use in
Corrosion Testing”

ASTM G 57-1995, “Standard Test Method for Field Measurement of Soil


Resistivity Using the Wenner Four-Electrode Method”

ASTM G 59-1997, “Standard Practice for Conducting Potentiodynamic


Polarization Resistance Measurements”

ASTM G 109-1999, “Standard Test Method for Determining the Effects of


Chemical Admixtures on the Corrosion of Embedded Steel in Concrete Exposed
to Chloride Environments”

AASHTO Standard Test Methods:

AASHTO T 22-1997, “Compressive Strength of Cylindrical Concrete


Specimens”

AASHTO T 24-1997 “Obtaining and Testing Drilled Cores and Sawed Beams of
Concrete”

AASHTO T 106-2000, “Compressive Strength of Hydraulic Cement Mortar”

AASHTO T 121-1997, “Unit Weight, Yield, and Air Content (Gravimetric) of


Concrete”

AASHTO T 160-1997, “Shrinkage of Portland Cement Concrete”

238
AASHTO T 161-2000, “Standard Test Method for Resistance of Concrete to
Freezing and Thawing”

AASHTO T 193-1999, “Standard Method of Test for The California Bearing


Ratio”

AASHTO T 197-2000, “Time of Setting of Concrete Mixtures by Penetration


Resistance”

AASHTO T 222-1990, “Nonrepetitive Static Plate Load Tests of Soils and


Flexible Pavement Components for Use in Evaluation and Design of Airport and
Highway Pavements”

AASHTO T 292-1991, “Standard Method of Test for Resilient Modulus of


Subgrade Soils and Untreated Base/Subbase Materials”

AASHTO T 276-1997, “Developing Early Age Compression Test Values and


Projecting Later Age Strengths"

AASHTO T 277-1996, “Rapid Determination of the Chloride Permeability of


Concrete”

Corp of Engineers Test Methods:

CRD-C611-1980, “Method of Test for Flow of Grout Mixtures (flow cone


methods)”

239
Appendix B: Setting/Hardening Test Results

Table B-1: Needle penetration (NP) and soil penetrometer (SP) results

Mix # 1 Mix # 2 Mix # 1r Mix # 9


Time NP SP Time NP SP Time NP SP Time NP SP
(hrs) (MPA) (MPA) (hrs) (MPA) (MPA) (hrs) (MPA) (MPA) (hrs) (MPA) (MPA)
1.5 0.00 2 2 0.00 3.33 0.01
2 0.01 3 3 0.00 4.25 0.11 0.01
2.5 0.01 4 4 0.50 0.02 6.17 0.22 0.02
3 0.03 5 5 1.10 0.04 7 0.25 0.02
3.5 0.07 6.17 0.58 0.02 6 1.76 0.08 8 0.28 0.02
4 1.21 0.08 6.67 0.76 0.01 7 2.10 0.12 22.33 0.74 0.12
4.5 1.43 0.37 7.67 - 0.03 8 2.43 0.15 26.17 0.83 0.12
5 1.43 0.17 8.17 0.72 0.04 9 2.32 0.17
5.5 2.07 0.19 20 4.91 0.07 24 2.89 0.27
6 2.32 0.25 24 3.20
8.25 2.32 0.27
24 4.69 0.46
Mix # 15 Mix # 4 Mix # 5 Mix # 3
Time NP SP Time NP SP Time NP SP Time NP SP
(hrs) (MPA) (MPA) (hrs) (MPA) (MPA) (hrs) (MPA) (MPA) (hrs) (MPA) (MPA)
2 0.00 0.44 0.05 3 1.24 0.01 4 0.44 0.01
3 0.01 4.25 0.58 0.05 4.67 1.32 0.02 5 0.74 0.04
4 0.17 0.02 5.58 0.69 0.09 5.83 1.49 0.02 6 0.61 0.04
5 0.11 0.03 6.33 0.77 0.08 7.67 1.76 0.03 7.5 1.24 0.04
6 0.28 0.04 7 0.94 0.10 22.33 2.70 0.20 8.5 1.41 0.05
7 0.33 0.05 9 1.49 0.12 24.33 2.78 0.26 11 2.37 0.13
8 0.33 0.06 24 3.86 - 12 2.78 0.15
9 0.33 0.06 24 8.41
10 0.39 0.07
24 0.39 0.09
Mix # 8 Mix # 10 Mix # 12 Mix #7
Time NP SP Time NP SP Time NP SP Time NP SP
(hrs) (MPA) (MPA) (hrs) (MPA) (MPA) (hrs) (MPA) (MPA) (hrs) (MPA) (MPA)
3 0.06 3 1.38 0.08 3 1.02 0.03 3.5 0.03 0.00
4.22 0.08 0.09 4 1.46 0.07 4.33 0.96 0.03 4.33 0.08 0.01
5.17 0.11 0.01 5.33 1.65 0.05 5.5 1.10 0.07 6.73 0.22 0.02
6.5 0.14 0.02 6.83 1.79 0.11 6.67 1.43 0.09 8 0.25 0.02
7.5 0.22 0.02 8.5 2.45 0.15 8 1.93 0.10 9.33 0.28 0.02
8.5 0.28 0.03 24 3.03 0.30 24 8.06 - 24 0.63 0.1
24.67 0.94 0.13

