You are on page 1of 469

Mathanraj Theivakularatnam

Durability Assessment and Permanent


Deformation Characterisation of Lightly
Stabilised Granular Base Materials

Thesis submitted to

School of Engineering and Information Technology

University of New South Wales, Canberra

for the degree of Doctor of Philosophy

August 2015
Abstract

The mechanical properties of lightly stabilised granular base materials are considerably
affected by in-situ environmental conditions such as freeze-thaw (f-t), wet-dry (w-t)
actions and changes in moisture content. Unlike cemented materials, durability studies
related to lightly stabilised granular materials are very limited and their durability
against these environmental variables are unknown. The existing laboratory tests which
are relevant for assessing durability of stabilised soils are not well received by the
pavement engineers as they have testing procedures which is subjective and requires
time that is longer than the acceptable time frame.

The main objectives of this research are to assess the durability of lightly stabilised
granular materials using different laboratory experiments focussing on various
mechanical properties and to determine the suitable test method and testing protocols.
Specimens of two granular base parent materials stabilised with varying binder contents
were subjected to different environmental actions and tested to obtain various
mechanical properties. Laboratory tests such as unconfined compressive strength
(UCS), repeated load triaxial (RLT), flexural beam test and vacuum saturation test
(VST) were included in the testing program. Furthermore, non-destructive tube suction
test (TST) was also added because of its potential acceptability. Along with the
assessment of durability, permanent deformation (PD) characterisation of lightly
stabilised granular base materials was also investigated using RLT and accelerated
pavement model (APM) tests. The shakedown theory was mainly used to classify the
material responses obtained from the RLT and to determine limits classifying different
range of responses. A finite element (FE) analysis was used to back-calculate the
moduli of stabilised base and subgrade materials of the APM test and its PD was
analysed with RLT tests to form a relationship between these two tests and to develop
equations to predict the PD of PM for different stress combinations.

In first set of durability study, two granular materials, Canberra materials (CM) and
Queensland materials (QM), stabilised with 1.5% of cement-flyash (CF) were chosen
for the experiments. Standard UCS cylindrical specimens moulded with different
moisture contents were used to investigate four testing procedures such as UCS after 7-
day, 12 f-t cycles, vacuum saturation and TST. UCS measured after 12 f-t cycles
directly related to f-t durability tests whereas the other three tests were added to the
testing programs due to their acceptability in predicting the durability of the stabilised
materials. F-t durability test showed that the specimens remained integrated after 12 f-t
cycles and showing resistivity against f-t cycles with no visible fissures or cracks on
them. However, comparison with 7-day UCS revealed that the 12 f-t cycles deteriorated
the material’s UCS. Another series of specimens tested for UCS after subjecting them to
vacuum and saturation also lasted the entire testing program without showing any
disintegration. The regression analysis showed that the outcomes from 7-day UCS, f-t
durability test and VST produced better correlation among them and replacing f-t
durability test with either VST or 7-day UCS is viable for evaluating f-t durability of
lightly stabilised granular materials. However, the dielectric values (DVs) measured
using TST did not relate well with the UCS obtained from these tests and further study
is required before TST is acceptable as a potential test to evaluate f-t durability of
lightly stabilised materials.

Despite its limited ability to assess the durability of the lightly stabilised granular
material, TST revealed outcomes that gave more insight about the materials behaviour.
The DVs showed dependency to moulding moisture content and binder content in
which decrease in moisture content resulted in decrease in DVs and increasing binder
content showed reduction in DVs. UCS of lightly stabilised granular material after TST
showed that the DVs are not very well related to UCS, but the trend depicted that the
material with lower DVs tend to show higher UCS. This research showed that the use of
quality porous disk and having smooth top surface of the specimen are essential for
obtaining reliable results from TST.

In the next series of tests, different factors that could influence the laboratory f-t
durability tests were investigated using UCS, repeated load triaxial (RLT) and flexural
beam tests. This is to obtain the best combinations of variables that could be used to
assess the f-t durability of lightly stabilised granular base materials. In the RLT test,
resilient moduli (Mr) and permanent deformation (PD) were determined to evaluate the
resilient and permanent deformation behaviour of specimens subjected to f-t cycles.

ii
Parent granular material, method of curing, curing time (or age), binder content, f-t
testing procedure and number of f-t cycles were selected as variables.

UCS and RLT tests showed that, irrespective of the testing procedure or variables,
lightly stabilised material performed well against destructive f-t actions. However it was
found that there were two factors that needed attention in the selection of a better
laboratory f-t durability test; one is number of days of curing and the other is the
moisture content. The UCS and RLT tests revealed that the curing age has an influence
on the properties obtained after f-t cycles. To avoid strength gain during the f-t cycles, a
better approach is to cure in an accelerated environment. Supplying moisture during the
thawing phase was found to be important to replicate the field condition and to observe
the adverse f-t effects within a reasonable amount of time. Compared with resilient
modulus, PD from RLT test clearly identified the f-t effects when the specimens were
subjected to higher deviatoric stresses.

F-t durability study performed on flexural properties of specimens stabilised with 1.5%
and 3% of CF showed that f-t actions deteriorated the tensile capacity, considerably
higher than other properties, of the material.

Similar to f-t test durability study, the w-d durability study was performed using UCS,
RLT and flexural test. All three tests performed on QM specimens stabilised with 1.5%
CF showed heavy deterioration after few w-d cycles and not a single specimen lasted 12
w-d cycles. However, increasing the CF binder content to 3% produced enough
resistance to w-d actions and UCS after 12 w-d cycles was significantly higher than the
7-day UCS of the same material not subjected to w-d cycles. Due to the high
temperature in the drying phase of a w-d cycle, the strength gain due to accelerated
cementitious and pozzalanic reaction was evident particularly in specimens cured for
less number of days. Therefore, it was possible to notice strength gain instead of loss in
the first few w-d cycles. Furthermore, similar to the outcome of f-t durability test, the
MoR and SSM were found to be more vulnerable to w-d cycles than other properties
with single w-d cycle causing significant reduction in tensile capacity of QM specimens
stabilised with 1.5% of CF.

iii
Characterising the PD behaviour using RLT test was performed on CM stabilised with
0.5% and 1.5% of CF and cured for 7 days. Specimens stabilised with 0.5% of CF were
tested for moisture contents of 8% and 6.5% for four and three different confining
pressures respectively and three moisture contents of 8%, 6.5% and 5% were selected
for specimens stabilised with 1.5% of CF binder at one confining pressure of 50 kPa.
The PD behaviour from each stress sequences were classified based on the accumulated
permanent strain from 3000 to 5000 load cycles into range A, B or C shakedown
behaviour. Based on this classification, shakedown limit calculations were performed
to obtain plastic creep line (PCL) and plastic shakedown line (PSL). The established
shakedown limit equations can be used to predict the material’s performance for given
stress combinations while considering the impact of moisture content.

Furthermore, log and power based permanent deformation prediction models were
developed to relate permanent strain with number of loading cycles. These two
equations were fitted with the experimental data to obtain the regression coefficients
and the log model showed better fit to the experimental data than the power model. The
regression coefficients were further analysed to obtain their dependency on stress values
and an equation was established to predict the PD in terms of bulk stress and confining
stress.

Finally, APM testing was performed on a pavement model constructed in a cylindrical


tank having 600 mm of subgrade and 150 mm of lightly stabilised base. The model
pavement was instrumented with linear variable differential transformers (LVDTs),
moisture probes (MPs) and earth pressure cells (EPCs) for measuring vertical
deformations, moisture and pressures respectively. The measurement from MPs showed
that soil-specific calibration is essential to improve the accuracy of the measurements
and MP 306 moisture sensors can be used for continuous moisture measurement. The
obtained deformations data were used to back calculate the moduli of both subgrade and
base using a FE model developed in ABAQUS. Vertical and horizontal pressure
distributions of experimental and finite element model showed that EPCs can be used to
measure pressures to a reasonable accuracy.

The permanent deformation (PD) measurement showed that the rut depth measured in
the subgrade was found to be approximately three times that of the lightly stabilised

iv
granular base of the pavement. The study also emphasizes the need to characterise the
deformation behaviour for every stress sequence to avoid continuing the test for higher
number of load cycles which would reveal very little about the material’s behaviour.
Finally, using RLT test data and FE analysis, a model in terms of bulk-stress was
developed for predicting the accumulated permanent deformation of pavements for
different combinations of stresses.

v
Keywords

Stabilisation

Cementitious binders

Lightly stabilised materials

Characterisation

Stiffness modulus

Durability

Freeze-thaw

Wet-dry

Fatigue life

Permanent deformation

Back-calculation

vi
Acknowledgement

I would like express my sincere gratitude to my supervisor Dr. C. T. Gnanendran for his
invaluable advice, guidance throughout my candidature. I’m grateful for his advice,
suggestions, support and encouragement that helped me to produce this thesis. I am
very pleased to have him as my supervisor and feel fortunate to work with him on my
project. I also would like to extend my gratitude to my co-supervisor Dr. Robert Lo for
his support and mentoring.

I would like to extend my gratitude to the technical officers Mr. Jim Baxter, Mr.
Matthew Barrett and Mr. David Sharp for helping me to carry out my experiments. I
would also like thank mechanical workshop technicians for providing support for the
experimental setups.

I would like extend my thanks to Ms. Denise Russell for her time in editing the
manuscript of this thesis.

I am grateful to work with few undergraduate students Terry, Daniel, Josh and Ben and
would like to thank for their help in the experimentation. I wish to thank my colleagues
Dr. Dalim Kumar Paul, Dr. Asafuddoula Asaf, Dr. Ohiduzzaman, Jahid Iftekhar Alam
Dr. Shakhaout Hossain and Dr. Ariful Islam for their help related to my thesis.

I would like to thank all of my friends and well-wishers for their support, interaction
and encouragement during the course of study. This bunch of friends include: Dr.
Sobers Xavior,Arunkumar Krishnamoorthi, Dr. Ramkumar Rathakrishanan, Dr.
Shahidul Islam, Dr. Anup Chakrabortty,, Dr. Nurul Islam, , Dr. Sayem Uddin, Dr.
Khairul Alam and Hu Chen.

Many thanks also go to my family and friends residing different parts of the world their
love, support, encouragement and sacrifice throughout the course of this study.

I would like to acknowledge the financial support provided by the University of New
South Wales, Canberra for supporting my tuition fees, living allowance and other
expenses related to my research.

vii
Dedication

To My Family

viii
ix
x
List of Publications

Journals

 Paul D. K., Theivakularatnam M., and Gnanendran C. T. 2015. “Damage


Study of a Lightly Stabilised Granular Material Using Flexural Testing”. Indian
Geotechnical Journal, Volume 45, Issue 4, pp. 441-448,
http://dx.doi.org/10.1007/s40098-015-0158-2

Refereed Conferences

 Theivakularatnam M., and Gnanendran C. T. “Durability of lightly stabilised


granular material subjected to freeze-thaw and wet-dry cycles”. Geotechnical
Special Publication, In Geo-Congress 2014, Atlanta, USA 2014, pp. 1410 –
1419, http://dx.doi.org/10.1061/9780784479087.127

 Theivakularatnam M., and Gnanendran C. T. “Effects of Freeze-thaw on


Lightly Stabilized Material’s Flexural Properties”. Geotechnical Special
Publication, In Geo-Congress 2014, Atlanta, USA 2014, pp. 3465 – 3475,
http://dx.doi.org/10.1061/9780784413272.336

 Theivakularatnam M., and Gnanendran C. T. “Evaluating the Durability of


Lightly Stabilised Granular Base Materials Subjected to Wet-Dry Cycles Using
Different Element Testings” Proceedings of the 15th Pan-American Conference
on Soil Mechanics and Geotechnical Engineering, pp. 1560 – 1567,
http://ebooks.iospress.nl/publication/41687

 Theivakularatnam M., and Gnanendran C. T. “Tube Suction Test to Evaluate


Moisture Susceptibility of Lightly Stabilised Granular Materials”. Proceedings
of the 15th Pan-American Conference on Soil Mechanics and Geotechnical
Engineering, pp. 446-453, http://ebooks.iospress.nl/publication/41546

xi
Contents

Abstract ...........................................................................................................................iii 

Keywords ........................................................................................................................ vi 

Acknowledgement ......................................................................................................... vii 

Dedication .....................................................................................................................viii 

Originality Statement........................................................ Error! Bookmark not defined. 

List of Publications......................................................................................................... xi 

Contents ......................................................................................................................... xii 

List of Figures .............................................................................................................. xxii 

List of Tables ............................................................................................................. xxxv 

List of Abbreviations ............................................................................................... xxxix 

List of Symbols .............................................................................................................. xli 

Chapter 1 Introduction .................................................................................................. 1 

1.1  General Overview............................................................................................... 1 

1.2  Objectives ........................................................................................................... 3 

1.3  Contribution to new knowledge ......................................................................... 4 

1.4  Organisation of Thesis........................................................................................ 5 

Chapter 2 Literature Review ........................................................................................ 8 

2.1  Introduction ........................................................................................................ 8 

2.2  Mechanical –Empirical (M-E) Analysis and Design........................................ 10 

2.3  Enhanced Integrated Climatic Model (EICM) ................................................. 13 

2.4  Parameters Required to Run Enhanced Integrated Climatic Model ................. 14 

2.5  Outputs from Enhanced Integrated Climatic Model (EICM) ........................... 14 

xii
2.6  Pavement Distress Modes and Performance Models ....................................... 16 

2.7  Stabilisation of Base Materials ......................................................................... 19 

2.7.1  Cementitious Stabilisation ........................................................................... 20 

2.7.1.1  General Purpose (GP) and General Blended (GB) Cements ................. 20 

2.7.1.2  Fly Ash .................................................................................................. 21 

2.7.1.3  Slag ........................................................................................................ 22 

2.7.1.4  Lime ....................................................................................................... 23 

2.7.2  Granular Stabilisation .................................................................................. 24 

2.7.3  Bituminous Stabilisation .............................................................................. 25 

2.7.4  Chemical Stabilisation ................................................................................. 25 

2.8  Categorisation of Cementitiously Stabilised Materials .................................... 26 

2.9  Characterisation of (lightly) Stabilised Materials............................................. 27 

2.9.1  Characterisation using Unconfined Compressive Strength Test ................. 28 

2.9.2  Characterisation using Tensile Properties ................................................... 30 

2.9.3  Characterisation using Resilient Modulus ................................................... 31 

2.10  Durability of (Lightly) Stabilised Materials ..................................................... 32 

2.10.1  Laboratory Freeze-Thaw Durability Tests ............................................ 33 

2.10.1.1  Standard Freeze-Thaw Test of Cement-Stabilised Soils ................... 33 

2.10.1.2  Freeze-Thaw Durability Studies using Unconfined Compressive


Strength Test.................................................................................................... 35 

2.10.1.3  Freeze-Thaw Durability Studies Using Resilient Modulus Testing .. 35 

2.10.1.4  Standard Freeze-Thaw Test of Soils Stabilised with Pozzolans


(Vacuum Saturation Test) ............................................................................... 37 

xiii
2.10.1.5  Freeze-Thaw Durability Studies of Flexural Properties .................... 38 

2.10.1.6  Freeze-Thaw Studies using Tube Suction Test.................................. 39 

2.10.2  Wet-Dry Laboratory Durability Tests ................................................... 45 

2.11  Moisture content fluctuation within the Pavement........................................... 48 

Chapter 3 Research Methodology .............................................................................. 50 

3.1  Introduction ...................................................................................................... 50 

3.2  Materials ........................................................................................................... 54 

3.2.1  Base Materials ............................................................................................. 54 

3.2.1.1  Queensland Granular Material (QM) .................................................... 54 

3.2.1.2  Canberra Granular Material (CM) ......................................................... 59 

3.2.2  Subgrade Material ........................................................................................ 60 

3.2.3  Lightly Stabilised Granular Base ................................................................. 64 

3.2.3.1  Cementitious Stabilisation ..................................................................... 66 

3.2.3.1.1  General Blended Cement ................................................................ 68 

3.2.3.1.2  Fly Ash............................................................................................ 69 

3.2.3.1.3  Slag–Lime (SL)............................................................................... 70 

3.3  Laboratory Testing Methods ............................................................................ 71 

3.3.1  Freeze-Thaw Durability Testing .................................................................. 71 

3.3.1.1  Assessing Freeze-Thaw Durability using Unconfined Compression


Strength Test.................................................................................................... 71 

3.3.1.2  Assessing Freeze-Thaw Durability using Repeated Load Triaxial Test72 

3.3.1.3  Assessing Freeze-Thaw Durability using Flexural Beam Test ............. 73 

xiv
3.3.1.4  Assessing Durability using Vacuum Saturation Test ............................ 73 

3.3.2  Wet-Dry Durability Test .............................................................................. 74 

3.3.3  Moisture Susceptibility Test ........................................................................ 74 

3.3.3.1  Assessing Moisture Susceptibility Using TST ...................................... 74 

3.3.3.2  Assessing Moisture Susceptibility using Unconfined Compressive


Strength and Flexural Test .............................................................................. 75 

3.3.4  Characterising Deformation Responses ....................................................... 75 

3.4  Summary .......................................................................................................... 76 

Chapter 4 Evaluating Suitability of Direct and Indirect Laboratory Tests for


Assessing Freeze-Thaw and Moisture Effects on Lightly Stabilised Granular
Materials.................................................................................................................... 77 

4.1  Introduction ...................................................................................................... 77 

4.2  Experimental Program ...................................................................................... 79 

4.2.1  Materials and Binders .................................................................................. 79 

4.2.2  Freeze-Thaw Durability Test ....................................................................... 79 

4.2.3  7-day Unconfined Compressive Strength Testing ....................................... 81 

4.2.4  Vacuum Saturation Test .............................................................................. 82 

4.2.4.1  Experimental Procedure ........................................................................ 82 

4.2.5  Tube Suction Test ........................................................................................ 85 

4.2.5.1  Experimental Procedure ........................................................................ 86 

4.2.6  Flexural and UCS Testing to study Effects of Moisture Content ................ 92 

4.3  Results and Discussion ..................................................................................... 95 

4.3.1  Freeze-Thaw Durability Test Result ............................................................ 95 

xv
4.3.2  7-day Unconfined Compressive Strength Test Results ............................... 96 

4.3.3  Vacuum Saturation Test Result ................................................................... 96 

4.3.4  Tube Suction Test Results ......................................................................... 101 

4.3.4.1  Effect of Moulding Moisture Content on Dielectric Value ................. 101 

4.3.4.2  Tube Suction Test with Different Binders and Binder Percentages .... 107 

4.3.4.3  Effect of Porous Disk on Dielectric Measurement .............................. 112 

4.3.4.4  Comparison of Dielectric Values measured from TST Setups 1 and 2


113 

4.3.4.5  Comparison of Dielectric Values and Moisture Content .................... 115 

4.3.5  Comparison of Durability Test Results ..................................................... 118 

4.3.6  Flexural and UCS varying with moisture content ..................................... 123 

4.4  Summary ........................................................................................................ 126 

Chapter 5 Assessing Durability of Lightly Stabilised Granular Material against


Freeze-Thaw and Wet-Dry Cycles Using Different Element Testings and Testing
Procedures ............................................................................................................... 129 

5.1  Introduction .................................................................................................... 129 

5.2  Experimental Program .................................................................................... 131 

5.2.1  Unconfined Compressive Strength Test .................................................... 131 

5.2.1.1  Introduction ......................................................................................... 131 

5.2.1.2  Sample Preparation and Testing .......................................................... 134 

5.2.2  Repeated Load Triaxial Test ...................................................................... 135 

5.2.2.1  Introduction ......................................................................................... 135 

5.2.2.2  Sample Preparation and Curing ........................................................... 137 

xvi
5.2.2.3  Installation of Specimen ...................................................................... 140 

5.2.2.4  Deformation Measurement .................................................................. 142 

5.2.2.5  Stress Control Dynamic Test ............................................................... 146 

5.2.2.6  Repeated Load Triaxial Resilient Modulus Test ................................. 150 

5.2.2.7  Repeated Load Triaxial Permanent Deformation Test ........................ 162 

5.2.3  Flexural Test .............................................................................................. 164 

5.2.3.1  Introduction ......................................................................................... 164 

5.2.3.2  Sample Preparation.............................................................................. 168 

5.2.3.3  Experimental Setup ............................................................................. 169 

5.2.4  Freeze-Thaw and Wet-Dry Cycles ............................................................ 171 

5.2.4.1  F-T Cycles ........................................................................................... 172 

5.2.4.2  W-D Cycles ......................................................................................... 173 

5.3  Results and Discussion ................................................................................... 176 

5.3.1  Effect of Stress and Loading Frequency on Resilient Modulus ................ 176 

5.3.2  F-T Durability Test .................................................................................... 183 

5.3.2.1  F-T Durability Study using Unconfined Compression Strength Test . 184 

5.3.2.2  Freeze-Thaw Durability Study using Repeated Load Triaxial Test .... 202 

5.3.2.2.1  Freeze-Thaw Effects on Resilient Modulus.................................. 203 

5.3.2.2.2  Effects of Freeze-Thaw on Repeated Load Triaxial Permanent


Deformation Test ....................................................................................... 217 