240
Table B-1: Needle penetrAtion (NP) and soil penetrometer (SP) results
(cont’d)

Mix #3r Mix # 4r Mix # 24 Mix # 17


Time NP SP Time NP SP Time NP SP Time NP SP
(hrs) (MPA) (MPA) (hrs) (MPA) (MPA) (hrs) (MPA) (MPA) (hrs) (MPA) (MPA)
3.00 0.55 0.01 3.67 0.86 0.02 3.40 0.08 0.01 3.00 0.28 0.02
5.00 1.32 0.03 5.00 0.77 0.03 4.67 0.33 0.01 4.33 0.36 0.05
7.00 1.43 0.05 6.00 1.41 0.04 5.50 0.30 0.01 5.33 0.39 0.05
8.50 1.49 0.07 8.25 1.86 0.08 6.50 0.50 0.02 9.50 0.66 0.08
10.50 2.43 0.11 9.33 1.88 0.11 8.00 0.61 0.02 23.00 0.83 0.16
24.00 8.96 - 26.00 5.41 0.42 24.00 2.87 0.18 26.50 0.91 0.21
Mix # 23 Mix # 18 Mix # 14 Mix # 11
Time NP SP Time NP SP Time NP SP Time NP SP
(hrs) (MPA) (MPA) (hrs) (MPA) (MPA) (hrs) (MPA) (MPA) (hrs) (MPA) (MPA)
5.00 0.08 0.00 4.50 0.08 0.01 2.50 0.30 0.02 3.00 0.41 0.02
6.50 0.14 0.01 5.67 0.22 0.01 3.50 0.50 0.03 4.17 0.58 0.02
8.50 0.33 0.01 6.67 0.30 0.02 4.50 0.66 0.05 5.25 1.10 0.03
22.00 0.61 0.28 8.00 0.47 0.03 5.50 0.99 0.06 6.17 1.43 0.08
24.00 0.88 0.05 14.00 1.38 0.11 8.50 3.03 0.19 7.67 2.07 0.08
28.00 0.77 0.23 31.00 1.60 0.16 26.00 14.76 - 24.00 5.02 0.37
Mix # 2r Mix # 29 Mix # 30 Mix # 6
Time NP SP Time NP SP Time NP SP Time NP SP
(hrs) (MPA) (MPA) (hrs) (MPA) (MPA) (hrs) (MPA) (MPA) (hrs) (MPA) (MPA)
3.00 0.17 0.01 2.50 0.25 0.02 3.00 0.22 0.02 3.00 0.00 0.00
4.33 0.41 0.02 3.50 0.44 0.05 4.00 0.36 0.03 4.33 0.06 0.01
5.67 1.02 0.10 4.00 0.61 0.08 5.00 0.44 0.04 6.33 0.55 0.02
7.17 3.39 0.30 4.50 0.72 0.08 6.00 0.47 0.05 8.33 0.58 0.03
8.50 5.24 - 7.00 1.27 0.17 9.67 0.74 0.09 9.83 0.80 0.03
22.67 14.76 24.00 2.98 0.34 24.00 1.19 0.12 12.50 1.16 0.05
24.17 2.95 0.20

241
Table B-1: Needle penetration (NP) and soil penetrometer (SP) results
(cont’d)

Mix # 16 Mix # 21 Mix # 22 Mix # 22r


Time NP SP Time NP SP Time NP SP Time NP SP
(hrs) (MPA) (MPA) (hrs) (MPA) (MPA) (hrs) (MPA) (MPA) (hrs) (MPA) (MPA)
3.5 0.39 0.03 3 3.66 4.83 0.30 0.01
4.67 0.30 0.03 4.33 0 0.00 5 0.25 0.01 6.5 0.72 0.05
7.33 0.50 0.03 5.33 0.03 0.01 6 0.39 0.01 7.67 0.94 0.10
9.67 0.80 0.04 8.33 0.22 0.01 7 0.58 0.03 9.3 1.49 0.15
10.17 1.05 0.06 11 0.39 0.02 9 0.91 0.03 15 1.93 0.21
24.25 1.10 0.08 21.5 0.61 0.04 11 1.05 0.06 24.56 2.34 0.28
26.83 0.61 0.04 25.33 1.21 0.17
Mix # 5r Mix # 26 Mix # 16r Mix # 13
Time NP SP Time NP SP Time NP SP Time NP SP
(hrs) (MPA) (MPA) (hrs) (MPA) (MPA) (hrs) (MPA) (MPA) (hrs) (MPA) (MPA)
3 0.83 0.03 4 0.41 0.02 2.5 0 0 4.5 0.83 0.12
4 1.02 0.08 5 0.44 0.03 3.75 0.11 0.01 6 1.05 0.15
5 1.05 0.14 6.17 0.91 0.05 5.25 0.11 0.02 7.6 1.10 0.17
7 1.71 0.15 8.17 1.16 0.11 7 0.25 0.03 9 1.27 0.20
9 2.15 0.17 9.17 1.21 0.14 9 0.41 0.03 10.33 1.43 0.22
24.33 4.41 11.17 1.31 0.17 18.67 0.66 0.08 21 2.41 0.31
24.17 2.07 0.25 22.21 0.66 0.11 24.5 2.90 0.37
Mix # 25 Mix # 19 Mix # 20 Mix # 27
Time NP SP Time NP SP Time NP SP Time NP SP
(hrs) (MPA) (MPA) (hrs) (MPA) (MPA) (hrs) (MPA) (MPA) (hrs) (MPA) (MPA)
3 3.33 0.47 0.04 3 0.55 0.03 2 0 0.00
4 0.01 4.33 0.47 0.04 5 0.61 0.06 3 0.11 0.01
5.33 0.03 0.01 5.67 0.55 0.04 7.5 0.91 0.14 4 0.25 0.02
7.33 0.19 0.02 7.17 0.77 0.06 9.5 1.08 0.16 5 0.30 0.02
8.83 0.25 0.02 9.33 0.72 0.07 19.67 1.86 0.22 8 0.50 0.03
10.67 0.36 0.03 25 0.97 0.14 24 1.86 0.30 21.67 2.12 0.21
18.33 0.83 0.06 24 2.34 0.23
Mix # 20r Mix # 28
Time NP SP Time NP SP
(hrs) (MPA) (MPA) (hrs) (MPA) (MPA)
3 0.25 0.02 2 0.25 0.03
5 0.41 0.04 3 0.52 0.04
7 0.61 0.05 4 1.05 0.10
10.67 0.69 0.08 5 1.46 0.11
23.5 1.41 0.27 6 1.85 0.11
25.83 1.45 0.31 9.17 - 0.16
11.33 - 0.24
24.33 6.21 0.27