5.3.2.2.3  Freeze-Thaw Effects on Residual Unconfined Compressive


Strength ...................................................................................................... 231 

xvii
5.3.2.3  Freeze-Thaw Effects on Modulus of Rupture and Static Stiffness
Modulus ......................................................................................................... 236 

5.3.3  Wet-Dry Durability Test using Different Element Testings...................... 244 

5.3.3.1  W-d Durability Study using UCS Test ................................................ 245 

5.3.3.2  W-d Durability Study using RLT Test ................................................ 253 

5.3.3.3  Wet-Dry Durability Test using Flexural Test...................................... 261 

5.4  Summary ........................................................................................................ 264 

Chapter 6 Characterising Permanent Deformation Behaviour of Lightly Stabilised


Granular Material using Repeated Load Triaxle (RLT) Test ........................... 269 

6.1  Introduction .................................................................................................... 269 

6.2  Experimental Methodology .............................. Error! Bookmark not defined. 

6.2.1  Granular Materials and Binders ................................................................. 270 

6.2.2  Multi-stage Permanent Deformation Testing with Varying Moisture


Contents ............................................................................................................ 271 

6.3  Results and Discussion ................................................................................... 273 

6.3.1  Classification of Material Responses using Shakedown Limits ................ 273 

6.3.2  Calculation of Shakedown Limits ............................................................. 277 

6.3.3  Analysis of Resilient Deformation ............................................................ 281 

6.3.4  Influence of Moisture on Resilient Modulus ............................................. 287 

6.3.5  Characterisation of Permanent Deformation Behaviour............................ 290 

6.3.6  Relationship between Permanent and Elastic Deformations ..................... 295 

6.4  Summary ........................................................................................................ 297 

xviii
Chapter 7 Characterising Deformation Behaviour of Lightly Stabilised Granular
Material using Accelerated Pavement Model (APM) Test ................................. 299 

7.1  Introduction .................................................................................................... 299 

7.2  Accelerated Pavement Model Testing Facility............................................... 301 

7.2.1  Measurement of Vertical Deformation ...................................................... 305 

7.2.2  Pressure Measurements ............................................................................. 308 

7.2.3  In-situ Moisture Measurements ................................................................. 308 

7.2.3.1  Time Domain Reflectometer ............................................................... 310 

7.2.3.2  Neutron Probe...................................................................................... 310 

7.2.3.3  Standing-Wave Ratio (STR)................................................................ 311 

7.2.3.4  Capacitance Probe ............................................................................... 311 

7.2.3.5  Moisture Probe (MP306) ..................................................................... 312 

7.2.3.6  Laboratory Calibration of MP306 Probes ........................................... 313 

7.2.4  Preparation of Subgrade and Base Layers ................................................. 316 

7.2.5  Data acquisition ......................................................................................... 320 

7.2.6  Testing ....................................................................................................... 321 

7.3  Repeated Load Triaxial Test .......................................................................... 324 

7.4  Results and Discussion ................................................................................... 325 

7.4.1  Accelerated Pavement Model Test Results ............................................... 325 

7.4.1.1  Permanent Deformation Responses ..................................................... 326 

7.4.1.2  Resilient Deformation Responses........................................................ 328 

7.4.1.3  Moisture Content Readings ................................................................. 331 

xix
7.4.2  Finite Element Modelling and Back Calculation of Mechanical Properties
333 

7.4.2.1  Introduction ......................................................................................... 333 

7.4.2.2  Development of Finite Element Model ............................................... 336 

7.4.2.3  Boundary Conditions and Loading...................................................... 338 

7.4.2.4  Back-Calculation Considering Isotropic Elastic Properties ................ 340 

7.4.2.5  Back-Calculation Considering Cross-anisotropic Elastic Properties .. 345 

7.4.2.6  Comparison of Vertical Pressure Distributions ................................... 349 

7.4.2.7  Comparison of Horizontal Pressure Distributions............................... 353 

7.4.3  Repeated Load Triaxial Test Results ......................................................... 357 

7.4.3.1  Permanent and Resilient Response of Repeated Load Triaxial Test ... 357 

7.4.3.2  Selecting Stress Values from Experiments.......................................... 358 

7.4.3.3  Selection Stress Values from ABAQUS ............................................. 359 

7.5  Summary ........................................................................................................ 363 

Chapter 8 Summary, Conclusions and Recommendations .................................... 366 

8.1  General ........................................................................................................... 366 

8.2  Conclusions .................................................................................................... 366 

8.2.1  Evaluating Suitability of Direct and Indirect Laboratory Tests for Assessing
Freeze-Thaw and Moisture Effects on Lightly Stabilised Granular Materials . 366 

8.2.2  Assessing Durability of Lightly Stabilised Granular Material against


Freeze-Thaw and Wet-Dry Cycles Using Different Element Testings and
Testing Procedures ............................................................................................ 368 

8.2.3  Characterising Permanent Deformation Behaviour `of Lightly Stabilised


Granular Material using Repeated Load Triaxle (RLT) Test ........................... 369 

xx
8.2.4  Characterising Deformation Behaviour of Lightly Stabilised Granular
Material using Accelerated Pavement Model (APM) Test ............................... 370 

8.3  Recommendations for Further Research ........................................................ 372 

References .................................................................................................................... 374 

Appendix A ..................................................................................................................... A 

Appendix B ..................................................................................................................... B 

xxi
List of Figures
Figure 2.1: Typical flexible pavement ............................................................................ 10 
Figure 2.2: Typical rigid pavement ................................................................................. 10 
Figure 2.3: EICM scheme (Quintero 2007) .................................................................... 15 
Figure 2.4: Scanning electron microscope image of fly ash ........................................... 22 
Figure 2.5: Strains under repeated loads (Huang 2004) .................................................. 32 
Figure 3.1: Queensland granular material (QM) ............................................................. 55 
Figure 3.2: Particle size distributions of QM materials .................................................. 56 
Figure 3.3: Reconstituted particle size distributions of CM and QM ............................. 58 
Figure 3.4: OMC-MDD relationship of QM ................................................................... 59 
Figure 3.5: OMC-MDD relationship of CM ................................................................... 60 
Figure 3.6: Canberra granular material (CM) ................................................................. 61 
Figure 3.7: Queensland subgrade material ...................................................................... 61 
Figure 3.8: Particle size distribution of Queensland subgrade material.......................... 62 
Figure 3.9: OMC-MDD relationship of Queensland subgrade material ......................... 63 
Figure 3.10: Guide for selecting binder based on PI....................................................... 67 
Figure 4.1: Sealed container used for curing specimens tested for vacuum saturation test
................................................................................................................................. 84 
Figure 4.2: Pictorial view of vacuum saturation equipment ........................................... 85 
Figure 4.3: Percentages of moisture loss with number of days....................................... 89 
Figure 4.4: TST experimental setup 1 ............................................................................. 89 
Figure 4.5: Specimen prepared for TST experimental setup 1 ....................................... 90 
Figure 4.6: TST experimental setup 2 ............................................................................. 90 
Figure 4.7: Specimen prepared for TST experimental setup 2 ....................................... 91 
Figure 4.8: Percometer and surface probe used for DV measurement ........................... 91 
Figure 4.9: Five surface probe reading locations ............................................................ 92 
Figure 4.10: UCS of CM specimens stabilised with 1.5% CF and subjected to 12 f-t
cycles varying with different moulding moisture contents ..................................... 98 

xxii
Figure 4.11: UCS of QM specimens stabilised with 1.5% CF and subjected to 12 f-t
cycles varying with different moulding moisture contents ..................................... 98 
Figure 4.12: UCS of CM specimens stabilised with 1.5% CF and subjected to vacuum
saturation varying with different moulding moisture contents ............................... 99 
Figure 4.13: UCS of QM specimens stabilised with 1.5% CF and subjected to vacuum
saturation varying with different moulding moisture contents ............................... 99 
Figure 4.14: 7-day UCS of CM specimens stabilised with 1.5% CF varying with
different moulding moisture contents ................................................................... 100 
Figure 4.15: 7-day UCS of QM specimens stabilised with 1.5% CF varying with
different moulding moisture contents ................................................................... 100 
Figure 4.16: DVs of CM stabilised with 1.5% CF varying with time, measured using
TST setup 1 ........................................................................................................... 103 
Figure 4.17: DVs of CM stabilised with 0.5% CF varying with time, measured using
TST setup 1 ........................................................................................................... 104 
Figure 4.18: DVs of QM stabilised with 1.5% CF varying with time, measured using
TST setup 1 ........................................................................................................... 104 
Figure 4.19: DVs of QM stabilised with 0.5% CF varying with time, measured using
TST setup 1 ........................................................................................................... 105 
Figure 4.20: DVs of CM stabilised with 1.5% CF varying with time, measured using
TST setup 2 ........................................................................................................... 105 
Figure 4.21: DVs of CM stabilised with 0.5% CF varying with time, measured using
TST setup 2 ........................................................................................................... 106 
Figure 4.22: DVs of QM stabilised with 1.5% CF varying with time, measured using
TST setup 2 ........................................................................................................... 106 
Figure 4.23: DVs of QM stabilised with 0.5% CF varying with time, measured using
TST setup 2 ........................................................................................................... 107 
Figure 4.24: Comparison of DVs of different mixes of stabilised materials ................ 109 
Figure 4.25: DVs of CM stabilised with 0.5%, 1.5% and 3% CF varying with time ... 109 
Figure 4.26: DVs of CM stabilised with 3%, 4.5% and 6% SL varying with time ...... 110 
Figure 4.27: DVs of QM stabilised with 0.5%, 1.5% and 3% CF varying with time ... 110 
Figure 4.28: DVs of QM stabilised with 3%, 4.5% and 6% CF varying with time ...... 111 
Figure 4.29: Relationships between UCS and maximum DV for CM and QM ........... 111 
Figure 4.30: Different types of porous disks used for TST test .................................... 112 
Figure 4.31: DVs measured using different types of porous disks ............................... 113 

xxiii
Figure 4.32: Comparison of DVs measured from TST setups 1 and 2 ......................... 115 
Figure 4.33: Moisture contents measured from bottom and top of specimen............... 116 
Figure 4.34: Correlations between moisture contents measured after TST and respective
DVs measured during TST.................................................................................... 117 
Figure 4.35: Correlations between UCS obtained VST and f-t durability tests ............ 120 
Figure 4.36: Correlations between UCS obtained from vacuum saturation and 7-day
UCS tests ............................................................................................................... 121 
Figure 4.37: Correlations between UCS and final DVs obtained from vacuum saturation
test and TST respectively ...................................................................................... 121 
Figure 4.38: Correlations between UCSs and final DVs obtained from vacuum
saturation test and TST respectively ..................................................................... 122 
Figure 4.39: Correlations between UCSs and final DVs obtained from vacuum
saturation test and TST respectively ..................................................................... 122 
Figure 4.40: Correlations between UCS and final DVs obtained from vacuum saturation
test and TST respectively ...................................................................................... 123 
Figure 4.41: MoR varying with moisture content ......................................................... 124 
Figure 4.42: Flexural modulus varying with moisture content ..................................... 125 
Figure 4.43: UCS varying with moisture content ......................................................... 125 
Figure 5.1: Dry and wet mixes of Queensland granular base material (QM) ............... 138 
Figure 5.2: Split mould used to mould 100 mm diameter by 195 mm height cylindrical
specimens .............................................................................................................. 139 
Figure 5.3: Triaxial cylindrical specimen used for the f-t and w-d testings ................. 139 
Figure 5.4: Schematic diagram of triaxial testing system ............................................. 141 
Figure 5.5: Photographic view of triaxial testing system.............................................. 142 
Figure 5.6: Internal and external LVDT arrangements (Ping et al. 2003) .................... 144 
Figure 5.7: On-specimen deformation measurement (Ekblad 2006) ............................ 145 
Figure 5.8: On-specimen LVDT arrangement .............................................................. 147 
Figure 5.9: Control program TestStarTM ....................................................................... 148 
Figure 5.10: TestWare-SX program used to input stress sequences ............................. 149 
Figure 5.11: Typical load-time history ......................................................................... 149 
Figure 5.12: Typical deformation-time history ............................................................. 150 

xxiv
Figure 5.13: Percentage differences between measured and applied cyclic stresses of
lightly stabilised material and unbound material tested based on NCHRP (2008) Mr
standard ................................................................................................................. 158 
Figure 5.14: Percentage differences between LVDT readings of lightly stabilised
material and unbound material tested based on NCHRP (2008) Mr standard ...... 158 
Figure 5.15: Stress dependency of Mr ........................................................................... 160 
Figure 5.16: Stress paths followed in PD testing .......................................................... 164 
Figure 5.17: Photographic view of steel mould, 76 mm by 76 mm by 284 mm used for
preparing flexural beam specimens ....................................................................... 169 
Figure 5.18: Compacted 76 mm by 76 mm by 284 mm flexural beam specimen ........ 169 
Figure 5.19: Experimental setup for flexural testing .................................................... 171 
Figure 5.20: Photographic view of flexural testing arrangement (Paul et al. 2015) ..... 171 
Figure 5.21: Environmental chamber used for f-t cycles .............................................. 173 
Figure 5.22: Triaxial specimen placed on top of moisture pad ..................................... 174 
Figure 5.23: Testing set-up for w-d process ................................................................. 175 
Figure 5.24 Load frequencies (a) 10 Hz (b) 5 Hz (c) 1 Hz ........................................... 178 
Figure 5.25: Variations in Mr with different frequencies and bulk stresses.................. 180 
Figure 5.26: Variations in Mr with different deviatoric stresses and confining pressures
(a) 10 Hz (b) 5 Hz (c) 1 Hz ................................................................................... 182 
Figure 5.27: Variations in UCS with number of f-t cycles for QM and CM with 1.5%
CF binder............................................................................................................... 189 
Figure 5.28: Variations in UCS variations with number of f-t cycles for QM and CM
with 1.5% CF binder ............................................................................................. 189 
Figure 5.29: Strength-gaining patterns of QM lightly stabilised granular material with
(a) 1.5% CF binder (b) 3.0% CF binder ................................................................ 191 
Figure 5.30: Influence of temperature on strength of concrete cast and cured at the
temperature indicated (Neville and Brooks 2010) ................................................ 195 
Figure 5.31: Percentage changes in UCS relative to 0 cycles of normally cured QM
specimens and those subjected to f-t procedures 1 and 2...................................... 198 
Figure 5.32: Variations in moisture content with number of f-t cycles for QM with 1.5%
and 3% CF binders ................................................................................................ 198 
Figure 5.33: Variations in moisture content with number of f-t cycles for QM and CM
subjected to f-t procedures 1 and 2 ....................................................................... 199 
Figure 5.34: Variations in UCS with different moisture contents ................................ 199 

xxv
Figure 5.35: Comparisons of QM and CM ................................................................... 201 
Figure 5.36: Particle size distributions of CM and QM ................................................ 202 
Figure 5.37: Variations in Mr with bulk stress of QM stabilised with 1.5% CF binder,
cured for 7 days and subjected to f-t procedure 1 ................................................. 205 
Figure 5.38: Variations in Mr with number of f-t cycles for typical stress sequences 4, 8,
12 and 16 (Series 1, QM) ...................................................................................... 206 
Figure 5.39: Variations in Mr with bulk stress of CM stabilised with 1.5% CF binder,
cured for 7 days and subjected to f-t procedure 1 ................................................. 207 
Figure 5.40: Variations in Mr with number of w-d cycles for typical stress sequences 4,
8, 12 and 16 (Series 1, QM) .................................................................................. 207 
Figure 5.41: Mr varying with bulk stress of QM stabilised with 1.5% CF binder, cured
for 28 days and subjected to f-t procedure 1 ......................................................... 209 
Figure 5.42: Mr varying with number of w-d cycles for typical stress sequences 4, 8, 12
and 16 (Series 2, QM) ........................................................................................... 209 
Figure 5.43: Mr varying with bulk stress of CM stabilised with 1.5% CF binder, cured
for 28 days and subjected to f-t procedure 1 ......................................................... 210 
Figure 5.44: Mr varying with number of w-d cycles for typical stress sequences 4, 8, 12
and 16 (Series 2, CM) ........................................................................................... 211 
Figure 5.45: Mr of Canberra material stabilised with 1.5% CF binder, cured for 7 days
and subjected to f-t procedure 2 ............................................................................ 213 
Figure 5.46: Mr varying with number of w-d cycles for typical stress sequences 4, 8, 12
and 16 (Series 3, CM) ........................................................................................... 213 
Figure 5.47: Mr of CM stabilised with 1.5% CF binder, cured for 7 days at 40oC and
subjected to f-t procedure 2 ................................................................................... 214 
Figure 5.48: Mr varying with number of w-d cycles for typical stress sequences 4, 8, 12
and 16 (Series 3, CM) ........................................................................................... 215 
Figure 5.49: Accumulation of differences in Mr (MPa) of selected stress sequences .. 216 
Figure 5.50: PDs of QM stabilised with 1.5% CF binder, cured for 7 days and subjected
to f-t procedure 1 ................................................................................................... 218 
Figure 5.51: PD of QM stabilised with 1.5% CF binder, cured for 7 days and subjected
to f-t procedure 1 ................................................................................................... 220 
Figure 5.52: PDs of CM stabilised with 1.5% CF binder, cured for 7 days and subjected
to f-t procedure 1 ................................................................................................... 220 
Figure 5.53: PDs of QM stabilised with 1.5% CF binder, cured for 28 days and
subjected to f-t procedure 1 ................................................................................... 221 

xxvi
Figure 5.54: PDs of CM stabilised with 1.5% CF binder, cured for 28 days and
subjected to f-t procedure 1 ................................................................................... 221 
Figure 5.55: PD of CM stabilised with 1.5% CF binder, cured for 7 days at 40oC and
subjected to f-t procedure 2 ................................................................................... 222 
Figure 5.56: PD of CM stabilised with 1.5% CF binder, cured for 7 days and subjected
to f-t procedure 2 ................................................................................................... 222 
Figure 5.57: Elasto-plastic responses of material under repeated loading (Johnson 1986)
............................................................................................................................... 226 
Figure 5.58: Vertical permanent strain rates versus vertical permanent strains - range A
............................................................................................................................... 227 
Figure 5.59: Vertical resilient strain varying with number of load cycles - range A .... 227 
Figure 5.60: Vertical permanent strain rates versus vertical permanent strain_ range C
............................................................................................................................... 228 
Figure 5.61: Vertical resilient strain varying with number of load cycles_range C ..... 229 
Figure 5.62: Vertical permanent strain rates versus vertical permanent strain_ range B
............................................................................................................................... 230 
Figure 5.63: Percentage declines in RUCS/UCS with number of f-t cycles of specimens
cured for 7 days and subjected to f-t procedure 1 ................................................. 234 
Figure 5.64: Percentage declines in RUCS/UCS with number of f-t cycles of specimens
cured for 28 days and subjected to f-t procedure 1 ............................................... 234 
Figure 5.65: Percentage decline in RUCS/UCS with number of f-t cycles (procedure 2)
of specimens cured at 400C and normally cured specimens ................................. 235 
Figure 5.66: Relationship between RUCS and UCS..................................................... 235 
Figure 5.67: Variations in MoR with number of cycles for 1.5% CF binder ............... 240 
Figure 5.68: Variations in SSM with number of cycles for 1.5% CF binder................ 240 
Figure 5.69: Variations in flexural modulus with number of cycles for 3% CF binder 241 
Figure 5.70: Variations in MoR with number of cycles for 3% CF binder .................. 241 
Figure 5.71: Flexural stress-strain relationship of specimens stabilised with 1.5% of CF
binder and subjected to various f-t cycles ............................................................. 242 
Figure 5.72: Flexural stress-strain relationship of specimens stabilised with 3% of CF
binder and subjected to various f-t cycles ............................................................. 242 
Figure 5.73: Relationship between flexural modulus and MoR of specimens subjected to
f-t cycles and specimens not subjected to f-t cycles ............................................. 243 
Figure 5.74: UCS specimens after 1 w-d cycle curing (1.5% CF binder - QM) ........... 246 