242
Table B-2: Needle penetration (NP), soil penetrometer (SP) and vane shear
tester (VS) Results

Mix E-1 Mix E-2


Time NP SP VS Time NP SP VS
(MPa) (MPa) (MPa) (MPa) (MPa) (MPa)
3.18 0.23 0.07 0.10 0.80 0.14 0.02 0.13
5.22 0.41 0.16 0.18 3.33 0.28 0.12 0.15
7.18 1.45 0.18 0.29 5.97 0.77 0.12 0.15
11.02 2.13 0.35 0.43 9.80 0.83 0.12 0.22
22.40 3.26 0.48 21.18 1.21 0.17 0.25
Mix E-3 Mix E-4
Time NP SP VS Time NP SP VS
(MPa) (MPa) (MPa) (MPa) (MPa) (MPa)
6.22 0.00 0.00 0.04 2.28 0.50 0.02 0.00
8.27 0.14 0.01 0.05 3.87 0.66 0.03 0.05
10.22 0.17 0.01 0.05 5.95 1.08 0.07 0.08
20.71 0.22 0.10 0.11 7.95 1.24 0.17 0.11
24.02 0.33 0.27 0.06 18.28 2.48 0.31 0.16
21.62 2.04 0.27 0.24
Mix E-5 Mix E-6
Time NP SP VS Time NP SP VS
(MPa) (MPa) (MPa) (MPa) (MPa) (MPa)
3.95 0.34 3.25 0.14 0.02
6.59 0.86 0.04 0.04 5.90 0.36 0.01 0.01
9.50 1.71 0.08 0.10 7.73 0.79 0.03 0.02
11.50 1.71 0.11 0.11 10.82 0.83 0.04 0.03
22.08 2.43 0.18 0.16 12.73 1.10 0.07 0.04
26.17 2.70 0.23 0.22 23.35 2.07 0.15 0.08
Mix E-7 Mix E-8
Time NP SP VS Time NP SP VS
(MPa) (MPa) (MPa) (MPa) (MPa) (MPa)
2.98 0 0.02 0.01 2.98 0.72 0.03 0.03
5.45 0.50 0.03 0.04 5.45 1.55 0.12 0.10
7.31 0.75 0.03 0.04 7.31 2.13 0.22 0.10
11.52 1.10 0.08 0.08 11.52 2.18 0.24 0.10
13.49 1.57 0.09 0.09 13.49 2.40 0.24 0.11
22.82 0.73 0.27 0.11 22.82 0.62 0.36 0.21
21:37 2.04 0.27 0.24

243
References

1. American Concrete Institute, Committee 229, Controlled Low-Strength

Materials (CLSM), ACI 229R-94 Report, 1994.

2. Adaska, W. S., “Controlled Low Strength Materials,” Concrete International,

Vol. 19, No. 4, 1997.

3. Funston, J. J., Krell, W. C. and Zimmer, F. V., “Flowable Fly Ash, A New

Cement Stabilized Backfill,” Civil Engineering, ASCE, 1984.

4. Krell, W. C., “Flowable Fly Ash,” Transportation Research Record, No.

1234, 1989.

5. Krell, W. C., “Flowable Fly Ash,” Concrete International, Vol. 15, No. 7,

1989.

6. Larsen, R. L., “Sound Use of CLSMs in the Environment,” Concrete

International, Vol. 15, No. 7, 1993.

7. Larsen, R. L., “Use of Controlled Low Strength Materials in Iowa,” Concrete

International, Vol. 10, No. 7, 1988.

8. ASTM, The Design and Application of Controlled Low-Strength Materials

(Flowable Fill), Howard, A. K. and Hitch, J. L., eds, STP 1331, 1998.

9. Adesanya, S. A., Analysis of Controlled Low Strength Backfill Material for

Utility Trenches, M. E. Thesis, University of Louisville, 1988.

10. Bhat, S. T., Use of Coal Combustion Residues and Foundry Sands in

Flowable Fill, Ph. D. Dissertation, Purdue University, 1996.

244
11. Burns, F., Flowable Fly Ash Backfill for Buried Pipelines, M. S. Thesis,

University of North Carolina at Charlotte, 1990.

12. Hagerty, J. L., Investigation and Testing of Controlled Low Strength Material

for Use in the Restoration of Street Cuts, M. S. Thesis, University of

Louisville, 1989.