xxvii
Figure 5.75: UCS specimens after 8 w-d cycles (1.5% CF binder) .............................. 247 
Figure 5.76: UCS specimens at end of curing (3% CF binder) .................................... 247 
Figure 5.77: UCS specimens after 12 w-d cycles (3% CF binder) ............................... 248 
Figure 5.78: UCS of specimens subjected to w-d cycles and normally cured specimens
with 1.5% CF binders............................................................................................ 250 
Figure 5.79: UCS of specimens subjected to w-d cycles and normally cured specimens
with 3% CF binders............................................................................................... 250 
Figure 5.80: Triaxial specimens after 0 and 1 w-d cycles ........................................... 255 
Figure 5.81: Triaxial specimens after 2 and 4 w-d cycles ............................................ 255 
Figure 5.82: Resilient modulus varying with bulk stress of specimens subjected to w-d
cycles ..................................................................................................................... 257 
Figure 5.83: Resilient modulus varying with number of w-d cycles for typical stress
sequences 4, 8, 12 and 16 ...................................................................................... 258 
Figure 5.84: Permanent deformation with different numbers of w-d cycles ................ 259 
Figure 5.85: Permanent strain rates versus vertical permanent of specimen subjected to
2 w-d cycles .......................................................................................................... 260 
Figure 5.86: Permanent strain rates versus vertical permanent of loading sequence 4. 260 
Figure 5.87: MoR for different numbers of w-d cycles ................................................ 262 
Figure 5.88: Specimen subjected to 4 w-d cycles and the specimen subjected to 8 w-d
cycles is being tested for failure load only ............................................................ 262 
Figure 6.1: Shakedown limit lines of specimens moulded with 8% moisture content and
1.5% CF binder ..................................................................................................... 278 
Figure 6.2: Interpolation for determining exact point of shakedown limit line ............ 278 
Figure 6.3: Shakedown limit lines of specimens moulded with 6.5% moisture content
and 1.5% CF binder .............................................................................................. 279 
Figure 6.4: Comparison of equations for PSL and PCL of specimens compacted at 8%
and 6.5% moisture contents .................................................................................. 280 
Figure 6.5: Typical resilient deformation response in shakedown range A.................. 283 
Figure 6.6: Typical resilient deformation responses in shakedown range B ................ 284 
Figure 6.7: Typical resilient deformation response in shakedown range C just before
failure of specimen ................................................................................................ 285 
Figure 6.8: Resilient strain varying with stress ratio for specimens stabilised with 1.5%
CF binder and compacted at different moisture contents...................................... 286 

xxviii
Figure 6.9: Resilient strain varying with stress ratio for specimens stabilised with 0.5%
CF binder and compacted at different moisture contents...................................... 287 
Figure 6.10: Resilient modulus obtained for 8% and 6.5% moisture contents ............. 290 
Figure 6.11: Regression constant ‘b1’ varying with bulk stress (MC=8%) .................. 294 
Figure 6.12: Regression constant ‘b1’ varying with bulk stress (MC=6.5%) ............... 294 
Figure 6.13: Permanent strain rate varying with resilient strain of specimens compacted
at 8.0% moisture content ....................................................................................... 296 
Figure 6.14: Permanent strain rate varying with resilient strain of specimens compacted
at 6.5% moisture content ....................................................................................... 297 
Figure 7.1: Schematic diagram of accelerated pavement model testing facility........... 303 
Figure 7.2: Cylindrical tank used for APM testing ....................................................... 304 
Figure 7.3: Loading actuator assembly ......................................................................... 304 
Figure 7.4: ‘I’ section used for beam and columns in the steel frame .......................... 305 
Figure 7.5: 184-mm diameter loading plate with steel plates attached on which to mount
LVDTs for measuring top surface deformation .................................................... 306 
Figure 7.6: In-depth measurement device for measuring vertical deformations .......... 307 
Figure 7.7: Locations at which vertical deformation measurements were measured ... 307 
Figure 7.8: Locations of installed EPCs........................................................................ 309 
Figure 7.9: Moisture probe (MP306) selected for experiments .................................... 314 
Figure 7.10: MP306, data logger and SMM ................................................................. 314 
Figure 7.11: MP306 moisture sensors’ calibration curves for base material ................ 315 
Figure 7.12: MP306 moisture sensors’ calibration curves for subgrade material......... 315 
Figure 7.13: Locations of installed MP306s ................................................................. 316 
Figure 7.14: Subgrade construction in the cylindrical tank .......................................... 318 
Figure 7.15: Aluminium grid for supporting LVDT holders ........................................ 322 
Figure 7.16:Typical loading cycle with rest period ...................................................... 323 
Figure 7.17: Typical outputs showing LVDT readings and cyclic loadings................. 323 
Figure 7.18: Surface and subgrade permanent deformations varying with number of
load cycles ............................................................................................................. 327 
Figure 7.19: Base deformations varying with number of load cycles .......................... 327 
Figure 7.20: Surface and subgrade resilient deformations varying with number of load
cycles ..................................................................................................................... 329 

xxix
Figure 7.21: Base resilient deformation varying with number of load cycles .............. 330 
Figure 7.22: Vertical permanent strain rate varying with vertical permanent strain .... 330 
Figure 7.23: Moisture content in top surface varying with time ................................... 332 
Figure 7.24: Moisture contents at interface and 100 mm in subgrade varying with time
............................................................................................................................... 333 
Figure 7.25: Configuration of APM .............................................................................. 336 
Figure 7.26: Schematic of axisymmetric model selected for finite element modelling 337 
Figure 7.27: Part module of axisymmetric model......................................................... 337 
Figure 7.28: FE mesh used for axisymmetric model .................................................... 338 
Figure 7.29: Matching ABAQUS surface vertical resilient deformations and measured
experimental deformations (loading pressure 750 kPa) ........................................ 341 
Figure 7.30: Matching ABAQUS middle and interface vertical resilient deformations
and measured experimental deformations (loading pressure 750 kPa) ................. 341 
Figure 7.31: Matching ABAQUS surface vertical resilient deformations and measured
experimental deformations (loading pressure 1000 kPa) ...................................... 342 
Figure 7.32: Matching ABAQUS middle and interface vertical resilient deformations
and measured experimental deformations (loading pressure 1000 kPa) ............... 342 
Figure 7.33: Matching ABAQUS surface vertical resilient deformations and measured
experimental deformations (loading pressure 1250 kPa) ...................................... 343 
Figure 7.34: Matching ABAQUS middle and interface vertical resilient deformations
and measured experimental deformations (loading pressure 1250 kPa) ............... 343 
Figure 7.35: Matching ABAQUS surface vertical resilient deformations and measured
experimental deformations (loading pressure 1500 kPa) ...................................... 344 
Figure 7.36: Matching ABAQUS surface vertical resilient deformations and measured
experimental deformations (loading pressure 1500 kPa) ...................................... 344 
Figure 7.37: Layered compaction makes cross-anisotropic pavement layers ............... 346 
Figure 7.38: Comparison of moduli of base materials .................................................. 348 
Figure 7.39: Comparison of moduli of subgrade materials .......................................... 348 
Figure 7.40: Variations in EPC measurements at base-subgrade interface with number
of cycles ................................................................................................................ 349 
Figure 7.41: Vertical pressure distributions measured using EPCs at different depths 350 
Figure 7.42: Vertical pressure distributions obtained from isotropic analysis.............. 351 
Figure 7.43: Vertical pressure distributions obtained from cross-anisotropic analysis 351 

xxx
Figure 7.44: Vertical pressure distributions obtained from cross-anisotropic analysis for
1500 kPa cyclic loading ........................................................................................ 352 
Figure 7.45: Locations at which horizontal pressure distributions obtained from cross-
anisotropic analysis ............................................................................................... 354 
Figure 7.46: Horizontal pressure distributions obtained at B-B from cross-anisotropic
analysis .................................................................................................................. 355 
Figure 7.47: Horizontal pressure distributions obtained at boundary wall from cross-
anisotropic analysis ............................................................................................... 356 
Figure 7.48: Horizontal pressure distributions in base layer obtained from cross-
anisotropic analysis for 1500 kPa cyclic loading .................................................. 356 
Figure 7.49: Typical RLT permanent and resilient deformation response ................... 357 
Figure 7.50: Regression parameters a and b varying with bulk stress .......................... 359 
Figure 7.51: An imaginary RLT specimen size cylindrical in the accelerated pavement
model ..................................................................................................................... 361 
Figure 7.52: Vertical pressure distribution along A-B .................................................. 361 
Figure 7.53: Horizontal pressure distribution along C-D.............................................. 362 
Figure 7.54: Relationship between regression coefficients........................................... 363 

Figure A. 1: Flexural stress-strain relationship of specimen stabilised with 1.5% CF


binder and subjected to 0 f-t cycle ........................................................................ A-1 
Figure A. 2: Flexural stress-strain relationship of specimen stabilised with 1.5% CF
binder and subjected to 1 f-t cycle ........................................................................ A-1 
Figure A. 3: Flexural stress-strain relationship of specimen stabilised with 1.5% CF
binder and subjected to 2 f-t cycles ....................................................................... A-2 
Figure A. 4: Flexural stress-strain relationship of specimen stabilised with 1.5% CF
binder and subjected to 4 f-t cycl .......................................................................... A-2 
Figure A. 5: Flexural stress-strain relationship of specimen stabilised with 1.5% CF
binder and subjected to 8 f-t cycles ....................................................................... A-3 
Figure A. 6: Flexural stress-strain relationship of specimen stabilised with 1.5% CF
binder and subjected to 12 f-t cycles ..................................................................... A-3 
Figure A. 7: Flexural stress-strain relationship of specimen stabilised with 3% CF
binder and subjected to 0 f-t cycle ........................................................................ A-4 
Figure A. 8: Flexural stress-strain relationship of specimen stabilised with 3% CF
binder and subjected to 1 f-t cycle ........................................................................ A-4 

xxxi
Figure A. 9: Flexural stress-strain relationship of specimen stabilised with 3% CF
binder and subjected to 2 f-t cycles ....................................................................... A-5 
Figure A. 10: Flexural stress-strain relationship of specimen stabilised with 3% CF
binder and subjected to 4 f-t cycles ....................................................................... A-5 
Figure A. 11: Flexural stress-strain relationship of specimen stabilised with 3% CF
binder and subjected to 8 f-t cycles ....................................................................... A-6 
Figure A. 12: Flexural stress-strain relationship of specimen stabilised with 3% CF
binder and subjected to 12 f-t cycles ..................................................................... A-6 

Figure B. 1: Resilient strain measured in RLT tests varying with number of cycles of
specimens compacted at 8% moisture content with 50 kPa confining pressure for
(a) stress sequence 1 and (b) stress sequence 2 ..................................................... B-1 
Figure B. 2: Resilient strain measured in RLT tests varying with number of cycles of
specimens compacted at 8% moisture content with 50 kPa confining pressure for
(a) stress sequence 3 and (b) stress sequence 4 ..................................................... B-2 
Figure B. 3: Resilient strain measured in RLT tests varying with number of cycles of
specimens compacted at 8% moisture content with 100 kPa confining pressure for
(a) stress sequence 1 and (b) stress sequence 2 ..................................................... B-3 
Figure B. 4: Resilient strain measured in RLT tests varying with number of cycles of
specimens compacted at 8% moisture content with 100 kPa confining pressure for
(a) stress sequence 3 and (b) stress sequence 4 ..................................................... B-4 
Figure B. 5: Resilient strain measured in RLT tests varying with number of cycles of
specimens compacted at 8% moisture content and tested with with 100 kPa
confining pressure for stress sequence 5 ............................................................... B-5 
Figure B. 6: Resilient strain measured in RLT tests varying with number of cycles of
specimens compacted at 8% moisture content with 140 kPa confining pressure for
(a) stress sequence 1 and (b) stress sequence 2 ..................................................... B-6 
Figure B. 7: Resilient strain measured in RLT tests varying with number of cycles of
specimens compacted at 8% moisture content with 140 kPa confining pressure for
(a) stress sequence 3 and (b) stress sequence 4 ..................................................... B-7 
Figure B. 8: Resilient strain measured in RLT tests varying with number of cycles of
specimens compacted at 8% moisture content with 140 kPa confining pressure for
stress sequence ...................................................................................................... B-8 
Figure B. 9: Resilient strain measured in RLT tests varying with number of cycles of
specimens compacted at 8% moisture content with 180 kPa confining pressure for
(a) stress sequence 1 and (b) stress sequence 2 ..................................................... B-9 

xxxii
Figure B. 10: Resilient strain measured in RLT tests varying with number of cycles of
specimens compacted at 8% moisture content with 180 kPa confining pressure for
(a) stress sequence 3 and (b) stress sequence 4 ................................................... B-10 
Figure B. 11: Resilient strain measured in RLT tests varying with number of cycles of
specimens compacted at 6.5% moisture content with 50 kPa confining pressure for
(a) stress sequence 1 and (b) stress sequence 2 ................................................... B-11 
Figure B. 12: Resilient strain measured in RLT tests varying with number of cycles of
specimens compacted at 6.5% moisture content with 50 kPa confining pressure for
(a) stress sequence 3 and (b) stress sequence 4 ................................................... B-12 
Figure B. 13: Resilient strain measured in RLT tests varying with number of cycles of
specimens compacted at 6.5% moisture content with 50 kPa confining pressure for
stress sequence 5 ................................................................................................. B-13 
Figure B. 14: Resilient strain measured in RLT tests varying with number of cycles of
specimens compacted at 6.5% moisture content with 100 kPa confining pressure
for (a) stress sequence 1 and (b) stress sequence 2 ............................................. B-14 
Figure B. 15: Resilient strain measured in RLT tests varying with number of cycles of
specimens compacted at 6.5% moisture content with 100 kPa confining pressure
for (a) stress sequence 3 and (b) stress sequence 4 ............................................. B-15 
Figure B. 16: Resilient strain measured in RLT tests varying with number of cycles of
specimens compacted at 6.5% moisture content with 100 kPa confining pressure
for stress sequence 5............................................................................................ B-16 
Figure B. 17: Resilient strain measured in RLT tests varying with number of cycles of
specimens compacted at 6.5% moisture content with 140 kPa confining pressure
for (a) stress sequence 1 and (b) stress sequence 2 ............................................. B-17 
Figure B. 18: Resilient strain measured in RLT tests varying with number of cycles of
specimens compacted at 6.5% moisture content with 140 kPa confining pressure
for (a) stress sequence 3 and (b) stress sequence 4 ............................................. B-18 
Figure B. 19: Resilient strain measured in RLT tests varying with number of cycles of
specimens compacted at 6.5% moisture content with 140 kPa confining pressure
for stress sequence 5............................................................................................ B-19 
Figure B. 20: Resilient strain measured in RLT tests varying with number of cycles of
specimens compacted at 8% moisture content with 50 kPa confining pressure for
(a) stress sequence 1 (b) stress sequence 2.......................................................... B-20 
Figure B. 21: Resilient strain measured in RLT tests varying with number of cycles of
specimens compacted at 8% moisture content with 50 kPa confining pressure for
(a) stress sequence 3 (b) stress sequence 4.......................................................... B-21 

xxxiii
Figure B. 22: Resilient strain measured in RLT tests varying with number of cycles of
specimens compacted at 8% moisture content with 50 kPa confining pressure for
stress sequence 5 ................................................................................................. B-22 
Figure B. 23: Resilient strain measured in RLT tests varying with number of cycles of
specimens compacted at 6.5% moisture content with 50 kPa confining pressure for
(a) stress sequence 1 (b) stress sequence 2.......................................................... B-23 
Figure B. 24: Resilient strain measured in RLT tests varying with number of cycles of
specimens compacted at 6.5% moisture content with 50 kPa confining pressure for
(a) stress sequence 3 (b) stress sequence 4.......................................................... B-24 
Figure B. 25: Resilient strain measured in RLT tests varying with number of cycles of
specimens compacted at 6.5% moisture content with 50 kPa confining pressure for
stress sequence 5 ................................................................................................. B-25 
Figure B. 26: Resilient strain measured in RLT tests varying with number of cycles of
specimens compacted at 5% moisture content with 50 kPa confining pressure for
(a) stress sequence 1 (b) stress sequence 2.......................................................... B-26 
Figure B. 27: Resilient strain measured in RLT tests varying with number of cycles of
specimens compacted at 5% moisture content with 50 kPa confining pressure for
(a) stress sequence 3 (b) stress sequence 4.......................................................... B-27 
Figure B. 28: Resilient strain measured in RLT tests varying with number of cycles of
specimens compacted at 5% moisture content with 50 kPa confining pressure for
stress sequence 5 ................................................................................................. B-28 
Figure B. 29: Permanent strains measured in RLT tests varying with number of cycles
of specimens compacted at 8% moisture content with (a) 50 kPa confining pressure
and (b) 100 kPa confining pressure ..................................................................... B-29 
Figure B. 30: Permanent strains measured in RLT tests varying with number of cycles
of specimens compacted at 8% moisture content with (a) 140 kPa confining
pressure and (b) 180 kPa confining pressure ...................................................... B-30 
Figure B. 31: Permanent strains measured in RLT tests varying with number of cycles
of specimens compacted at 8% moisture content with (a) 50 kPa confining pressure
and (b) 100 kPa confining pressure ..................................................................... B-31 
Figure B. 32: Permanent strains measured in RLT tests varying with number of cycles
of specimens compacted at 8% moisture content with 140 kPa confining pressure
............................................................................................................................. B-32 

xxxiv
List of Tables
Table 2.1: Level-1 inputs from material testings ............................................................ 12 
Table 2.2: Level-2 inputs from relationship with other material properties ................... 12 
Table 2.3: Presumptive values for Level-3 inputs .......................................................... 13 
Table 2.4: Typical chemical compositions of Portland cement, slag and fly ash (Smith
and Hansen 2003).................................................................................................... 23 
Table 2.5: Properties of lime (AustSTAB 2010) ............................................................ 24 
Table 2.6: Categorisation based on 28-day strength ....................................................... 27 
Table 2.7: UCS used for modulus calculation for level-2 inputs in AASHTO (2008) ... 29 
Table 2.8: Maximum allowable weight loss % ............................................................... 34 
Table 2.9: Minimum compressive strength requirements for lime-pozzolan stabilised
soils ......................................................................................................................... 37 
Table 2.10: Proposed relationship between DVs and field performance (Scullion and
Saarenketo 1997)..................................................................................................... 41 
Table 3.1: Breakdown of thesis chapters, experiments and objectives ........................... 53 
Table 3.2: Impacts of gradation on MoR of lightly stabilised granular material ............ 56 
Table 3.3: Impacts of gradation on SSM of lightly stabilised granular material ............ 57 
Table 3.4: Physical properties of Queensland subgrade material ................................... 63 
Table 3.5: Categories of stabilisation .............................................................................. 65 
Table 3.6: Binders and binder percentages used for different tests ................................ 65 
Table 3.7: Performance comparison of GB cement against AS 3972 requirements ....... 69 
Table 3.8: Chemical composition of fly ash used for experiments ................................. 70 
Table 4.1: Testing program for moisture and durability test .......................................... 80 
Table 4.2: Compaction efforts to achieve maximum dry density ................................... 81 
Table 4.3: Additional TST program ................................................................................ 86 
Table 4.4: Details of TST methodologies obtained from published literature ................ 93 
Table 4.5: Experimental results from different moisture and durability tests................. 97 
Table 4.6: DVs measured with different moulding moisture and binder contents ....... 103 
Table 4.7: DVs measured for different binder contents ................................................ 108 
Table 4.8: The relationship between different durability tests...................................... 119 
Table 4.9: Paired t-test for sample means of VST and F-t durability test..................... 119 

xxxv
Table 4.10: Paired t-test for sample means of VST and 7-day UCS test ...................... 120 
Table 4.11: Paired t-test for sample means of 7-day UCS and f-t durability test ......... 120 
Table 5.1: Mr test of lightly stabilised granular material based on NCHRP testing
protocol ................................................................................................................. 155 
Table 5.2: Mr test of unbound granular material based on NCHRP testing protocol ... 156 
Table 5.3: Two different Mr tests of lightly stabilised material based on NCHRP (2004)
............................................................................................................................... 159 
Table 5.4: Mr tests of lightly stabilised material based on AASHTO testing standard 161 
Table 5.5: Selected stress combinations for Mr testing program .................................. 162 
Table 5.6: Stress combinations for PD testing .............................................................. 163 
Table 5.7: Mr versus loading frequencies ..................................................................... 179 
Table 5.8: Calculated Mr based on regression equation................................................ 180 
Table 5.9: Testing program for UCS testing of materials stabilised with CF binder.... 185 
Table 5.10 Variations in UCS with number of f-t cycles for QM stabilised with 1.5%
and 3% CF binders ................................................................................................ 187 
Table 5.11: Variations in UCS with number of f-t cycles for specimens cured for 7 days
............................................................................................................................... 188 
Table 5.12: UCS after different numbers of f-t cycles for QM and CM specimens cured
for 28 days ............................................................................................................. 190 
Table 5.13: UCS after different numbers of f-t cycles for QM and CM specimens tested
using f-t procedure 2 and cured at 400C ................................................................ 192 
Table 5.14: UCS after different numbers of f-t cycles for QM and CM specimens tested
using f-t procedure 2 ............................................................................................. 193 
Table 5.15: Physical properties determined from sieve analysis and MDD ................. 201 
Table 5.16: Program for RLT testing of materials stabilised with 1.5% CF binder ..... 203 
Table 5.17: Mr of QM stabilised with 1.5% CF binder, cured for 7 days and subjected to
f-t procedure 1 ....................................................................................................... 205 
Table 5.18: Mr of CM stabilised with 1.5% CF binder, cured for 7 days and .............. 206 
Table 5.19: Mr of QM stabilised with 1.5% CF binder, cured for 28 days and subjected
to f-t procedure 1 ................................................................................................... 208 
Table 5.20: Mr of CM stabilised with 1.5% CF binder, cured for 28 days and subjected
to f-t procedure 1 ................................................................................................... 210 
Table 5.21: Mr of CM stabilised with 1.5% CF binder, cured for 7 days and subjected to
f-t procedure 2 ....................................................................................................... 212 

xxxvi
Table 5.22: Mr of CM stabilised with 1.5% CF binder, cured for 7 days at 40oC and
subjected to f-t procedure 2 ................................................................................... 214 
Table 5.23: Regression parameters of two empirical equations correlating PD with
number of load cycles ........................................................................................... 224 
Table 5.24: Classification of PD response .................................................................... 231 
Table 5.25: Residual UCS ............................................................................................. 233 
Table 5.26: Testing Program for flexural test ............................................................... 237 
Table 5.27: Variations in MoR and flexural modulus with number of f-t cycles for QM
stabilised with 1.5% CF binder ............................................................................. 238 
Table 5.28: Variations in MoR and flexural modulus with number of f-t cycles for QM
stabilised with 3% CF binder ................................................................................ 239 
Table 5.29: Testing program for QM stabilised with CF binder subjected to w-d cycles
............................................................................................................................... 245 
Table 5.30: UCS obtained after different numbers of w-d cycles................................. 249 
Table 5.31: Specimens subjected to 1 w-d cycle and cured for different numbers of days
............................................................................................................................... 253 
Table 5.32: Resilient modulus of stabilised specimens with different numbers of w-d
cycles ..................................................................................................................... 256 
Table 5.33: Calculated Mr based on regression equation.............................................. 257 
Table 5.34: MoR and SSM for different numbers of w-d cycles .................................. 261 
Table 6.1: RLT testing program .................................................................................... 271 
Table 6.2: Compaction efforts....................................................................................... 272 
Table 6.3: Stress combinations for multi-stage permanent deformation testing ........... 272 
Table 6.4: Classification of shakedown ranges of specimens compacted with 8%
moisture content and stabilised with 0.5% CF ...................................................... 275 
Table 6.5: Classification of shakedown ranges of specimens compacted with 6.5%
moisture content and stabilised with 0.5% CF ...................................................... 276 
Table 6.6: Classification of shakedown ranges of specimens compacted with 8.0%,
6.5% and 5.0% moisture contents and stabilised with 1.5% CF ........................... 276 
Table 6.7: Shakedown limit equations .......................................................................... 280 
Table 6.8: Resilient modulus obtained for CM ............................................................. 289 
Table 6.9: Regression coefficients obtained for 8% moisture content.......................... 292 
Table 6.10: Regression coefficients obtained for 6.5% moisture content..................... 293 

xxxvii
Table 6.11: Regression constant obtained for b1........................................................... 295 
Table 6.12: Regression constant obtained for exponential and power equations ......... 295 
Table 7.1: Applied loading pressures and numbers of cycles ....................................... 305 
Table 7.2: Comparisons of different types of soil moisture sensors ............................. 312 
Table 7.3: MP306 specification .................................................................................... 313 
Table 7.4: Requirements for subgrade and base materials ............................................ 317 
Table 7.5: Comparisons of UCS obtained from two different curing procedures ........ 320 
Table 7.6: Cyclic loading maximum and minimum pressures...................................... 324 
Table 7.7: RLT testing program .................................................................................... 325 
Table 7.8: Proportion of deformation of base and subgrade ......................................... 329 
Table 7.9: Back-calculated elastic moduli matching resilient deformations ............... 345 
Table 7.10: Input parameters required for cross-anisotropic analysis .......................... 347 
Table 7.11: Back-calculated cross-anisotropic properties ............................................ 347 
Table 7.12: Comparison of stress values at interface and mid-depth of base material . 352 
Table 7.13: Horizontal pressures at boundary wall and 184 mm from centre .............. 354 
Table 7.14: RLT test regression coefficients ................................................................ 358 
Table 7.15: Representative uniform pressure distributions from cross-anisotropic
analysis and comparison of regression coefficients .............................................. 362 

xxxviii
List of Abbreviations
AASHTO American Association of State Highway Transportation Officials
APM Accelerated pavement model
ARRB Australian Road Research Board
CF Cement-flyash
CKD cement kiln dust
CM Canberra granular material
DPP Dry power polymers
DPP Dry powder polymers
DSM dynamic stiffness modulus
DV Dielectric value
EICM Enhanced Integrated Climatic Model
EPC Earth Pressure Cell
FE Finite element
FEA Finite element analysis
FEM Finite element model
FWD Falling weight deflectometer
GB General Blended
GGBFS ground granulated blast furnace slag
GMC gravimetric moisture content
GP General Purpose
GWT groundwater table
HMA hot mix asphalt
HVS Heavy vehicle simulators
IDT indirect diametrical tensile
LVDT Linear variable differential transformers
MDD Maximum dry density
MEPD mechanical empirical pavement design
MoR Modulus of rupture
MoR Modulus of Rupture
MP Moisture probe
OMC Optimum moisture content
PCC Portland cement concrete
PCL Plastic creep limit
PCMC post compaction moisture content

xxxix
PD Permanent deformation
PI Plasticity index
PSL Plastic shake down limit
QM Queensland granular material
RLT Repeated load triaxial
SL Slag-lime
SMM soil moisture meter
SSM static stiffness modulus
STR standing wave ratio
TDR Time domain reflectometry
TMI Thornthwaite moisture index
TST Tube suction test
UCS/qu Unconfined compressive strength
VMC Volumetric moisture content
VST vacuum saturation test

xl
List of Symbols

N’ Allowable number of repetitions of load

 Vertical strain at top of subgrade

 a (n) Permanent deformation of a layer/sublayer

n Number of traffic repetitions

 GB Calibration factor

 , , 0 Material parameters

H Thickness of layer/sublayer

r Resilient strain

v Average vertical resilient strain in layer/sublayer

 Tensile strain produced by the load

Vb Percentage of volume of bitumen in asphalt (%)

Smix Asphalt modulus (MPa)

RF Reliability factor for asphalt fatigue

Ein Initial flexural modulus

Mr Resilient modulus

∆C Change in crack depth due to a cooling cycle

(∆K) Change in stress intensity factor due to cooling cycle

xli
A, n Fracture parameters for HMA mixture

EFlex Flexural modulus

k Constant value between 1000 to 1250

r Relative permittivity

o Permittivity of vacuum

D Relative dielectric permittivity

C Change in capacitance measured by probe

Ca Active capacitance of probe

E(t) Modulus of elasticity at time t

w Mixture density

fc Compressive strength at time t

d Axial deviatoric stress

r Axial recoverable strain

E Flex , 28 Flexural modulus of field beams at 28 days of moist curing

L Span length

B Average width of the specimens

H Average depth of the specimens

Pmax the maximum applied load in N

xlii
P Load

ν Poisson’s ratio

' Effective stress

 Total stress

Xw Bishop’s effective stress parameter

Ua, Uw Pore air and pore-water pressure respectively

ep Permanent strain

N Number of load cycles

a1, a2, b1,b2 Regression parameters

p Bulk stress

y, ɷ Regression constants

ep Permanent deformation

a, b Regression Parameters

Gxy Shear modulus

E Elastic modulus

qu Unconfined compressive strength

xliii
1 Chapter 1

Introduction

1.1 General Overview

The flexible pavement with thin asphalt layer requires unbound granular base course to
act as a structural layer and distributes the traffic load to the subgrade. The aggregates in
an unbound granular base generally consist of crushed stone or quarry rock, sand, gravel
and/or any other durable materials of mineral origin. Limited availability of good-
quality aggregates and costs of transporting materials from its origin to working site
resulted in materials of sub-standard quality being used for stabilisation. Moreover, the
introduction of heavy vehicles with increased tyre pressure has demanded pavement
engineers to build roads which can resist huge axle pressures. Improving the mechanical
properties of pavement materials through stabilisation to build roads resistant to ever-
increasing axle pressures has been practised for many years.

Stabilisation is performed in both new and rehabilitated pavements and, in the latter, an
existing material can be stabilised and used to obtain cost benefits (Smith and Vorobieff
2007). The advantages of rehabilitation using stabilisation can be categorised as cost,
social and environmental and it is even used to transform crushed concrete obtained
from demolished buildings into a good pavement material (Marradi and Fausto 2008).

A pavement’s ability to withstand traffic loading depends on the strength and stiffness
of its layers which are considerably affected by environmentally driven variables such
as temperature, moisture, wet-dry (w-d) and freeze-thaw (f-t). Chemical stabilisation of
granular materials is in practice for many years to obtain enhanced mechanical
properties, but environmental factors posing threats to materials’ performance.

Durability, which is defined as the capability of a material to retain stability and


integrity over years of exposure to the destructive forces of weathering studies, show
long-term performance of a chemical stabilization (Dempsey and Thompson 1968).
Chapter 1 Introduction

It, ensures the long-term performance of a pavement, is one of the most important
aspects of a granular base material. Assessing the durability of a stabilised material is
essential and is performed in the mix design stage. A typical design of a soil-cement
mixture uses strength and durability as the key engineering properties to evaluate the
quality of the mixture. The Unconfined compressive strength (UCS) test helps in the
design of a mix that meets the minimum strength requirement. However, there is no
uniform approach for assessing the durability of stabilised materials as the standard
testing procedure is not well accepted.

Considering a lightly stabilised granular material, the most difficult aspect is ensuring
the sufficient permanency of stabilisation throughout a pavement’s life. Unlike
cemented materials, for lightly stabilised materials, neither strength nor durability
requirements are imposed because the purpose of light stabilisation can be to improve
some other property of the material (for example, permeability). Therefore, emphasising
durability has never been a concern and the applicability of existing durability test
methods of cemented material to lightly stabilised materials is unknown. However, as
the improved elastic properties of lightly stabilised materials are used in pavement
design, understanding environmental effects is necessary.

Therefore, in this thesis, assessing durability of lightly stabilized granular materials are
performed using different laboratory tests focusing on different mechanical properties of
the material. Different experimental procedures were presented to study the f-t, w-d
cycles and its sensitiveness to moisture content changes. Moreover, different variables
and experimental procedures were compared to obtain a better laboratory test to
evaluate the durability of lightly stabilized granular materials.

In addition to the durability study, deformation characteristics of lightly stabilised


granular materials using repeated load triaxial (RLT) and accelerated pavement model
(APM) testing also performed to increase the understanding of the materials’ rutting
behaviour. Shake down theory was mainly used to classify the materials responses and
to define equations of plastic shake down limit (PSL) and plastic creep limit (PCL).

2
Chapter 1 Introduction

1.2 Objectives
The thesis presents comprehensive investigation on durability, moisture susceptibility
and deformation characterisation of lightly stabilised granular material using element as
well as the accelerated pavement model testing. The following objectives are identified
and investigated in this thesis.

 To present a comprehensive literature review on pavement design practices,


granular material stabilisation, characterisation, durability and moisture study on
cementitious stabilised materials.
 To assess the durability against f-t cycles using f-t durability test and to compare
outcome with three other laboratory tests that could also be suitable durability
testing procedures.
 To perform additional tests using tube suction test (TST) investigating the
impacts on moisture content and binder content on dielectric value (DV). Also
comparing DVs measured from two different TST setups and emphasising the
need for clean porous disk and smooth surface for consistence DV measurement.
 To propose a suitable stress combinations for performing resilient modulus (Mr)
test on lightly stabilised granular material and to investigate the effects of
loading frequency and stress values on Mr.
 To perform UCS test to assess the impacts of type of granular material, number
of curing days, binder percentage, curing method, number of f-t cycles and type
of f-t procedures on UCS values obtained after subjecting the specimens to f-t
cycles.
 Extend the investigation to RLT test and study the resilient and rutting
responses. Interpreting the results along with their stress values to determine the
effect of stress on deformation responses.
 To study the changes in UCS, Mr and permanent deformation behaviour of
lightly stabilised material subjected to w-d cycles and compare their outcomes.
 To evaluate the f-t and w-d effects on flexural properties of the lightly stabilised
material using flexural beam test. Extend the investigation to assess the impacts
of number of f-t, w-d cycles and binder content on static stiffness modulus

3
Chapter 1 Introduction

(SSM) and modulus of rupture (MoR) and to propose relationship between


them.
 To perform RLT test to investigate the permanent deformation (PD) response of
lightly stabilised granular material. Classify the material behaviour using
shakedown theory and defining the PSL and PCL for the material.
 To model the PD using log based model and to establish a relationship between
PD and number of load cycles in terms of confining pressure and bulk stresses.
 To study the relationship between permanent and elastic deformations and
influence of molding moisture content on Mr.
 To perform APM test for charactering the deformation responses and to back-
calculate the materials properties using finite element (FE) analysis.

1.3 Contribution to new knowledge


The following new knowledge obtained from the thesis expected to contribute the
understanding of the lightly stabilised granular materials.

 The f-t and w-d durability studies will show the lightly stabilised granular
materials’ ability to resist f-t and cycles.
 The effects of lightly stabilised materials’ properties such as UCS, flexural and
resilient modulus to moisture content will be established and the effects of
moisture content on f-t effects will be shown in series of UCS test.
 Comparison between different f-t laboratory testing procedures such as f-t
durability test, vacuum saturation test (VST) and 7-day UCS will help to form a
relationship between them and to identify the better laboratory test to evaluate
the f-t durability of the material.
 The TST on lightly stabilised granular material will help to understand its
suitability to assess the f-t durability and moisture sensitivity of the material.
The impacts of moisture, binder content, type of porous disk, smoothness of the
surface and TST setup will reveal their influence on DVs and the importance of
controlling some of these variables to obtain consistence results.
 The preliminary study will suggest the suitable stress values that could be used
in the Mr testing procedure to obtain the Mr of lightly stabilised granular

4
Chapter 1 Introduction

materials. The Mr test with different combinations of stress and frequency will
give their influence on Mr of lightly stabilised granular material.
 The UCS test on specimens subjected to f-t cycles will evaluate the materials
ability to resist the f-t cycles and the changes in its strength values with number
of f-t cycles. It will also show the effects of type of granular material, number of
curing days, binder percentage, curing method and type of f-t procedures on
UCS values obtained after f-t cycles.
 The Mr test will show the effects of f-t cycles on Mr and assess the influence of
all the variables that will be tested in UCS test.
 The w-d durability tests will show the materials’ ability to resist w-d cycles.
 Analysing the PD obtained from RLT tests of lightly stabilised granular material
using shakedown theory will be shown.
 Modelling the PD using log based model and defining a predictive equation in
terms of number of load cycles, bulk stress and confining pressure.
 The APM test performed on lightly stabilised material will increase the
understanding of deformation responses of the material. The continuous
measurement of moisture content and the pressure is expected to give the
moisture and pressure distribution within the pavement model.
 The developed FE model will show a method to back-calculate the Mr of base
and subgrade materials. The comparison of isotropic and cross-anisotropic
analysis will show differences in the outcome and the error that could result
from isotropic analysis.

1.4 Organisation of Thesis

The thesis comprised of eight chapters including this introductory chapter. The contents
of the other chapters are summarised below.

Chapter 2 presents a comprehensive literature study focussing on pavement design


concepts, stabilisation of pavement materials, characterisation and durability of
stabilised material. The back ground study on pavement design compare the past and
present pavement design methods and emphasise the importance of incorporating

5
Chapter 1 Introduction

environmental variables in the mechanical empirical (M-E) analysis and design.


Durability tests cover different laboratory tests that were performed on stabilised
materials to obtain resistance against f-t and w-d cycles.

Chapter 3 focusses on research methodologies that are used to achieve the objectives
outlined earlier in this chapter. It provides a framework of different testing procedures
and includes a brief outline of each them. It also presents physical and chemical
properties of granular materials and binders that were used for the experiments.

Chapter 4 presents four laboratory tests such as the f-t durability, vacuum saturation, 7-
day UCS tests and TST to evaluate the durability and moisture susceptibility of the
lightly stabilised granular material. TST is a non-destructive test which measures the
DV of a material whereas other three measure the UCS. Individual test results evaluated
the material’s ability to resist f-t cycles and its sensitivity to moisture whereas the
comparison between different tests asses the benefits and drawbacks of each test
methods and present the relationship among them.

Chapter 5 is also dedicated to the durability of lightly stabilised granular material.


However, it examined the factors that influence the f-t testing procedure using three
element testings such as UCS test, RLT test and flexural beam test. Along with the
different element testings, variables such as the type of granular material, number of
curing days, binder percentage, curing method and type of f-t procedures were
considered for obtaining better laboratory test to evaluate the durability of lightly
stabilised granular materials. The durability of lightly stabilised granular material
against w-d cycles were also presented in this chapter. Other than the durability tests,
suitable stress sequences that can be used to perform the Mr test on the lightly stabilised
granular material were also discussed in this chapter.

Chapter 6 examines the PD characterisation of lightly stabilised granular material using


RLT test. The PD response obtained from the granular material stabilised with 0.5% of
cement-flyash were analysed using shakedown theory. Initially, the classification of
material responses were outlined, which then followed by the equations defining PSL
and PCL were established in this chapter. It also presents the relationship between
permanent and elastic deformations and influence of moisture on Mr.

6
Chapter 1 Introduction

Chapter 7 is devoted to accelerated pavement model (APM) testing for charactering


material’s deformation behaviour. Moisture and pressures sensors were also installed
for continuous measurement of moisture and pressure respectively. Measured
deformations were used to back-calculation of elastic moduli using finite element model
(FEM) using isotropic and cross-anisotropic analysis. Finally, using RLT test data and
Finite element analysis (FEA), relationship between regression coefficients obtained
from accumulated permanent deformation was developed in terms of bulk-stress.

Chapter 8 summaries all the outcomes obtained from the experimental study and also
recommends potential research areas that could be explored in the future.

7
2 Chapter 2

Literature Review

2.1 Introduction

A well-developed and maintained road network is one of the essential requirements of


every country for the transfer of goods and services. A country not only allocates a
substantial investment in the development of sustainable road networks, it also spends
huge amounts of money on their maintenance. Obtaining good returns on these
investments, necessitates continuous improvements in material characterisation,
construction practices and pavement design.

The primary function of a pavement comprised of more than one layer is to distribute
applied vehicle loads to the subgrade while providing acceptable riding quality. A
pavement’s layers are designed in such a way that the pressures from wheel loads are
sufficiently distributed to enable the pavement to operate for its intended lifetime. There
are two main categories of pavements, flexible and rigid.

 Flexible pavement (Figure 2.1): has a bitumen-bonded surfacing or asphalt


concrete and a road base constructed on top of its subgrade as its foundation. As
wheel loads are spread across a wide area and less pressure is transferred to the
bottom layers, the top layers are stronger than the bottom ones.

 Rigid pavement (Figure 2.2): has a concrete surface slab which can be un-
reinforced, joint reinforced or continuously reinforced and relies heavily on its
flexural strength to resist wheel loads. In pavement analysis, the plate theory is
used to analyse a concrete slab, the major distress failure mode of which is
fatigue cracking.
Chapter 2 Literature Review

The types of materials available that can be used to form these pavement layers can be
categorised as follows.

 Unbound granular material


Unbound materials can be natural gravels, crushed rocks, sand, etc., and are used mainly
for constructing bases or sub-bases. The performance of an unbound pavement is
influenced by its moisture content or, more specifically, the degree of saturation of its
materials. An unbound granular material is characterised based on its resilient modulus
(Mr) which can be obtained using repeated load triaxial (RLT) testing in a laboratory.