13. Lucht, D. A., Thermal Performance of Flowable Fill Mixtures for Horizontal

GSHP System, M. S. Thesis, South Dakota State University, 1995.

14. McElroy, N. H., Testing of Controlled Low Strength Material Used as a

Utility Cut Backfill, M. S. Thesis, University of Louisville, 1993.

15. Mididaddi, V. G., Evaluation of Utility Cut Restoration Using Unshrinkable

Backfill, M. S. Thesis, Arizona State University. 1990.

16. Monson, T., Engineering Properties of Controlled Low-Strength Material

Using Fly Ash/AMD Sludge, M. S. Thesis, West Virginia University, 1997.

17. Nantung, T. E., Design Criteria for Controlled Low Strength Materials, Ph. D.

Dissertation, Purdue University, 1993.

18. Okwuegbu, A., Recycle of Industrial By-Products in Concrete Construction,

Ph. D. Dissertation, Michigan State University, 1995.

19. Wakim, A. B., Strength and Durability Characteristics of Flowable Fill, M. S.

Thesis, University of Tennessee, Knoxville, 1993.

20. Brewer, W. E., “The Design and Construction of Small Span Bridges and

Culverts Using Controlled Low Strength Material (CLSM),” Columbus, OH,

Ohio Department of Transportation, 1992.

245
21. Lundvall, J. F., “Mitigation of Roadway Settlement above Buried Culvert and

Pipes,” FHWA/WA-97/01, Cheyenne, WY, Wyoming Department of

Transportation, 1997.

22. Rojas-Gonzalez, L. F., Knott, D. L. and Newman, F. B., “Demonstration

Applications in the Use of Flowable Backfills for Bridge Abutments,”

Research Projects 90-12, Gai-Consultants, Inc., 1993.

23. AASHTO, “A Look at What Some States Are Doing,” AASHTO Quarterly,

Vol. 64, No. 4, 1985.

24. Hoopes, R. J., “Engineering Properties of Air-Modified Controlled Low

Strength,” The Design and Application of Controlled Low-Strength Materials

(Flowable Fill), ASTM STP 1331, West Conshohocken, PA, 1998.

25. Folliard, K. J.,Trejo, D., Du, L., and Sabol, S. A., “Controlled Low-Strength

Material for Backfill, Utility Bedding, Void Fill, and Bridge Approaches,”

NCHRP 24-12 Interim report, 1999.

26. Folliard, K. J.,Trejo, D., Du, L., and Sabol, S. A., “Controlled Low-Strength

Material for Backfill, Utility Bedding, Void Fill, and Bridge Approaches,”

NCHRP 24-12(1) Interim report, 2001.

27. Ayers, P. H., Charlton, C. B. and Frishette, C. W., “Investigation of Flash

Fill® as a Thermal Backfill Material,” Midwest Power Conference,

Proceedings of the Midwest Power Conf., Vol. 57//V1, 1995.

28. Fox, T. A., “Use of Coarse Aggregate in Controlled Low Strength Materials,”

Transportation Research Record, No. 1234, 1989.

246
29. Bhat, S. T. and Lovell, C. W., “Design of Flowable Fill: Waste Foundry Sand

as a Fine Aggregate,” Transportation Research Record, No. 1546, 1996.

30. Bhat, S. T. and Lovell, C. W., “Mix Design for Flowable Fill,” Transportation

Research Record, No. 1589, 1997.

31. Bhat, S. T. and Lovell, C. W., “Flowable Fill Using Waste Sand: a Substitute

for Compacted or Stabilized Soil,” Testing Soil Mixed with Waste or Recycled

Materials, ASTM STP 1275, West Conshohocken, PA, 1997.

32. Stern, K., “The Use of Spent Foundry Sand in Flowable Fill in Ohio,”

Foundry Management & Technology, Vol. 123, No.9, 1995.