 Modified granular material


Materials with small percentages of binders or modified by adding other granular
materials are classified as modified materials. Similar to unbound materials, they are
used to construct base or sub-base layers and are characterised using the resilient
modulus (Mr).

 Cemented material
A stabilised granular material with substantial flexural strength is classified as a
cemented material and is used to construct rigid pavements, with laboratory flexural
beam testing conducted to obtain its flexural properties. Shrinkage cracking due to
temperature changes and flexural cracking due to traffic loading are the two main
distress modes of this material.

 Asphalt
Asphalt is a mixture of a bituminous binder and other aggregates, such as stone, sand or
gravel. The quality of an asphalt layer depends on the quality of the bitumen binder and
it is more common to add polymers to the binder to obtain better performance.
Permanent deformation (PD) and tensile strain are the two limiting factors of asphalt
layers.

9
Chapter 2 Literature Review

Surface course
Base

Sub-base

Subgrade

Figure 2.1: Typical flexible pavement

Concrete base

Sub-base

Subgrade

Figure 2.2: Typical rigid pavement

2.2 Mechanical –Empirical (M-E) Analysis and Design

Pavement analysis and design is the process of determining the appropriate thicknesses
of pavement layers. A combination of asphalt, base/sub-base and subgrade layers is
expected to resist the stress generated by wheel loads under varying environmental
conditions. Such analysis requires inputs from different sources, such as pavement
loadings, traffic data, climatic data, the materials’ mechanical properties and
performance models. During the structural design of a pavement, the layers’ thicknesses
are selected in such a way that the stresses due to traffic remain below the allowable

10
Chapter 2 Literature Review

limits, with the environmental conditions the pavement would experience during its
lifetime incorporated.

The American Association of State Highway Transportation Officials (AASHTO)


published series of design guides for pavement structures in 1972, 1986 and 1993. All
were based on an AASHO road test carried out in the late 1950s in some parts of
America (Carvalho and Schwartz 2006). These guidelines are empirical in nature, with
their analyses and designs based on experiments and experiences.

As an empirical approach is site-specific, extrapolating its results is likely to result in


erroneous outcomes. As existing empirical correlations were derived from specific
loading and environmental conditions, their applicability to different sets of conditions
is questionable. Over the years, the introduction of heavy vehicles, use of new materials,
new construction techniques and changing climatic conditions have all raised questions
about the validity of relying on an empirical approach for pavement design. Therefore,
it is necessary to design pavements regionally and incorporate local conditions rather
than relying on generalised pavement design methods.

AASHTO with the National Cooperative Highway Research Program (NCHRP) and
initiated a projects called NCHRP 1-37A and 1-40A with the objective of introducing a
mechanistic-empirical (M-E) pavement design approach and published its interim guide
AASHTO (2008) which, as the name implies, has both mechanistic components and
empirical elements. In this M-E design, structural responses are mechanistically
calculated using a simple multi-layer or finite element program, with the obtained
outputs then used in empirically derived performance models to determine the allowable
number of passes.

The AASHTO (2008), interim guide provides a designer with the discretion to select the
levels of inputs required for the design. In a detail design, mechanical empirical
pavement design (MEPD) requires details of the materials’ properties, which may
require many laboratory tests to determine, and collections of data regarding
environmental and traffic conditions. Unless a project can devote substantial resources
to conducting these tests and data collections, two other levels of inputs can be selected.

11
Chapter 2 Literature Review

Tables 2.1 to 2.3 illustrate the different levels of inputs suggested for chemically
stabilised materials. For level 2, as shown in Table 2.2, the modulus and resilient
properties are related to the unconfined compressive strength (UCS) which can be easily
obtained. However, for level 1, except for a few materials, it is necessary to perform
appropriate laboratory testing which is relatively more complex than UCS testing but
improves accuracy. Although level-3 input values can be used directly from the table,
there is a good chance these may vary from the actual values. Therefore, hierarchical
levels of inputs serve different purposes and provide a designer with the freedom of
choice based on the limitation and importance of each parameter.

Table 2.1: Level-1 inputs from material testings

Material Mechanical property Test method


Lean concrete and cement-
Elastic modulus ASTM-C469 (2014)
treated aggregate
Lime-cement-fly ash Elastic modulus Levels 2 or 3

Soil cement Elastic modulus Levels 2 or 3


AASHTO-T307-99
Lime-stabilised soils Resilient modulus
(2007)

Table 2.2: Level-2 inputs from relationship with other material properties

Material Relationship for modulus Test method


Lean concrete and cement- AASHTO-T22
E = 57000(fc’)0.5
treated aggregate (2014)
Open-graded cement-
Use input level 3 None
stabilised aggregate
ASTM-C593
Lime-cement-fly ash E = 500+qu
(2011)
ASTM-D1633
Soil cement E = 1200(qu)
(2000)
ASTM-D5102-09
Lime-stabilised soils Mr = 0.124(qu)+9.98
(2009)

* E, Mr, fc’, qu in psi

12
Chapter 2 Literature Review

Table 2.3: Presumptive values for Level-3 inputs

psi/MPa

Lean concrete (E) 2,000,000/13,790


Cement-stabilised aggregate
1,000,000/6,895
(E)
Open-graded cement-
750,000/5,171
stabilised aggregate
Soil cement (E) 500,000/345

Lime-cement-fly ash (E) 1,500,000/10,342

Lime-stabilised soils (Mr) 45,000/3,102

2.3 Enhanced Integrated Climatic Model (EICM)

As the properties of the materials in pavements are sensitive to environmental variables,


measures are taken to limit their effects by selecting appropriate design parameters. The
new MEPD guide introduced the enhanced integrated climatic model (EICM) which
breaks down the layered pavement design process into two stages. In the first stage, the
environmental effects are predicted using the EICM and, in the second, the outcome of
such an analysis is used to quantify changes in the mechanical properties of the
pavement materials.

Unlike previous design methods, the MEPD guide considers environmental factors a
serious threat and the EICM plays a vital role in defining the material properties in it;
for example, based on different input values, particular water table and soil water
characteristic model, the EICM can predict the equilibrium moisture content that could
prevail for most of a pavement’s design life. This is then used to select the appropriate
Mr which is a function of the moisture content.

Before implementing the EICM, it’s good practice to validate its outputs against field
data pertaining to the particular region (Ahmed et al. 2005; Bulut et al. 2013; NCHRP
2008; Quintero 2007). However, this requires extensive field output data as well as the
parameters required to run the EICM.

13
Chapter 2 Literature Review

2.4 Parameters Required to Run Enhanced Integrated Climatic


Model

The parameters required to run the EICM depend on the accuracy of the analysis. For a
level-1 analysis, more detailed parameters are required than for a level 3 in which the
existing in-built relationship in the EICM based on basic soil properties can be used.
NCHRP (2008) provides details of all the parameters required for an analysis in four
tables ,a few of which are discussed below.

 Input parameters for model initialisation


Climatic measurements taken on an hourly basis, that is, the air temperature (F), wind
speed (mi/h), percentage of sunshine (%), precipitation (in), relative humidity (RH) (%),
depth of the groundwater table (GWT), Thornthwaite moisture index (TMI) and soil
suction profile.

 Input parameters required for natural, in-situ material properties


The specific gravity, saturated hydraulic conductivity, maximum dry unit weight, dry
thermal conductivity, heat capacity, plasticity index, optimum gravimetric water content
and sieve analysis results. Similarly, there are several input material properties required
for compacted materials such as hot mix asphalt (HMA) and Portland cement concrete
(PCC). For a level-1 analysis, most of the input parameters needed are determined using
laboratory experiments.

2.5 Outputs from Enhanced Integrated Climatic Model (EICM)

The EICM produces the following outputs at the end of an analysis which can be used
to quantify their effects on mechanical properties. Figure 2.3 shows the whole process
involved in an EICM analysis.

1. Temperature distribution profile over time


2. Moisture content measurements profile over time
3. Frost penetration depth
4. CBR

14
Chapter 2 Literature Review

Figure 2.3: EICM scheme (Quintero 2007)

15
Chapter 2 Literature Review

2.6 Pavement Distress Modes and Performance Models

Each pavement layer has one or more distress mechanisms and, when their allowable
limits are reached, the pavement becomes unfavourable for moving traffic. Performance
models for estimating the numbers of allowable passes are designed based on these
failure criteria. Generally, rutting and fatigue criteria are used to determine the
allowable number of load repetitions and there are numerous models based primarily on
empirical approaches available for predicting allowable number of load cycles of which
some of them are presented in this section.

 Rutting

Rutting is very common in asphalt, unbound and subgrade layers as a result of the
continuous accumulation of PD. This is the principal distress mechanism for subgrade
materials and one of the two main ones for asphalt layers. Rutting in asphalt can be seen
in the longitudinal direction under the wheel paths which is a result of continuous
accumulation of PD. There are different rutting criteria available for predictions, with
Austroads (2010a) suggesting equation 2.1 for subgrade materials but not considering
asphalt rutting for design purposes. However, rutting in asphalt, which generally
develops during hot seasons when the asphalt is softer than at other times, can be
controlled using a good-quality asphalt binder.

7
 9300 
N   Eq. (2.1)
  
where:
N = allowable number of repetitions of load
 = vertical strain at top of subgrade

In AASHTO (2008), each pavement layer is sub-layered to calculate PDs, including that
of an unbound granular base. The following equation is used for both base and subgrade
soils.



   n 
 a ( N )   GB  0 e   vh Eq. (2.2)
r 

16
Chapter 2 Literature Review

where:
 a (n) = PD of a layer/sublayer (in)
n = number of traffic repetitions
 GB = calibration factor
 , , 0 = material parameters
H = thickness of layer/sublayer (in)
r = resilient strain imposed in laboratory test to obtain material
properties listed above (in/in)
v = average vertical resilient strain in layer/sublayer obtained from
primary response model (in/in)

 Fatigue cracking

This is very common in an asphalt layer and cemented materials and is caused by
repeated traffic loadings whereby PD continuously accumulates over a number of
loading cycles, with cracks indicating fatigue failure of the asphalt or cemented layer.
There are numerous empirical models and models based on dissipated energy
approaches available for predicting the allowable number of repetitions (Mallick and El-
Korchi 2013). The asphalt fatigue performance model suggested in Austroads (2010a)
for moderate to heavily trafficked pavements is:

5
 6918(0.856Vb  1.08) 
N '  RF   Eq. (2.3)
S mix 
0.36
 

where:

N’ = allowable number of repetitions of the load

 = tensile strain produced by the load (micro strain)

Vb = percentage of volume of bitumen in asphalt (%)

Smix= asphalt modulus (MPa)

RF = reliability factor for asphalt fatigue

17
Chapter 2 Literature Review

The fatigue criteria defined by Austroads (2010a) for a cemented material is:

0.804
113000 / E in 191
N '  RF[ ] Eq. (2.4)


where:

N’ = allowable number of repetitions of load

µe= tensile strain produced by load (micro strain)

Ein = initial flexural modulus (MPa)

RF = reliability factor

 Thermal cracking

Transverse thermal cracking commonly occurs in asphalt concrete pavements due to


sharp temperature drops or after repeated temperature cycling, particularly in cold
regions. It is classified in three levels by the Strategic Highway Research Program
(1993a) based on its severity as:

1. Low severity: mean crack width less than 6 mm;


2. Moderate severity: mean crack width between 6 mm and 19 mm or less than 19
mm, but adjacent to low severity crack;
3. High severity: mean crack width exceeding 19 mm or less than 19 mm, but
adjacent to moderate to high severity crack.

AASHTO (2008) recommends the following model for estimating a change in crack
depth due to the temperature cycle.

∆C = A(∆K)n Eq. (2.5)

where:

∆C = change in crack depth due to a cooling cycle

18
Chapter 2 Literature Review

(∆K) = change in stress intensity factor due to cooling cycle

A, n = fracture parameters for HMA mixture

There are other types of cracking, such as the reflection of shrinkage cracking, potholes,
bleeding, frost heave/thaw cracking and joint reflection cracking, which may also cause
problems in a pavement.

2.7 Stabilisation of Base Materials

The increasing use of heavy vehicles with multiple axle wheels and increased tyre
pressures which are required to meet the ever-growing demands and needs of society
has put pressure on pavement engineers to build roads which can resist huge axle
pressures. Improving the mechanical properties of pavement materials through
stabilisation to build roads resistant to ever increasing axle pressures has been practised
for many years. Moreover, the limited availability of good quality aggregates and costs
associated with transporting materials from one place to another has resulted in
materials of sub-standard quality being used for stabilisation.

The term ‘stabilisation’ refers to the addition of a liquid or solid substance to a


pavement material to improve its performance and its purpose can range from simply
providing a working platform to significantly improving the mechanical properties of
the material. It also enhances the quality of certain sub-standard materials to make them
suitable for pavement construction (Chakrabatri et al. 2002). Moreover, even good-
quality materials can be stabilised to improve their mechanical properties which can
reduce a pavement’s thickness. Stabilised materials offer many advantages apart from
improving strength, stiffness and durability; for example, using their by-products, such
as fly ash and slag, benefits the environment (Sahu and Gayathri 2014)

There are many ways of classifying the types of stabilisation, cementitious, granular,
bituminous and chemical, as discussed in the following sections.

19
Chapter 2 Literature Review

2.7.1 Cementitious Stabilisation

Cementitious stabilisation uses Portland cement either alone or with other


supplementary materials, such as lime, fly ash or slag. The cementitious binders that
react with water and form a cementitious paste that helps to bind the solid soil particles.
In the presence of a continuous supply of water and suitable temperature, strength gains
occur as the paste hardens which is called hydration. The use of fly ash activates
pozzolanic reactions in the presence of an alkaline to produce cementitious materials.

One important consideration when selecting a particular cementitious binder and its
composition is its setting time. As a cement reaction is an exothermic one in which heat
is dissipated into the environment, the higher the setting time, the better the results. A
high setting time enables better compaction and provides safety against uncertainties
during construction (Smith and Hansen 2003). Therefore, currently, in a cementitious
stabilisation, a particular percentage of Portland cement is replaced by other
supplementary cementitious materials which are slow reactors in order to obtain
extended working hours. Four commonly used cementitious binders are discussed in the
next sections.

2.7.1.1 General Purpose (GP) and General Blended (GB) Cements

While Portland cement remains as the major constituent in GP and GB cements, the
quantities of the other minerals added to them differentiate them into two classes. AS
3972 (AS-3972 (2010) defines GP cement as hydraulic cement containing mainly
Portland cement a with the addition of a single mineral not exceeding 5% or a
combination of minerals not exceeding 7.5% which could be lime, fly ash or slag. In a
GB cement, the amount of Portland cement is limited to 92.5%, and a quantity of fly ash
or ground granulated iron blast furnace slag or a combination of them exceeding 7.5%
can be added as well as up to 10% of amorphous silica at the discretion of the
manufacturer. Including fly ash or slag with Portland cement provides GB cement with
the following advantages over GP cement (Vorobieff 2000).

 As its fly ash and slag are by-products, there are environmental benefits.

20
Chapter 2 Literature Review

 Its low initial reaction allows additional working hours for compaction, ride and
level control. According to Vorobieff (1997b), additional 1-2 hours of working
time can be obtained by using GB cement.

 It is relatively less costly

 Due to its slow setting nature, reflective cracking is avoided.

Due to its abovementioned benefits, GB cement has been identified as the most
favourable option and is successfully used throughout Australia (Vorobieff 1997a). As a
result, manufacturers produce GB cement with specific proportions consisting of fly ash
and blast-furnace slag.

2.7.1.2 Fly Ash

Fly ash is a by-product generated during the combustion of pulverised coal in a boiler in
a power station. As shown in Figure 2.4, its particles are very fine grain and spherical in
shape which helps to increase workability. It is a pozzolanic material defined in the ACI
concrete terminology as “a siliceous or siliceous and aluminous material that in itself
possesses little or no cementitious value but that will, in finely divided form and in the
presence of moisture, chemically react with calcium hydroxide (lime) at ordinary
temperatures to form compounds having cementitious properties; they are both artificial
and natural”(ACI 2010). Its low reaction rate reduces the heat of hydration and the
possibility of drying shrinkage and, when it replaces cement, has a ‘greener’ effect on
the environment and is commonly known as a supplementary cementitious material.

The hydration of cement produces a base medium (a supply of calcium hydroxide) and
activates the pozzolanic reaction of fly ash, as shown in equations (2.6) and (2.7).

3 [Ca(OH)2] + 2 [SiO2] = [3(CaO) 2(SiO2)3(H2O)] Eq. (2.6)

3 [Ca(OH)2] + Al2O3 = [3(CaO) Al2O3 6(H2O)] Eq. (2.7)

21
Chapter 2 Literature Review

In Australia, the fly ash used as a supplementary cementitious material has to comply
with AS-3582.1 (1998) in which the requirements for its grading, alkali and chloride
iron contents, strength and density are mentioned.

Figure 2.4: Scanning electron microscope image of fly ash

2.7.1.3 Slag

Slag is a supplementary cementitious material similar to fly ash that can be used to
partially replace clinker in cement production. AS-3582.2 (2001) defines slag as: “the
glassy granular material resulting from the rapid chilling of molten iron blast-furnace
slag”. It is a by-product of the iron and steel manufacturing processes and produced in
enormous quantities throughout the world. The initial iron manufacturing process in
steel production generates molten iron metal and molten slag, with the slag separated
from the molten metal in liquid form and generally known as blast furnace slag. Then,
high-pressure water is used to produce a sandy material called granulated blast furnace
slag which, when ground to cement fineness is known as ground granulated blast
furnace slag (GGBFS) and used for stabilisation.

22
Chapter 2 Literature Review

The chemical composition of slag is similar to those of cement and lime but it does not
react with water (Table 2.4). As the hydration of Portland cement and lime provides
calcium hydroxide which acts as an activator for the slag to react and produce
cementitious products, it is generally mixed with either GP cement or lime and used for
stabilisation. Moreover, slag give slightly higher strength than the same proportion of
fly ash in a mix (Smith and Hansen 2003)

One of the major benefits of using slag products in a stabilised pavement is its slower
rate of cementitious reaction, less heat generation and less drying shrinkage. Using slag
as a substitute for clinker promotes a ‘greener cement’ as clinker is the most energy-
and greenhouse-intensive component in cement. It’s also cost effective as it is a by-
product of steel production and should meet the requirements stated in for a
supplementary cementitious material for use with Portland cement in road stabilisation

Table 2.4: Typical chemical compositions of Portland cement, slag and fly ash
(Smith and Hansen 2003)

Chemical
Portland cement Slag Fly ash
compound
CaO 64.0 41.0 4.4
SiO2 22.0 32.6 55.0
Al2O3 4.5 12.8 25.0
Fe2O3 3.5 1.3 9.0
MgO 1.4 7.2 1.5
Na2O,K2O 0.7 2.3 1.8
SO3 2.4 0.03 0.4

2.7.1.4 Lime

Quicklime (CaO), lime slurry and hydrated lime (Ca(OH)2) are three forms of lime
available for soil stabilisation and their properties are shown in Table 2.5. Ca(OH)2 is
produced from quicklime lime by adding water either in the field or at the
manufacturing plant. Lime stabilises a material in two ways: in the short term, Ca2+
cations replace other cations and form bigger particles through flocculation and
agglomeration (AustSTAB 2010), with the newly formed soil particles generating

23
Chapter 2 Literature Review

increased soil friction and showing more resistance to loadings; and in the long term, it
creates an alkaline environment for the pozzolanic reaction. Adding lime to the reactive
silica and aluminate present in a soil enables the pozzolanic reaction required to produce
cementitious products. Otherwise, pozzolanic materials, such as fly ash and slag, can be
added to supply the necessary silica and aluminate.

Lime-stabilised soils show less plasticity, higher strength, and greater resistance to
fracture, fatigue and adverse moisture effects than other soils and are much more
effective in moderate to high plastic clays (Little et al. 2002). Lime stabilisation is
shown to be 50 % cheaper than the traditional ground improvement technique
(AustSTAB 2010)

Table 2.5: Properties of lime (AustSTAB 2010)

Hydrated lime Quicklime Lime slurry


Composition Ca(OH)2 CaO Ca(OH)2 slurry
Equivalent Ca(OH)2 1.00 1.32 0.56-0.33
Bulk density (t/m3) 0.45-0.56 0.90-1.30 1.25

2.7.2 Granular Stabilisation

Granular stabilisation is used to improve the quality of the grading of a soil or minimise
the percentage of poor-quality elements in its parent material by adding one or more
materials which could be granular materials or industrial waste or river deposit products
(Austroads 2010a). Granular stabilisation is a low-cost choice for improving the
performances of roads, particularly unsealed low-volume ones in regional areas.
According to Yin (2011), both the lack of knowledge about locally available materials
and the limited money allocated for road maintenance hinder the benefits that could be
obtained from granular stabilisation.

24
Chapter 2 Literature Review

2.7.3 Bituminous Stabilisation

According to Austroads (2010a), bituminous stabilisation is performed using either a


foamed bitumen or bitumen emulsion with cement or lime added to increase the
stiffness of the mix. When hot bitumen has cold water added to it, its volume increases
10 to 15 times and produces foamed bitumen which coats the finer granular materials
and forms a cohesive granular mix which is workable and lasts sufficiently long to
complete compaction. Foamed bitumen stabilisation is best suited for low-plasticity
granular soils and is very effective when compacted on the dry side of the optimum
moisture content (OMC) (Tom Wilmot 2000).

Standard bitumen emulsions are comprised of 40% water and 60% bitumen although,
for a sprayed seal, 70% bitumen and 30% water is becoming popular. They are mixed
with soil and water to obtain a dispersive stabilised mix with increased cohesion and
plasticity.

2.7.4 Chemical Stabilisation

Chemical binders are available in both solid and liquid form and primarily used as dust
suppressants in unsealed roads with cementitious binders (Vorobieff 2000). According
to Vorobieff (2000), chemical binders added to an unsealed road minimise the
generation of dust which has been an issue in such roads next to agricultural land as it
detrimentally affect crops. Moreover, chemical stabilisation minimises the rutting that
occurs after rainfall when the material has a very limited bearing capacity. Polymers,
salts, and organic, ionic and biological binders are some categories of chemical binders
(Austroads 2010a).

Dry powder polymers (DPP) are suitable for water-susceptible gravels that lose strength
when they become wet. Upon water intrusion, the finer particles in the gravel act as
lubricants and allow the bigger particles to slide which causes rutting. DPP are water-
insoluble and produce a less permeable pavement material by creating a hydrophobic
soil matrix between the stones. This softening by the lubricating effect is referred to as
the ‘internal’ waterproofing of fine-grained particles (AustStab 2007). The treated

25
Chapter 2 Literature Review

material remains flexible and less permeable than untreated and acts as a barrier against
moisture for subgrade materials.

2.8 Categorisation of Cementitiously Stabilised Materials

As mentioned in the previous section, cementitious stabilisation is performed by adding


cementitious binders to granular materials. Depending on the type, amount and
enhancing quality of a binder, the stabilised material can be classified in one of three
different categories.

Modified Granular Base Material

The material resulting from the addition of a low percentage of binder is referred to as a
modified material. Its binder content is sufficiently low to not create significant tensile
strength but improves properties such as the plasticity index, shrinkage limit,
permeability and durability. According to Austroads (2010a), a modified material
possesses UCS of between 0.7 MPa and 1.5 MPa and, similar to unbound granular
materials, is characterised based mainly on its Mr for which, in the absence of laboratory
Mr test data, a presumptive value of 1000 MPa can be used.

Lightly Bound Material

Unlike a modified granular material, a lightly bound material possesses considerable


tensile strength but is still characterised based on its Mr. Apart from increasing its
tensile capacity, the purpose of stabilisation could be to improve its other mechanical
properties, such as the Mr or UCS. As it’s likely that a material would lose its tensile
capacity under a typical working load, the Mr is used for design purposes. As shown in
Table 2.6, earlier versions of Austroads classified lightly bound material as having an
UCS between 1 MPa and 4 MPa.

26
Chapter 2 Literature Review

Bound or Cemented Material

A fully bound material, which is also referred to as a cemented material, possesses


significant tensile strength as a result of a higher percentage of binder being added to a
mix. It is characterised based mainly on its flexural properties as its failure distress
mode is fatigue cracking.

Defining a category based on the percentage of binder needs to be avoided as the


enhancing capability of each binder is different and, at the same time, using UCS values
remains subjective. In fact, Austroads (2010a) does not recognise a category for lightly
bound materials, with its earlier version having the categories shown in Table 2.6.
Although the focus of this thesis is mainly on a lightly bound material, some test
combinations use 3% of binder which is likely to cause the material to behave as
cemented.

Table 2.6: Categorisation based on 28-day strength

Category of Design strength Design flexural


Degree of binding
stabilisation (MPa) modulus (MPa)
Modified material UCS1 < 1.0 1000

Austroads (2002) Lightly bound 1< UCS1 < 4 1500 – 3000

Heavily bound UCS1 > 4 5000

2.9 Characterisation of (lightly) Stabilised Materials

The characterisation of a pavement material is paramount as it provides the necessary


input details for the analysis and design of a pavement. Each material has a few
mechanical properties that are more important than others. Flexural characteristics are
important for a cemented material while unbound granular materials are characterised
based on the Mr. However, this does not eliminate the need to characterise the other
properties and, in fact, understanding the different mechanical properties is extremely
useful.

27
Chapter 2 Literature Review

Characterisation reveals the variables that influence a selected material’s properties and
may require laboratory studies to investigate their levels of dependency and the
necessity of including them when defining a material’s properties.

Although, for a lightly stabilised material, the main input parameter is the Mr, studies
focussing solely on it may not provide a better understanding of the material. Moreover,
relationships between different mechanical properties are very useful for calculating a
particular property in the absence of another. As previously mentioned, in the current
MEPDG, three levels of inputs are recommended at the discretion of the designer. A
comprehensive design requires a level-1 material property which is measured in a
laboratory whereas another could be performed with level-2 or -3 inputs. Level-2 inputs
generally relate complex mechanical properties in terms of a simple test, with level-3
commonly using presumptive values. Therefore, the characterisation of a lightly
stabilised material based on different mechanical properties is very important and
determining the empirical relationship among them is useful for differentiation.

The four tests most commonly used to characterise a lightly stabilised material are the
UCS, indirect diametrical tensile (IDT), flexural beam and RLT. The UCS is relatively
easy and used more widely than the other three tests which require sophisticated test
setups.

2.9.1 Characterisation using Unconfined Compressive Strength Test

The UCS test is an index test commonly used for the basic characterisation of any kind
of pavement material and is essential because the UCS is often related to other
mechanical properties. Equation (2.8) and Table 2.7 show the relationships based on the
UCS for obtaining the flexural and elastic modulus suggested by (Austroads 2010b) and
AASHTO (2008) respectively.

EFlex = k UCS Eq. (2.8)

where:
EFlex = flexural modulus (MPa)
k = values dependent on construction specifications and laboratory testing practices
UCS = unconfined compressive strengths of specimens at 28 days

28
Chapter 2 Literature Review

Table 2.7: UCS used for modulus calculation for level-2 inputs in AASHTO (2008)

Material Relationship for modulus


Lean concrete and cement-
E = 57000(qu)0.5
treated aggregate
Open-graded cement-
Use input level 3
stabilised aggregate
Lime-cement-fly ash E = 500+qu

Soil cement E = 1200(qu)

Lime-stabilised soils Mr = 0.124(qu)+9.98

* E, Mr, qu in psi

Although measuring the UCS is more common in a UCS test, the development of
internal measurement techniques for measuring axial deformations has helped studies of
modulus and stress-strain behaviours. Piratheepan et al. (2009) used the UCS with an
internal strain measurement to characterise a lightly stabilised material using different
modulus values. Paul and Gnanendran (2013) used a similar experimental setup to
investigate the non-linearity of stress-strain curves. Lim and Zollinger (2003) developed
the relationship between the compressive strength and elastic modulus of cement-treated
aggregate materials (the cement content was kept between 4% and 8%) in an extensive
study of 189 samples.

In a stabilised material, because of the consistency, reliability and ease of the UCS test,
it is preferred over the other tests for assessing the impacts of different variables. Ateş
(2013) studied the effects of polymer-cement stabilisation on the UCS of liquefiable
soils by changing the polymer and cement percentages of the mixes. White and
Gnanendran (2005) investigated the effects of the compaction method and density on
the UCS and modulus of a reclaimed pavement material stabilised with a slag-lime
binder.

29
Chapter 2 Literature Review

2.9.2 Characterisation using Tensile Properties

The tensile properties of a lightly stabilised material are characterised based mainly on
IDT and flexural beam tests. Although the latter is preferred, a lightly stabilised
granular material with a limited tensile capacity can be conveniently characterised using
the IDT. In both these tests, characterisation can be performed under monotonic and
dynamic loadings. In flexural testing, a monotonic test can be used to measure the
flexural strength and static stiffness modulus (SSM) and a cyclic one to obtain the
dynamic stiffness modulus (DSM) and determine the fatigue life. Similarly, in the IDT,
the indirect diametrical strength and SSM are measured under monotonic loadings while
a cyclic test is performed to determine the DSM and fatigue life.

Of the related works, Paul and Gnanendran (2012b) carried out monotonic beam tests
on a granular material stabilised with 1% to 3% of CF binder. An improved flexural
beam testing arrangement measured the flexural strength and modulus with great
accuracy. The author presented a relationship between the flexural modulus and strength
(equation (2.9) and also reported that the loading rate had a significant effect on flexural
properties.

Eflex = 4840*MoR Eq.


(2.9)

where:
Eflex = flexural modulus (MPa); and
MoR = flexural strength (MPa).

Using IDT, Gnanendran and Piratheepan (2008) performed static and dynamic tests on
specimens stabilised with 3 to 5% of slag-lime and cured for 28 days. In monotonic
tests, the authors measured the indirect tensile strength and SSM by varying the binder
content and moulding moisture content. Characterisation was also performed using the
DSM obtained under cyclic loading and presented equations for defining the
relationships between these mechanical properties.

30
Chapter 2 Literature Review

Similarly, Paul and Gnanendran (2015) characterised granular materials stabilised with
1% to 3% of CF binder using the flexural strength, SSM and DSM measured from
cyclic and monotonic flexural testing. The effects of the applied stress and binder
content on the Modulus of Rupture (MoR), SSM and DSM were evaluated and
relationships among these variables established. According to the authors, despite the
limited tensile capacities of the materials, the testing setup produced reliable and
consistent outcomes.

To obtain the flexural modulus of a cemented material, a flexural beam test of beam
specimens after curing for 28 days is preferred (Austroads 2010b). However, in the
absence of flexural modulus values, correlation of it with other mechanical properties
can be used.

Also, tensile fatigue tests of lightly stabilised materials are performed to determine the
number of cycles required for the material to ‘fail’ based on predefined criteria.
Generally, a reduction in the modulus to half its initial value is considered a failure
criterion. Gnanendran and Piratheepan (2010) performed an IDT test to determine the
fatigue life of a granular material lightly stabilised with 3 to 5% of slag-lime. Two
methods, namely, the energy ratio method and 50% modulus reduction, were defined as
the failure criteria, with both showing similar results.

2.9.3 Characterisation using Resilient Modulus

It is well known that granular pavement layers show non-linear and time-dependent
elastoplastic responses under traffic loading. As shown in Figure 2.5, for most loading
pressures, the plastic response decreases to very lower values while the resilient one
stabilises after a certain number of load cycles. Similarly, it is shown in Chapters 5 and
6, that a lightly stabilised material exhibits the same trend if the applied stress is
relatively lower than the strength of the material, with its ‘stable’ resilient response
based on its cyclic deviatoric stress defined as its Mr. Under monotonic loading, Paul
and Gnanendran (2013) showed that a lightly stabilised material exhibited non-linear
responses. However, defining elastic responses based on the secant or tangent modulus

31
Chapter 2 Literature Review

with the Poisson’s ratio is no longer used because of the wider acceptability of the Mr
obtained from a RLT.

Figure 2.5: Strains under repeated loads (Huang 2004)

For a lightly stabilised granular material, the Mr is the main input parameter for
pavement analysis (Austroads 2010b), with its value obtained using RLT in which a
cylindrical specimen of a lightly stabilised material is subjected to different
combinations of confining and vertical pressures, with the confining pressure generally
kept constant while applying cyclic deviatoric stresses. The load and deformation
readings are used to determine the Mr for each stress sequence.

As the Mr remains the single most material input parameter for an unbound material,
many studies are available based on it and the variables that could affect Mr whereas, as
there are only a few studies related to lightly stabilised materials, little information
about them is available.

2.10 Durability of (Lightly) Stabilised Materials

Durability, ensures the long-term performance of a pavement, is one of the most


important aspects of a granular base material. Environmental factors, such as f-t, w-d
and changes in moisture levels, cause a pavement’s mechanical properties to degrade
and deteriorate its performance. Assessing the durability of a stabilised material is
essential and is performed in the mix design stage. A typical design of a soil-cement
mixture uses strength and durability as the key engineering properties to evaluate the

32
Chapter 2 Literature Review

quality of the mixture. The UCS test helps in the design of a mix that meets the
minimum strength requirement. However, there is no uniform approach for assessing
the durability of stabilised materials as the standard testing procedure is not well
accepted.

In the next section, although existing durability tests are discussed in detail, it must be
remembered that most of them focus mainly on cemented materials, with only and
limited studies available on lightly stabilised materials. To gain a better understanding
of their applicability for assessing the durability of lightly stabilised materials, series of
tests were performed which are discussed later in results chapters.

2.10.1 Laboratory Freeze-Thaw Durability Tests

Laboratory f-t durability studies are performed to assess the strength and stiffness
variation of pavement layers due to f-t cycles. The degree to which f-t affects
mechanical properties depends on many factors, such as the moisture content of the soil,
depth of frost penetration, duration of freezing and thawing cycles, temperatures of
these two phases and number of f-t cycles. Although assessing f-t effects has to be
performed as part of a mix design, there is currently no widely accepted procedure
available. In the next sections, the f-t laboratory tests that are being used to assess f-t
effects are discussed in detail.

2.10.1.1 Standard Freeze-Thaw Test of Cement-Stabilised Soils

ASTM D560 (2003) is the recommended standard testing procedure for soils stabilised
with cement to determine the resistance of a stabilised material subjected to f-t cycles.
According to this procedure, cylindrical specimens are compacted and normally cured
for 7 days and then subjected to f-t cycles. The f-t procedure involves, subject the
specimen at a constant temperature not warmer than −10°F (−23°C) for 24 hours and
thawing for 23 hours at 70°F (21°C) with 100% relative humidity. At the completion of
the first f-t cycle, each specimen has to be given two firm strokes of approximately
13.3N with a wire scratch brush that cover the whole area. This procedure has to be

33
Chapter 2 Literature Review

continued for 12 f-t cycles, with the weight loss in the soil-cement mix measured at the
completion of the 12th cycle and compared against the maximum limits which depend
on the soil type and cement content. Table 2.8 shows an excerpt from ASTM D560
(2003) that defines the allowable limits.

However, ASTM D560 (2003) has become unattractive for the following reasons:
 the difficulty of maintaining consistency during brushing;
 the time required to complete a test being more than a month;
 as stabilisation is currently performed using many supplementary cementitious
additives with Portland cement, its limits on the percentage of cement loss are
of little use; and
 its limits do not work for lightly stabilised materials because their specified
percentages of cement contents are higher in values which are relevant to
cemented materials

For these reasons, many studies have avoided the brushing portion and replaced it with
UCS testing after 12 f-t cycles and compared the results with those from specimens not
subjected to f-t cycles to evaluate degradation, as presented in the next section.

Table 2.8: Maximum allowable weight loss %

AASHTO soil class Cement content (%) Maximum weight loss (%)
3 14
A-4 (5) 5 6
7 3
2 29
A-1-b (0)
3 3
5 19
A-6 (10) 7 10
10 2
8 34
A-4 (8) 10 15
12 8
A-4 (8) 7 9

34
Chapter 2 Literature Review

2.10.1.2 Freeze-Thaw Durability Studies using Unconfined Compressive Strength


Test

As previously mentioned, most f-t durability studies using the UCS follow ASTM-D560
(2003) but, instead of weight loss, UCS after 12 f-t cycles is measured. In fact, Shihata
and Baghdadi (2001) tested three different soils stabilised with cement and found that
the standard UCS and UCS after f-t cycles (i.e., residual UCS), correlated well with
standard ASTM-D560 (2003)’s weight loss and suggested that the standard UCS is a
better indicator of durability due to its shorter duration. A UCS test is preferred over
other mechanical tests because of the ease of performing it.

Similarly, Gutherie et al. (2008) reported that the UCS after 12 f-t cycles showed better
correlations with the VST of two aggregates stabilised with three stabilisers at different
levels of binder content. Miller and Zaman (2000) conducted a UCS test of samples
stabilised with cement kiln dust (CKD) and cured for 7 days to assess CKD’s capability
as a potential stabiliser compared with that of quicklime. Their UCS were tested after 0,
1, 3, 7 and 12 f-t cycles which demonstrated that lime-treated shale gradually gained
strength during f-t cycles whereas the CKD-treated specimens declined up to 7 f-t
cycles and increased after 12 f-t cycles. Although the authors presented the pattern of
strength gain, no attempt was made to compare the results from both stabilisers.
Moreover, Parker (2008), Roper (2007), and Parsons and Milburn (2002) also
performed UCS testing after different numbers of f-t cycles and compared their
correlations with the results from other durability procedures.

2.10.1.3 Freeze-Thaw Durability Studies Using Resilient Modulus Testing

Since the UCS can only be used as an index test because field conditions rarely
represent the unconfined state, performing RLT resilient modulus testing to assess
durability has been considered a better choice (Thompson and L 1990). However, there
are limited studies available that assess the f-t effect on the Mr of a stabilised material
because of the complexity of performing RLT testing. A typical RLT test involves
testing specimens subjected to different numbers of f-t cycles at multiple stress
sequences, commonly after certain intervals of f-t cycles (for example, 4, 8 and 12

35
Chapter 2 Literature Review

cycles), and compare the results against those for specimens not subjected to f-t cycles
to observe the pattern of deterioration.

Zaman. et al. (1999) studied the Mr of a CKD stabilised low quality aggregate base
subjected to f-t cycles using specimens 152.4 mm in diameter and 304.8 mm high
compacted using a vibratory compactor and cured for 7 days. These specimens were
then subjected to f-t cycles with a freezing phase lasting 24 hours at -150C and a
thawing phase of 24 hours at a 21.60C temperature and 95% relative humidity. They
were tested for the Mr after 0, 4, 8 and12 f-t cycles which showed decline in the Mr with
increasing numbers of f-t cycles. A similar observation was reported by Khoury and
Zaman (2007) for various stabilising agents, such as CKD, Class C fly ash and fluidised
bed ash, although the magnitudes of their losses varied from one aggregate to the other.
In cylindrical specimens cured for 28 days and tested after 0, 8, 16 and 30 f-t cycles, the
Mr decreased with increasing numbers of f-t cycles.

According to N and R (2010), the Mr shows different patterns depending on the number
of curing days. Cylindrical specimens were stabilised with 10% Class C fly ash, cured
for 3 and 28 days, and then subjected to f-t cycles. Specimens cured for 3 days showed
an increasing Mr with increasing numbers of f-t cycles up to 30 f-t cycles whereas the
28-day cured specimens showed increased values for the first 12 f-t cycles before
decreasing up to 30 f-t cycles. The UCS values of the 28- and 3-day stabilised
specimens also exhibited the same trends as those of the Mr. The tests were performed
using two f-t procedures and, in f-t procedure 1, a membrane was placed around each
specimen during the f-t cycles. Although both specimens showed similar trend, in f-t
procedure without the membrane they deteriorated higher than the one with membrane.

The f-t test presented by Simonsen et al. (2002) was very different from those in earlier
studies as the Mr test was performed in a custom-made triaxial assembly in which the
specimens were subjected to f-t cycles while their moisture contents were kept constant.
The tests were also performed at different temperatures to assess the influence of
temperature on the Mr.

36
Chapter 2 Literature Review

2.10.1.4 Standard Freeze-Thaw Test of Soils Stabilised with Pozzolans (Vacuum


Saturation Test)

Dempsey and Thompson (1973) first suggested the use of the VST as an alternative test
method for evaluating the durability of stabilised materials against f-t cycles. As,
according to the authors, the results for specimens stabilised with different binders, such
as cement, lime and lime-fly ash, correlated well with those from a f-t durability test,
they labelled the VST ‘rapid and economical’.

Currently, ASTM-D593 (2011)’s ‘Standard specification for fly ash and other pozzolans
for use with lime for soil stabilisation’, uses the VST to assess the quality of fly ash and
other pozzolans that can be used with lime to form effective pozzolanic reactions. As
shown in Table 2.9, this standard specifies testing procedures and limits for strength
development as well as requirements for durability against f-t cycles. By meeting the
minimum vacuum saturation strength, a lime-pozzolan soil is expected to resist f-t
cycles which show that the VST is an indirect f-t durability test of lime-pozzolan
stabilised soils. However, it is currently performed on specimens stabilised with other
types of binders and, as the limit values in Table 2.9 are not available for different types
of binders, it is often performed with other durability tests for comparison purposes.
These types of studies are discussed in the next section.