33. Kennedy, D. O. and Line, C. L. 1987. “Environmental and Economical

Aspects of Sand Reclamation System,” EPRI, Vol. 2, Palo Alato, CA

34. Federal Highway Administration (FHWA), Chesner, W. H., “User Guidelines

for Waste and By-Product Materials in Pavement Construction,” Turner-

Fairbank Highway Research Center, 1998.

35. Environmental Protection Agency (EPA), “Back Document for Proposed CPG

III and Draft RMAN III,” EPA Report EPA530-R-98-003, 1998.

36. Crouch, L. K., Gamble, R. Brogdon, J. F. and Tucker, C. J., “Use of High-

Fines Limestone Screenings as Aggregate for Controlled Low-Strength

Material (CLSM),” Testing Soil Mixed with Waste or Recycled Materials,

ASTM STP 1275, West Conshohocken, PA, 1998.

37. Crouch, L. K. and Gamble, R., “Putting More Fines in Flowable Fill,” Rock

Products, Vol. 100, No. 7, 1997.

247
38. Karim, A. K., Salgado, R. and Lovell, C. W., “Compaction of Fly and Bottom

Ash Mixtures,” 51st Purdue Industrial Waste Conference Proceedings, 1996.

39. Naik, T. R., Kraus, R. N., Sturzl, R. F. and Ramme, B. W., “Design and

Testing Controlled Low Strength Materials (CLSM) Using Coal Ash,” Testing

Soil Mixed with Waste or Recycled Materials, ASTM STP 1275, West

Conshohocken, PA, 1998.

40. Ohlheiser, T. R., “Utilization of Recycled Glass as Aggregate in Controlled

Low-Strength Material (CLSM),” Testing Soil Mixed with Waste or Recycled

Materials, ASTM STP 1275, West Conshohocken, PA, 1998.

41. Gandham, S., Seals, R. K. and Foxworthy, P. T., “Phosphogypsum as a

Component of Flowable Fill,” Transportation Research Record, No. 1546,

1996.

42. Federal Highway Administration (FHWA), “Fly Ash Facts for Highway

Engineers,” U. S. Department of Transportation, FHWA-SA-94-081, 1995.

43. McGrath, T. J. and Hoopes, R. J., “Bedding Factors and E Values for Buried

Pipe Installations Backfilled with Air-Modified CLSM,” The Design and

Application of Controlled Low-Strength Materials (Flowable Fill), ASTM

STP 1331, West Conshohocken, PA, 1998.

44. Nmai, C. K., McNeal, F. and Martin, D., “New Foaming Agent for CLSM

Application,” Concrete International, Vol. 19, No. 4, 1997.

45. Janardhanam, R., Burns, F. and Peindl, R. D., “Mix Design for Flowable Fly-

Ash Backfill Material,” Journal of Materials in Civil Engineering, Vol. 4, No.

3,1992.

248
46. Ayers, M. E., Wong, S. Z. and Zaman, M., “Optimization of Flowable Fill

Mix Proportions,” Controlled Low-Strength Materials, ACI SP 150, 1994.

47. Fesenmaier, T., “CLSM for New England Thruway,” Concrete International,

Vol. 10, No. 7,1988.

48. Brewer, W. E., “Durability Factors Affecting CLSM,” Controlled Low-

Strength Materials, ACI SP150, 1994.

49. Mulllarky, J. I., “Long Term Strength Gain of Controlled Low-Strength

Materials,” The Design and Application of Controlled Low-Strength Materials

(Flowable Fill), ASTM STP 1331, West Conshohocken, PA, 1998.

50. Washington Aggregates and Concrete Association, “Suggested Specifications

for Controlled Density Fill,” Seattle, Washington,1992.

51. Dandria, G. G., Frost, D. J., Ashmawy, A. and Patterson, K. R., “Potential

Factors Affecting Flow consistency Test Method for Controlled Low Strength

Materials,” Transportation Research Record, No. 1589, 1997.

52. Gray, D. D., “Filling Abandoned Mines with Fluidized Bed Combustion Ash

Grout,” The Design and Application of Controlled Low-Strength Materials

(Flowable Fill), ASTM STP 1331, West Conshohocken, PA, 1998.