Table 2.9: Minimum compressive strength requirements for lime-pozzolan


stabilised soils

At 7 days (cured at 54±2 °C) 4.1 MPa


After additional 21 days (cured at 23±2 °C) 4.1 MPa

Minimum compressive strength


(cured at 38±2 °C with 4 hours of soaking in water) 2.8 MP
Minimum vacuum saturation strength 2.8 MPa

Roper (2007) studied the durability of two aggregates, reclaimed aggregates and
crushed limestone, stabilised with three stabilisers, Class C fly ash, lime-fly ash and
Portland cement, using the f-t test, VST and tube suction test (TST). The results

37
Chapter 2 Literature Review

indicated that the VST has the potential to replace the f-t durability test for assessing the
durability of a material.

A similar study of two different subgrades, silty sand and a lean clay, stabilised with
four types of stabilisers, Class C fly ash, lime-fly ash, lime and Type I/II Portland
cement, were conducted by Parker (2008). Three types of test methods, the VST, TST
and f-t durability were performed on cylindrical specimens. The author also proposed
the VST as an alternative test method for assessing f-t durability based on a regression
analysis.

Gutherie et al. (2008) selected two aggregates stabilised with Class C fly ash, cement
and lime-fly ash at three concentration levels and performed the TST, VST and f-t
durability test based on ASTM-D560 (2003) (using the measured UCS instead of the
weight loss after 12 f-t cycles). Correlations among these three tests showed that only
the UCS measured after the f-t durability test and VST produced good regression
constants of approximately 0.7.

These studies suggest that the VST is a good indicator of the f-t durability of a stabilised
material, is not limited to lime-pozzolanic materials and its main benefits are that it can
be completed within 7 days and doesn’t require continuous monitoring. However, its
failure to simulate field conditions is one of its drawbacks.

2.10.1.5 Freeze-Thaw Durability Studies of Flexural Properties

A cemented material is characterised based mainly on its flexural properties as the


tensile strain in its bottom layers is the defining criteria of its fatigue life. Thus,
evaluating flexural properties against f-t is very important for characterisation.
However, as discussed in the last four sections, most f-t durability studies on cemented
materials have been carried out using a UCS or RLT test. The reason that very few
studies have focussed on flexural properties could be because of the unavailability of
suitable testing arrangements. It must be noted that obtaining the flexural properties of a
cemented material from its other material properties is very common, as previously
mentioned (Austroads 2010b). Until recently, characterising the elastic modulus of a

38
Chapter 2 Literature Review

cemented material using the Mr was common while the introduction of M-E designs
demanded the use of flexural properties in pavement analysis.

Khoury and Zaman (2006) examined the f-t durability of a limestone aggregate
stabilised with fly ash using cyclic loading. Beam specimens were stabilised with 10%
Class C fly ash and cured for 1 hour, 3 days and 28 days before being subjected to f-t
cycles. They were tested after different numbers of f-t cycles and it was found that, with
increasing numbers of f-t cycles, degradation increased. The study also investigated two
different f-t testing procedures in which supplying moisture during f-t cycles generated
an adverse effect compared with not supplying excess moisture.

Although a lightly stabilised granular material has a limited tensile capacity, Paul and
Gnanendran (2012b) and Piratheepan J. (2010) showed that the flexural beam test and
indirect diametrical test (IDT) respectively can be performed to characterise it.
Therefore, in this study, f-t durability testing was performed using flexural beam tests
and is discussed further in Chapter 5.

2.10.1.6 Freeze-Thaw Studies using Tube Suction Test

The TST was developed by the Finnish National Road Administration and Texas
Transportation Institute to identify poorly performing unbound aggregates. It measures
the dielectric value (DV) which is a measure of the unbound or free water within a soil
sample and doesn’t simply represent the total volumetric water content (Barbu and
Scullion 2005). The DV (relative permittivity) is the ratio of the permittivity of any
substance to the permittivity of a vacuum or free space (equation (2.10)).

A moist soil matrix comprises solid soil particles, water and air, the DVs of which are 5,
81 and 1 respectively. As the water controls the DVs of the soil, changes in the DVs can
be directly attributed to changes in the moisture content. The DVs are measured using
an Adek Percometer™ and a surface probe which measures DVs by determining the
differences in capacitance. As shown in equation (2.11), the DV is related to
capacitance of the probe and the change in capacitance measured by the probe.

39
Chapter 2 Literature Review

 r =o D Eq. (2.10)

where:

 r = relative permittivity

 o = permittivity of vacuum

D = relative dielectric permittivity (DV)

C  C a ( r1) Eq. (2.11)

where:
 C = change in capacitance measured by probe

C a = active capacitance of probe

 r = relative permittivity of material (DV)

Initially, the TST was used to identify well and poorly performing unbound aggregate
materials based on their moisture susceptibility but was later extended to study the
durability against f-t cycles of bound granular materials (Saarenketo 2000; Scullion and
Saarenketo 1997). In their earlier studies, Scullion and Saarenketo (1997) reported that
the DV measured from the TST could potentially be used to ‘flag’ poorly performing
materials and emphasised that the DV is a representation of the unbound moisture
content not the gravimetric moisture content.

As shown in Table 2.10, they reported the DVs, strength and deformation properties,
and moisture/frost susceptibility of aggregates they tested in both the field and
laboratory. Based on their observations, DVs greater than 16 should be avoided while
those of less than 10 are highly desirable.

40
Chapter 2 Literature Review

Table 2.10: Proposed relationship between DVs and field performance (Scullion
and Saarenketo 1997)

Strength and deformation Frost and moisture


DV
properties susceptibility

Low tensile strength, might be Not susceptible to frost and


<5
sensitive to PD moisture

Not susceptible to frost and


5-7 Optimum strength properties
moisture

Improper drainage may lead to


7-10 High shear strength
frost and moisture susceptibility

10-16 Less shear strength Sensitive to frost and moisture

Extremely sensitive to frost and


>16 Less shear strength and high PD
moisture

Further studies by Saarenketo (2000) and Guthrie and Scullion (2003) confirmed that
the above criteria can be simplified and used to classify a material’s performance as:

DV > 16 poor;

10 > DV > 16 marginal; and

DV < 10 good.

According to Syed et al. (2000), the mechanical classification and characterisation of


the moisture susceptibility of base and subgrade materials based on the particle size
distribution and plasticity index might not be adequate for identifying poorly
performing materials. Also, they showed that a material classified as acceptable based
on the traditional classification performs poorly in the field and eventually measures
high DVs. Therefore, they thought that it is important to include the TST with existing
methods to identify moisture-susceptible materials.

41
Chapter 2 Literature Review

Moreover, these poorly performing materials have been stabilised with cementitious
binders to obtain materials less susceptible to moisture. Syed et al. (2000) and Scullion
and Saarenketo (1997) reported that the material classified as poor by the TST showed
better performance after stabilisation. Syed et al. (2000) used a concentrated liquid
stabiliser to stabilise four poorly performing materials and obtained better
performances. Similarly, an in-situ study by Scullion and Saarenketo (1997) showed
that the material classification of the TST for identifying moisture-susceptible material
works well, with stabilisation one solution for limiting moisture sensitivity.

Zhang and Tao (2008) conducted the TST on low-plasticity soils moulded at six
different cement contents (2.5, 4.5, 6.5, 8.5, 10.5 and 12.5%) and four different
moulding moisture contents (15.4, 18.5, 21.5 and 24.5%), and found that an increase in
the binder (cement) content reduces the DV and delays the rate of capillary rise.
According to the authors, this delay could be a welcoming factor if the material is
exposed to moisture for a short period of time. Moreover, the specimens compacted on
the dry side of the OMC with a low cement content absorbed water at a higher rate than
those compacted on the wet side of the OMC.

Parker (2008) evaluated the moisture susceptibility of 7-day cured and stabilised silty
sand and lean clay specimens. Five different additives, namely, Class C fly ash, lime-fly
ash, lime and type I/II Portland cement at three different concentration levels were used
for stabilisation. From the outcomes, the performance of each stabiliser was evaluated.
A comparison of the DVs of the stabilised and unbound materials showed that
stabilisation reduces moisture susceptibility but its effectiveness depends on the type of
stabiliser.

Si and Herrera (2007) used CKD as a stabiliser for a limestone base by comparing its
performance with those of lime, cement and fly ash. The DVs of specimens cured for
short and long terms were tested and, according to the authors, the TST is not a good
indicator to use to evaluate the moisture susceptibility of a stabilised material but did
not clearly explain this conclusion. Barbu et al. (2004) studied the moisture
susceptibility of 28-day cured silty sand stabilised with 3.5% of cement and also
reported that stabilisation improves the moisture susceptibility of this material.

42
Chapter 2 Literature Review

The studies presented so far show that the TST has been used to classify a material
according to its moisture susceptibility which, they emphasise, does not simply
represent the gravimetric moisture content of a soil. Another important observation is
that field studies are important for classifying different responses for comparison with
outcomes from laboratory studies.

The success of its moisture prediction has prompted using the TST to test other
mechanical properties, such as the bearing capacity, frost heave, UCS after f-t cycles
and several other parameters, in an attempt to classify an aggregate based on its
performance. The fact that the final DVs are related to these mechanical properties is
useful for determining one property in the absence of others or developing criteria based
on a simple testing methodology; for example, Zhang and Tao (2008) compared the
performances of different soils compacted at different cement contents using the TST
and 7-day UCS test and reported that their results were equivalent in terms of
examining the effectiveness of stabilisation.

Little (2000) evaluated the moisture susceptibility of low, moderate and high-plasticity
soils using the TST. The DVs of specimens of untreated and lime-treated soils with
unsealed and controlled curing were measured until they reached equilibrium. It was
found that unsealed curing resulted in higher DVs and that sealed curing and the
outcomes related well with UCS and Mr of the same specimens.

Scullion and Saarenketo (1997) first related the DVs to the properties of frost-
susceptible materials, with the strength of a material and its capability to resist repeated
f-t cycling considered to be directly influenced by the unbound water in it. As the TST
is a measure of the unbound water, DVs can be expected to provide some indication of
the durability of a material.

Currently, it’s common to perform the TST with other durability studies to determine
the relationships between tests and examine their suitability for evaluating f-t durability.

Zhang and Tao (2008) performed the TST and 7-day UCS test on 5 different soils with
different cement contents and found that their DVs and UCS showed good correlations.
However, Parker (2008) performed an extensive study of sand and clay soils stabilised
with five different binders at different concentrations. The UCS after f-t cycles, VST

43
Chapter 2 Literature Review

and TST were conducted on the specimens to evaluate the permanence of the selected
binders. A strong correlation was identified between the UCS obtained after f-t cycles
and VST while the DVs showed less convincing relationships with both UCS values.
Also, Gutherie et al. (2008), (Roper 2007; Solanki et al. (2011)) reported that the TST
did not produce convincing relationships with other durability studies.

As the TST has not yet been standardised, there is no common practice and different
researchers have assessed the influences of different variables, including the; size of the
sample, compaction method, single or multiple layers, quality of the porous disk, water
quality, air temperature, salinity of mixing and soaking water, dielectric probe seating
force, number of days of drying before the start of capillary soaking, height of the water
bath and compaction energy.

Barbu et al. (2004) assessed the effect of the diameter on the final DV using three
different silty sand specimens stabilised with 3.5% of cement and cured for 28 days. A
height of 180 mm was kept constant while two different diameters, 152mm and 101.6
mm, were used for the specimens considered in the study. As the final DVs measured
after 500 hours showed little difference between the two diameters, it was concluded
that size has little effect on the final DV of a specimen.

Barbu et al. (2004) also studied the effect of compaction energy on a selected silt soil
and found that no effect on final DVs were observed. Specimens compacted at two
compaction energies achieved the same MDD for the selected moisture content which
could have resulted in same porous structures and same DVs. Barbu and Scullion
(2005) found that porous stones used at the bottom of a specimen must be clean to allow
unobstructed capillary absorption and the specimens’ surfaces need to be smooth to
provide consistent readings. According to the authors, distilled water, tab water had
little effect on the final DVs.

Guthrie and Shambaugh (2009) used two different aggregates (caliche and limestone) to
investigate seven factors, the air temperature, salinity of the mixing and soaking water,
dielectric probe seating force, number of days of drying before the start of capillary
soaking, fines content in the aggregate, height of the water bath and compaction energy.
Statistical hypothesis tests showed that the air temperature, drying time, fines content,

44
Chapter 2 Literature Review

water bath height and compaction energy were significant for the caliche aggregates
while only air temperature was important for the limestone ones. The probe seating
force and water salinity didn’t significantly influence the final DVs.

2.10.2 Wet-Dry Laboratory Durability Tests

W-d action is one of the environmental factors that cause problems for a pavement
constructed of chemically stabilised soils. As the moisture level within a pavement
fluctuates as a result of seasonal changes, the pavement’s materials are subjected to
wetting and drying phases, the continuous reversal of which adversely affects the
pavement. Therefore, w-d durability studies are performed in the mix design stage along
with strength characterisations to assess the materials’ capabilities to withstand such
adverse effects.

A laboratory w-d testing procedure simulates the moisture variations that occur within
pavement layers in a rapid time scale to evaluate the capabilities of the materials to
resist adverse w-d effects. The only relevant standard available for w-d durability
testing is ASTM-D559 (2007) which sets out the procedure for soil-cement mixes.
According to the testing procedure, cylindrical specimens subjected to w-d cycles are
given two firm strokes after each w-d cycle using a wire brush covering the whole
cylindrical surface and, after 12 w-d cycles, the percentage loss of soil-cement is
measured and compared against acceptable limits. One w-d cycle comprises soaking the
specimens in potable water at room temperature for a period of 5 hours and then placing
them in an oven at 160°F (71°C) for 42 hours. It can be recalled that this specification is
very similar to the f-t testing procedure in ASTM-D560 (2003) but f-t cycles are
replaced by w-d cycles. All the problems mentioned in section 2.8.1.1 regarding
ASTM-D560 (2003) are applicable to ASTM-D559 (2007) which makes this standard
less attractive.

An investigation of the durability of silt clay against w-d cycles by Zhang and Tao
(2008) is one of the few studies available that totally embraced measuring the soil-
cement loss. They studied the resistance to w-d cycles of silt clay specimens moulded at
four different moisture contents and stabilised with six different cement contents by

45
Chapter 2 Literature Review

measuring their percentage weight losses and concluded that percentage weight loss w-d
durability test and 7-day UCS tests showed good correlations and were equivalent for
predicting durability.

To eliminate the brushing portion in the ASTM-D559 (2007) testing procedure,


measuring the UCS after 12 w-d cycles instead of the soil-cement loss has been reported
in many studies (Ganne 2009; Miller and Zaman 2000; Parsons and Milburn 2002).
Miller and Zaman (2000) evaluated the durability of subgrade soils stabilised with
CKD. The compacted specimens were cured for 7 days and then subjected to w-d cycles
in compliance with ASTM-D559 (2007). However, after 12 w-d cycles, the UCS was
measured instead of the percentage of weight loss and was found to increase with the
number of w-d cycles.

Similarly, by following ASTM-D559 (2007), Ganne (2009) studied the durability


against w-d cycles of reclaimed asphalt mixes stabilised with cement and collected from
six different sites in Texas by measuring the UCS and volume change after 3, 7 and 14
w-d cycles to identify the best-performing mix. It was found that the volume change
was less than 1% and the UCS retained after w-d cycles showed good correlation with
the volumetric change in a specimen.

Parsons and Milburn (2002) compared the performances of 7 different clays and silts
soils stabilised with four different binders, lime, cement, fly ash and an enzymatic
stabiliser. The additive contents were chosen based on the ASTM minimum criteria for
different stabilisers. The durability of the stabilised materials against w-d cycles were
assessed based on ASTM-D559 (2007) but, after 12 w-d cycles, the UCS was measured
instead of the percentage loss of soil-cement.

The above studies show that the weight loss is no longer preferred as many used the
UCS. However, the un-confining nature of the UCS is criticised for not being
representative of field stress conditions. Khoury (2007) reported that Kalankamary and
Donald (1963) considered that the ASTM D 559 test method is too severe and not
representative of field conditions.

Using triaxial testing to evaluate the Mr with different stress sequences is considered
more relevant than the UCS as it simulates in-situ conditions better. As this Mr test is

46
Chapter 2 Literature Review

time consuming, few studies have focussed on it for evaluating the durability of
stabilised specimens subjected to w-d cycles. Of those available, Zaman. et al. (1999)
carried out tests on a limestone aggregate stabilised with CKD after 0, 4 and 12 w-d
cycles. According to the authors, as these numbers of w-d cycles deteriorated the
stabilised material, 8 to 12 cycles might be adequate for assessing their effects.

Khoury and Zaman (2002) investigated the effects of w-d cycles on aggregates
stabilised with 10% of coal fly ash. Specimens cured for 3 and 28 days were used for Mr
testing after 0, 4, 12 and 30 w-d cycles. The Mr of specimens cured for 3 days increased
up to 30 w-d cycles without showing any sign of deterioration while those cured for 28
days showed reductions after 12 w-d cycles which demonstrated the influence of the
number of curing days on w-d effects.

Khoury (2007) performed Mr testing of four aggregates stabilised with 15% of CKD,
10% of Class C fly ash and 10% of fluidised bed ash, cured for 28 days and subjected to
0, 8, 16 and 30 w-d cycles. It was found that the Mr values of three of the aggregates
decreased as the number of w-d cycles increased up to 30 cycles while the other showed
increased Mr values for 8 w-d cycles and then decreased ones up to 30 w-d cycles.
These studies demonstrated the capability of triaxial tests to evaluate the w-d durability
of a stabilised material.

Also, selection of the environmental parameters for a w-d durability test is important as
this may influence the final outcome of an experiment. A review of previous studies
reveals that no uniform approach has been adopted in selecting the environmental
parameters for the wet and dry phases. Parsons and Milburn (2002), Zhang and Tao
(2008) and Ganne (2009) selected the parameters suggested in ASTM-D559 (2007),
whereby compacted specimens cured for 7 days have to be submerged in water at room
temperature for 5 hours and then placed in an oven at a temperature of 71.8oC for 42
hours. However, Miller and Zaman (2000) used a 5-hour wetting phase followed by
oven-drying for 24 hours at 71oC. N and Zaman (2007) and Khoury (2007) considered
one w-d cycle consisting of placing a specimen in an oven at a temperature of 71.80C
for 24 hours and then submerging it in water for 24 hours at room temperature.

47
Chapter 2 Literature Review

This non-uniformity in selecting environmental parameters restricts comparisons of the


results from different studies and a selected parameter may be too severe for a particular
soil as it has been reported that specimens deteriorated even before reaching 12 w-d
cycles. Parsons and Milburn (2002) performed w-d tests on seven different clays and
silts stabilised with lime and cement, Class C fly ash and an enzymatic stabiliser, and
reported that ASTM-D559 (2007) was too severe for the selected materials at the
selected binder content since most specimens failed to survive 12 w-d cycles. The
remaining question is whether the particular material was incapable of resisting w-d
cycles or the selected parameters were too severe for it. Unless suitable parameters are
established, each w-d test result would be limited to that particular study. Moreover, as
almost all the studies presented in this section were performed on cemented materials
(materials with high binder percentages), it will be interesting to see the w-d resistance
of lightly stabilised granular materials.

2.11 Moisture content fluctuation within the Pavement

In arid and semi-arid regions, variations in the moisture content in a granular base are
critical. Water is needed as a lubricant to achieve high levels of compaction during
construction, with the moisture content at compaction affecting the strength and
stiffness properties of the soil. Excess moisture in a pavement causes excess pore-water
pressure which reduces the effective stress, and, hence, the pavement’s strength,
stiffness and resistance to deformation, thereby leading to its failure. The detrimental
effects of water can be a reduction in the strength of the subgrade and unbound granular
materials, pumping of a concrete pavement, differential heaving over swelling soils and
stripping of an asphalt mixture (Huang 2004) which subsequently lead to deterioration
of the pavement. Therefore, it is necessary to understand the sensitivity of the materials’
properties to different moisture contents.

Generally, construction specifications recommend that a granular base and subgrade be


constructed at the optimum moisture content (OMC) and attain the maximum dry
density (MDD) (Qian 2010). Although, during construction, these layers are compacted
close to the OMC, the moisture content beneath the pavement changes and reaches an
equilibrium moisture content (EMC) (Perera Y.Y et al. 2004). The EMC depends on the

48
Chapter 2 Literature Review

types of materials, environmental conditions, water table, permeability of capillary


action and drainage conditions, and is a constant often in the unsaturated zone.