53. Hepworth, H. K., Davidson, J. S. and Hooyman, J. L., “Admixture Enhanced

Controlled Low-Strength Material for Direct Underwater Injection with

Minimal Cross-Contamination,” The Design and Application of Controlled

Low-Strength Materials (Flowable Fill), ASTM STP 1331, West

Conshohocken, PA, 1998

249
54. Janardhanam, R., Burns, F. and Peindl, R. D., “Mix Design for Flowable Fly-

Ash Backfill Material,” Journal of Materials in Civil Engineering, Vol. 4, No.

2, 1992.

55. Smith, A., “Controlled Low-Strength Material,” Concrete Construction, May

1991.

56. McLaren, R. J. and Balsamo, N. J, Fly Ash Design Manual for Road and Site

Applications, Volume 2, Slurried Placement, Electric Power Research

Institute, 1986.

57. Bernard, R. D. and Tansley, R. S., “Laboratory Testing Program for

Development of a Lean Mix Backfill Specification,” PB83-122952,

Department of Housing and Urban Development, Washington, D. C. , 1981.

58. Ayers, M. E., Wong, S. Z. and Zaman, W., “Optimization of Flowable Fill

Mix Proportions,” Controlled Low-Strength Materials, ACI SP-150, 1994.

59. Dolen, T. P., “Properties of Low-Strength Concrete for Meeks Cabin Dam

Modification Project,” The Design and Application of Controlled Low-

Strength Materials (Flowable Fill), ASTM STP 1331, West Conshohocken,

PA, American Society for Testing Materials, 1998.

60. Abelleria, A., Berke, N. S. and Pickering, D. G., “Corrosion Activity of Steel

in Cementitious Controlled Low-Strength Materials vs. That in Soil,” The

Design and Application of Controlled Low-Strength Materials (Flowable

Fill), ASTM STP 1331, West Conshohocken, PA, American Society for

Testing Materials, 1998.

250
61. Ramme, B. W., Naik, T. R. and Kolbeck, H. J., “Use of Fly Ash Slurry for

Underground Facility Construction,” Construction and Building Materials,

Vol. 8, No. 1, pp. 63-67,1994.

62. Landwermeyer, J. S. and Rice, E. K., “Comparing Quick Set and Regular

CLSM,” Concrete International, Vol. 19, No.5, 1997.

63. Pons, F., Landwermeyer, J. S. and Kerns, L., “Development of Engineering

Properties for Regular and Quick-Set Flowable Fill,” The Design and

Application of Controlled Low-Strength Materials (Flowable Fill), ASTM

STP 1331, West Conshohocken, PA, American Society for Testing Materials,

1998.

64. Mullarky, J. I., “Long Term Strength Gain of Controlled Low-Strength

Materials,” The Design and Application of Controlled Low-Strength Materials

(Flowable Fill), ASTM STP 1331, West Conshohocken, PA, American

Society for Testing Materials, 1998.

65. Gress, D., “The Effect of Freeze-Thaw and Frost Heaving on Flowable Fill,”

UNH Civil Engineering #1096-1 for New Hampshire Department of

Transportation, University of New Hampshire, 1996.

66. Brewer, W. E., “Economic Considerations When Using Controlled Low

Strength Material (CLSM-CDF) as Backfill,” Transportation Research

Record, No. 1315, 1991.

67. Buss, W. E., “Iowa Flowable Mortar Saves Bridges and Culverts,”

Transportation Research Record, No. 1234, 1989.

251
68. Fougere, K. A., “Evaluation of Flowable Fill for Use in Highway and Bridge

Applications,” Technical Paper 94-6, Maine Department of Transportation,

Maine, 1994.

69. Hook, W. and Clem D. A., “Innovative Uses of Controlled Low Strength

Material (CLSM) in Colorado,” The Design and Application of Controlled

Low-Strength Materials (Flowable Fill), ASTM STP 1331, West

Conshohocken, PA, American Society for Testing Materials, 1998.

70. Mason, T. F., “Use of Controlled Density Fill to fill Underslab Void,” The

Design and Application of Controlled Low-Strength Materials (Flowable

Fill), ASTM STP 1331, West Conshohocken, PA, American Society for

Testing Materials, 1998.

71. Naik, T. R., Barun R., Ramme, B. W. and Lolbeck, H. J., “Filling Abandoned

Underground Facilities with CLSM Fly Ash Slurry,”Concrete International,

Vol. 12, No. 7, 1990.

72. Walker, M. P., and Ash, J. P., “Flowable Fill Backfill for Use in Sequential

Excavations in Contaminated Soil,” The Design and Application of Controlled

Low-Strength Materials (Flowable Fill), ASTM STP 1331, West

Conshohocken, PA, American Society for Testing Materials, 1998.

73. Ardani, A., “Bridge Approach Settlement,” Report No. CDOH-DTP-R-87-06,

Colorado Department of Highways, 1987.

74. Gardner, M. R., “Developing controlled Low-Strength Materials to Meet

Industry and Construction Needs,” The Design and Application of Controlled

252
Low-Strength Materials (Flowable Fill), ASTM STP 1331, West

Conshohocken, PA, American Society for Testing Materials, 1998.

75. Green, B. H., Staheli, K., Bennett, D. and Walley, D. M., “Fly-Ash-Based

Controlled Low-Strength Material (CLSM) Used for Critical Microtunneling

Application,” The Design and Application of Controlled Low-Strength

Materials (Flowable Fill), ASTM STP 1331, West Conshohocken, PA,

American Society for Testing Materials, 1998.

76. Kawasaki, H., Horiuchi, S., Akatsuka, M. and Sano, S., “Fly-Ash Slurry

Island, I. Theoretical and Experimental Investigation.” Journal of Materials in

Civil Engineering, Vol.4, No. 2. 1992.

77. Snethen, D. R. and Benson, J. M., “Construction of CLSM Approach

Embankment to Minimize the Bump at the End of the Bridge,” The Design

and Application of Controlled Low-Strength Materials (Flowable Fill),

ASTM STP 1331, West Conshohocken, PA, American Society for Testing

Materials, 1998.

78. Swaffar, K. M. and Price, H. R., “Tunnel Saved by Fly Ash,” Civil

Engineering, Vol. 57, No. 9, 1987.

79. Spangler, M. G. and Handy, R. L., Soil Engineering, Fourth Edition, Harper &

Row, New York, 1982.

80. Hegarty, J. R. and Eaton, S. J., “Flowable Fill Promotes Trench Safety and

Supports Drainage Pipe Buried 60 ft (18.3m) under New Runway,” The

Design and Application of Controlled Low-Strength Materials (Flowable

253
Fill), ASTM STP 1331, West Conshohocken, PA, American Society for

Testing Materials, 1998.

81. Portland Cement Association, “Soil-Cement Laboratory Handbook.” Skokie,

Ill, 59 p., 1992.

82. Lea’s Chemistry of Cement and Concrete, Hewlett, P. C., Editor, Fourth

Edition, 1998, P. 526.

83. McCuen, R. H., Statistical Methods for Engineers, Prentice-Hall, 1985.

254
Vita

Lianxiang Du was born in Emeishan City, Sichuan Province, China on

August 25, 1971, the son of Wuyin Du and Yueying Yuan. After completing his

work at Emei Second High School, Emeishan, Sichuan, in 1989, he entered

Hunan University in Changsha, Hunan Province, China, where he was awarded

the Zhao Zhiying Award. In 1993, he obtained the degree of Bachelor of Science.

In September of the same year, he entered the graduate school at Southwest

Jiaotong University, Chengdu, Sichuan Province, China. In June 1996, he earned

his master’s degree in science with the thesis titled, “Experimental Study on

Properties of Slurry Infiltrated Fiber Concrete (SIFCON).” In Fall 1997, he

received the prestigious Bloc Fellowship and entered The University of Delaware

to pursue his Ph. D. In Fall 1999, he transferred to The University of Texas at

Austin, where he worked as a teaching and research assistant.

Permanent address: Metal Materials Factory 3-7, Emeishan, Sichuan, China,

614200

This dissertation was typed by the author.

255

Вам также может понравиться