However, it can still fluctuate if the groundwater table reaches to within about three feet
of the ground surface and the edge of the pavement is within three to six feet of the
point of consideration (Perera Y.Y et al. 2004). Moreover, improper drainage, porous
asphalt layers, insufficient shoulder width and high water table could bring the base
layer closer to the saturated condition. Therefore, the design of a pavement requires a
good understanding of the materials’ reactions to moisture changes.

This investigation would/will be useful if it enables the EMC (the possible operating
moisture content) to become predictable. Recently, researchers have developed models
for predicting the EMC of granular base materials using different parameters which are
quite successful (NCHRP 2004; Perera Y.Y et al. 2004). The advantage of moisture
prediction models is that, based on predictions of the EMC, changes in a material’s
properties due to its moisture content changing from the OMC to EMC can be
evaluated. This offers the possibility of obtaining design parameters at the expected
EMC rather than OMC for use in a design which would solve the problem of moisture
variations.

49
3 Chapter 3

Research Methodology

3.1 Introduction

The primary purpose of a pavement is to resist wheel loadings and provide a good-
quality riding experience for users. It is designed with a particular life expectancy
within which a major failure could cause many socio-economic issues. Other than the
wheel loadings, the performance of a pavement is influenced by the environmental
conditions, the quality and methods of its construction and its material characteristics,
with environmental factors, such as freeze-thaw (f-t), wet-dry (w-d) and moisture
variations, having the potential to cause problems.

Stabilising granular materials to improve its various physical and mechanical properties
and ensure its quality performance in conjunction with other pavement layers is very
common. In a mix design, stabilised materials are assessed for their durability against f-
t, w-d and moisture using different laboratory tests and material characterisations, with
changes in the mix design continuing until the minimum criteria for durability limits are
satisfied.

Unlike for a cemented material, in a typical mix design of a lightly stabilised granular
material, durability studies are not performed, i.e., durability is considered a major
concern. However, for the better use of a lightly stabilised material, as its mechanical
properties are affected by environmental factors, it is necessary to understand its
behaviours. Therefore, in this experimental investigation, series of tests were performed
focussing on different properties of lightly stabilised granular materials to evaluate its
behaviour. Different testing procedures were implemented to find a suitable laboratory
test for evaluating the effects of environmental factors on lightly stabilised materials.
Chapter 3 Research Methodology

Of the related methods available for f-t durability test, ASTM D560 (2003) recommends
a procedure for granular materials stabilised with cement, ASTM C593 (2011) a
vacuum saturation test (VST) for soils stabilised with lime and lime-fly ash/pozzolans
and the w-d durability is performed using ASTM-D559 (2007). Apart from the VST, the
other two have limited use because of the ambiguity associated with their testing
procedures. As the applicability of these tests to lightly stabilised materials is unknown,
as shown in Table 3.1, various elements tests incorporating different variables were
performed.

Four tests, the f-t durability, VST, 7-day unconfined compressive strength (UCS) and
tube suction test (TST), are discussed in Chapter 4. F-t durability testing was based on
ASTM D560 (2003) but, instead of weight losses, the UCS after 12 f-t cycles were
measured. This was the only test in which the specimens were subjected to f-t cycles.
The testing procedures mentioned in ASTM C593 (2011) were adopted for the VST
while the 7-day UCS was obtained after 7 days of normal curing. The TST, an indirect
test which measures the dielectric value (DV), was also included with the other three
durability tests because of its wider acceptability. The purpose of these tests was to
compare their results and assess their acceptability for evaluating the durability of
lightly stabilised granular materials.

The presence of moisture and its fluctuations in a pavement system is another important
environmental consideration that needs careful attention. Water acts as a lubricant to
achieve high levels of compaction during construction, with the moisture content at
compaction affecting the strength and stiffness properties of a soil. However, as the
excess moisture in a pavement develops excess pore water pressure and reduces the
effective stress, the resultant inadequate strength, stiffness and deformation resistance
can lead to failure. To evaluate the influence of moisture on UCS and flexural
properties, another set of specimens were prepared and tested for their moisture
sensitivity.

Chapter 5 also presents studies of durability but its main focus is to analyse the changes
in the mechanical properties of specimens subjected to different numbers of f-t cycles,
with the different influencing factors investigated using three different element tests. In

51
Chapter 3 Research Methodology

addition to the f-t tests, w-d studies which investigated the flexural and compressive
properties of the lightly stabilised granular materials material are presented.
Comparisons of the mechanical properties tested, testing variables and procedures used
were expected to provide a comprehensive analysis of the durability of a lightly
stabilised granular material.

In chapter 6, the PD of the lightly stabilised material investigated using repeated load
triaxial (RLT) test is presented. The PD responses were analysed using shakedown
theory and the material responses are classified into different shakedown ranges. Based
on the classification of ranges, shakedown limits calculation to obtain plastic and elastic
shakedown limits are presented. Characterisation of deformation behaviour is also
studied using Accelerated pavement model (APM). Four axial pressures were used to
obtain deformation characterises of base and subgrade. The test instrumented with
deformation and pressure-measuring devices which were used for back-calculations of
base and subgrade modulus. A finite element (FE) model and outcomes from the RLT
tests were used to obtain a relationship between the APM and RLT test results.

Table 3.1 shows a summary of the tests presented in each chapter and the objectives
expected to be accomplished. Details of the procedures for each experiment, the
variables involved, as well as outcomes from, and comparisons of, different
experiments are presented in the respective chapters.

52
Chapter 3 Research Methodology

Table 3.1: Breakdown of thesis chapters, experiments and objectives

Thesis chapter Objectives Tests

 Obtaining UCS after f-t cycles,


vacuum saturation and 7-day
normal curing VST
UCS after 12 f-t cycles
 Measuring DVs and testing
UCS after 7 days
different procedures and
Chapter 4 TST
variables
 Comparing different test results Flexural beam test
 UCS and Flexural test to study UCS test
the moisture sensitivity of the
material

 Measuring properties after


different numbers of f-t cycles
 Changing variables to obtain
UCS test
better laboratory test
Chapter 5 RLT test
 Extending tests to w-d cycles Flexural test
 Comparing different mechanical
properties to find better
approach

 Analysis the PD behaviour using


shakedown theory
 Calculation of shake down
limits RLT test
Chapter 6
 Analysis of resilient deformation
 Comparing PD test outcome of
specimens stabilised with
different moisture content

 Performing APM testing with


different instrumentations
 Classifying responses using
shakedown approach APM test
Chapter 7
 Back-calculating elastic RLT test
properties
 Developing relationship
between RLT and APM test
results

53
Chapter 3 Research Methodology

3.2 Materials

For the experimental study, two base and one subgrade material were used, with details
of their physical properties presented in the next section.

3.2.1 Base Materials

A Queensland granular material (QM) and a Canberra granular material (CM) were the
two base materials selected for most of the element testings while the APM tests utilised
only CM. Their basic soil properties are presented in this section.

3.2.1.1 Queensland Granular Material (QM)

QM is widely used for road construction in Queensland and was extracted from
Wagners Wellcamp Quarry (Figure 3.1). According to the unified soil classification
system, the material is classified as sand with fines and has the dual symbol SM-SC.

A material’s gradation is one of its physical properties that influence its strength.
Particle size distribution of different samples suggested that obtaining a consistent
grading through a random sampling technique was not possible and the impact of
grading was found to be significant. The flexural properties of two different particle size
distributions (Figure 3.2) were tested after 14, 28 and 56 days of curing (details of the
test is presented later in this chapter). As shown in Table 3.2 and 3.3, the modulus of
rupture (MoR) and static stiffness modulus (SSM) of three specimens tested for each
grading varied considerably. Therefore, to eliminate the gradation effect, a reconstituted
material (i.e., material has predefined amount of each particle size) was obtained after a
series of size analyses, with the same particle size distribution used for all experiments.

The maximum particle size is important in the making of a small specimen as particles
greater than a particular size may affect the performance of the whole specimen.
Generally, the maximum particle size is defined by the standards based on the
specimen’s diameter; for example, AS-5101.4 (2008) recommends a particle size of less
than 19 mm for UCS specimens with diameters of 105 mm. Therefore, as shown in

54
Chapter 3 Research Methodology

Figure 3.3, boulders larger than 19 mm were sieved and removed from the reconstituted
particle size distribution. Moreover, according to ASTM-D1241-07 (2007), the selected
particle size distribution satisfied the requirements for that of type 1, gradation C road
base materials.

Figure 3.1: Queensland granular material (QM)

55
Chapter 3 Research Methodology

Grading 1
Grading 2
0.8
Percentage Passing(%)

0.6

0.4

0.2

0
0.01 0.1 1 10 100
Particle Size (mm)
Figure 3.2: Particle size distributions of QM materials

Table 3.2: Impacts of gradation on MoR of lightly stabilised granular material

14-day MoR 28-day MoR 56-day MoR


Grading Specimen
(kPa) (kPa) (kPa)
1 176 261 435
2 194 320 366
Grading 1
3 178 295 382
Average 183 292 394
1 292 433 530
2 277 406 499
Grading 2
3 273 401 506
Average 281 413 512
% difference 35 29 23

56
Chapter 3 Research Methodology

Table 3.3: Impacts of gradation on SSM of lightly stabilised granular material

14-day SSM 28-day SSM 56-day SSM


Grading Specimen
(MPa) (MPa) (MPa)
1 1050 1476 2534
2 1112 2293 2402
Grading 1
3 1323 1768 2813
Average 1162 1768 2583
1 2336 2760 3263
2 2192 2725 3578
Grading 2
3 2514 3379 3752
Average 2347 2955 3531
% difference 50 40 27

Almost all construction activities specify that the materials are to be compacted at their
optimum moisture contents (OMC s) and attain maximum dry densities (MDDs) greater
than 95%. A Proctor compaction test was performed to obtain the OMC-MDD
relationship for the unbound QM shown in Figure 3.4. As outlined later in this chapter,
different amounts of binders were used to stabilise the specimens. The inclusion of
powder like particle is likely to bring changes in the grading which could potentially
affect its OMC-MDD relationship. However, Paul (2012) showed that adding a small
percentage of binder has little impact on this relationship and the values obtained for the
unbound material can be used for lightly stabilised granular material. Therefore, OMC
of 9% was selected as the compaction moisture content for preparing specimens with
the attainment of MDDs of 2.3 kN/m3.

57
Chapter 3 Research Methodology

100

CM
QM
80 Grade C_min
Grade C_max
Percentage passing(%)

60

40

20

0
0.01 0.1 1 10 100
Particle size (mm)

Figure 3.3: Reconstituted particle size distributions of CM and QM

58
Chapter 3 Research Methodology

2320

2280

Dry Density (kg/m3) 2240

2200

2160

2120
4 6 8 10 12
Moisture Content (%)

Figure 3.4: OMC-MDD relationship of QM

3.2.1.2 Canberra Granular Material (CM)

The other granular material, CM, which was obtained from Mugga Quarry in Canberra
was supplied by Boral Resources (Country) Pty Ltd. It was a little coarser than QM and
required only 8% to reach a MDD of 2.1 kN/m3 as shown in Figure 3.5. Similar to QM,
its reconstituted particle size distribution was established and is shown in Figure 3.3.
However, for APM testing, due to the large volume required for model construction,
reconstituted particle size distribution could not be used.

A photographic view of this material, which was sandy with gravel and some fines, is
shown in Figure 3.6. It was reddish in colour and possessed sharp edge particles
probably due to the mechanical crushing of rocks undertaken to form crushed
aggregates.

59
Chapter 3 Research Methodology

2140

2120
Dry Density (kg/m3)

2100

2080

2060

2040
4 6 8 10
Moisture Content (%)

Figure 3.5: OMC-MDD relationship of CM

3.2.2 Subgrade Material

A subgrade material obtained from Queensland, Australia was used to construct a model
pavement for APM testing, with a photographic view of it shown in Figure 3.7 and its
particle size distribution in Figure 3.8. Its OMC of 33% and MDD of 1.4 kN/m3
obtained from the standard Proctor compaction test are graphically illustrated in Figure
3.9 and its plasticity index (PI) of 56.4% in Table 3.4

60
Chapter 3 Research Methodology

Figure 3.6: Canberra granular material (CM)

Figure 3.7: Queensland subgrade material

61
Chapter 3 Research Methodology

100

90
Percentage Passing (%)

80

70

60
0.001 0.01 0.1 1 10
Particle Size (mm)

Figure 3.8: Particle size distribution of Queensland subgrade material

62
Chapter 3 Research Methodology

1400

1350
Dry Density (kg/m3)

1300

1250

1200

1150

1100
10 20 30 40 50
Moisture Content (%)

Figure 3.9: OMC-MDD relationship of Queensland subgrade material

Table 3.4: Physical properties of Queensland subgrade material

Property Value
Liquid limit (%) 93.4
Plastic limit (%) 37
PI (%) 56.4
OMC (%) 33
MDD (kN/m3) 1.38
Soil classification as per USCS CL
Soil classification as per AASHTO A-7

63
Chapter 3 Research Methodology

3.2.3 Lightly Stabilised Granular Base

Stabilisation is a process for improving the quality of a material by adding binders or


mixing it with other granular materials. As the main interest of this thesis is limited to
cementitious binders, other stabilisation techniques briefly discussed in the previous
chapter are not presented here.

Depending on the amount used and its capability to improve the quality of a base
material, stabilisation can be classified in different categories. However, the definition
of each category is not ‘black and white’, as can be seen in Table 3.5 which shows that
the definition of different category of stabilisation defined by Austroads (2002) and
Austroads (2010a) based mainly on the UCS. Although categorisation is not well
defined, in other words, the exact UCS separating the different categories are not
universally accepted, it is essential to understand that, depending on the ‘degree of
cementation or bonding’, different properties of the material become significant in terms
of a pavement’s design.

As shown in Table 3.6, in this study, two different combinations of binders with
different percentages were used. However, as most tests were performed using 1.5%
and 0.5% of cement-fly ash (CF) binders, it was considered that the material would
most likely fall under the category of a lightly stabilised granular material as defined by
Austroads (2002). Therefore, the term lightly stabilised granular material is selected to
describe the resultant material.

As previously mentioned, although there is no universally accepted categorisation, it is


essential to understand the differences between fully and partially bonded materials. The
main characteristic of a fully bonded material, which can also be referred to as a
cemented material, is its significant flexural strength. For the stabilisation of cemented
materials, the selection of a particular binder and its amount are determined from a mix
design procedure, with importance given to strength and durability requirements. Then,
laboratory tests are performed to obtain mechanical properties (mainly flexural) which
are used in the design based on any relevant M-E design standards.

64
Chapter 3 Research Methodology

Table 3.5: Categories of stabilisation

Category of Design strength Design flexural


Degree of binding
stabilisation (MPa) modulus (MPa)
Modified material UCS1 < 1.0 1000
Lightly stabilised
1< UCS1 < 4 1500 - 3000
Austroads (2002) material
Heavily stabilised
UCS1 > 4 5000
material
Modified material 0.7 < UCS2 < 1.5 -
Austroads (2010)
Bound material UCS2 > 1.5 -
1 - 28-day test results, with standard compaction and moist curing according to AS-1141.51 (1996)
2 - Values determined from test specimens stabilised with GP cement and prepared using standard
compaction effort, normal curing for 28 days and 4 hours of soaking

Table 3.6: Binders and binder percentages used for different tests

Test Type of test Type of binder and binder %


UCS test Durability 1.5% CF, 3% CF
RLT test Durability 1.5% CF
Flexural beam test Durability 1.5% CF, 3% CF
0.5% CF, 1.5% CF, 3% CF
TST Durability
3.0 % SL, 4.5% SL, 6% SL
VST Durability 1.5% CF
RLT test Moisture 0.5% CF, 1.5% CF
UCS test Moisture 1.5% CF
Flexural beam test Moisture 1.5% CF
APM - 1.5 CF

In light stabilisation, engineers select the required properties and work on the amount
and type of binder necessary to improve them, with the possible objectives of:

 reducing the Plasticity index (PI);

 increasing the shrinkage limit;

65
Chapter 3 Research Methodology

 improving durability;

 improving permeability or moisture susceptibility;

 improving strength and stiffness;

 reducing the amount of clay/silt-sized particles; and

 reducing the number of volume changes.

As the purpose of light stabilisation can be any of the above, the material does not go
through a typical mix design process (to obtain the optimum binder content and satisfy
minimum strength and durability requirements) similar to that for a cemented material.
Instead, a trial and error approach is followed until the expected properties or
improvements are obtained. The chemical and physical properties of both the soil and
binder are essential for selecting a suitable binder or combination of binders for
stabilisation. It must be noted that, in a search to improve a selected property, it is likely
that the material’s other properties are enhanced and such improvements should be
considered wherever applicable during the design phase to achieve a better pavement
design. As it’s true that fully and partially bounded materials have differences, the
question remains unanswered is that the UCS value or any other mechanical property
that clearly separate the type of stabilisation.

3.2.3.1 Cementitious Stabilisation

The type and amount of binder required for stabilisation should be based on the type of
soil, availability and cost of the binder, working time, environmental concerns and
possible shrinkage cracking. For the selection of binders, guides showing the
effectiveness of each binder in different types of soils can be used. Figure 3.10, obtained
from Austroads (2002) is one of those type, shows that the plastic index is related to the
type of binder and, as can be seen, cementitious binders are very effective for most soil
types.

Cementitious stabilisation uses Portland cement either alone or with other


supplementary materials, such as lime, fly ash or slag, that react with water and form the

66
Chapter 3 Research Methodology

cementitious products calcium silicate hydrates, calcium aluminate hydrates, calcium


aluminate sulphate hydrates and calcium alumino-ferrite which bind the solid particles
together and provide increased strength and stiffness. Portland cement possesses all the
chemical compounds necessary for reacting with water and producing cementitious
products. However, pozzolanic stabilisers require an alkaline environment for active
reaction with the silicates and aluminates they possess; for example, fly ash and slag are
often used with cement or lime, which produces the hydroxide (OH-), to form a
reaction. Moreover, as lime requires silicates, which are often provided by the soil,
producing cementitious products is much more effective with reactive silicate-rich soils
(Vorobieff 1997b). Soils with insufficient silicates require pozzolanic materials, such as
fly ash or slag, to supply silicates for use the remaining lime.

Plasticity PI ≤ 10 10<PI<2 PI ≥ 20 PI≤ PI ≤ 10 PI > 10


PI×% Passing
index 0
75µm ≤60

Cement and
cementitious
blends
Lime

Usually suitable

Usually not suitable

Doubtful

Figure 3.10: Guide for selecting binder based on PI

In this study, most tests were performed on specimens stabilized with CF binder, except
for TST in which slag-lime (SL) was used for one series. The ratio of cement : flyash of
75%:25% was kept constant throughout testing while slag : lime was kept at 85%:15%.
As outlined below, there were several factors which contributed to the selection of the
binders and their proportions.

67
Chapter 3 Research Methodology

 It was decided not to use cement alone as it might not have promoted a ‘greener’
concept.
 As lightly stabilised granular materials are used across Australia in low-volume
roads where cost is a major concern, replacing a portion of cement with fly ash
is likely to result in greater savings.
 One of the important considerations when selecting a binder and its composition
is the setting time, with a high setting time enabling better compaction and
providing safety against uncertainties during construction (Smith and Hansen
2003). As using fly ash with cement will increase the setting time, these benefits
can be achieved.
 A high strength often causes extensive shrinkage cracking with resultant
declines in durability which can be avoided by using fly ash (Vorobieff 2000).

 However, the proportion of fly ash cannot be excessive because it may delay
strength development and give a lower strength gain. As most laboratory tests
needed to be performed quicker to make it acceptable, the fly ash content was
limited to 25%.
 As a SL ratio of 85:15 is commercially produced and sold to consumers as a
single product, it was used directly at the same proportion.

In the following section, detailed descriptions of the binders are presented.

3.2.3.1.1 General Blended Cement

Two types of cements are used in Australia, general purpose (GP) and general blended
(GB). A GB cement contains Portland cement and a quantity greater than 7.5% of fly
ash or slag or both, and up to and including 10% of amorphous silica. According to
Vorobieff (2000), GB cements provide 1 to 2 hours additional working than GP cement.
Moreover, their previously mentioned benefits such as cost savings, environmental
benefits and less reflective cracking, have encouraged manufacturers to produce GB
cements with specific proportions consisting of fly ash and blast furnace slag.

68
Chapter 3 Research Methodology

In this study, a 20 kg bag of commercially available GB cement, a Blue Circle® product,


was used. According to the manufacturer, it satisfies the requirements of AS-3972
(2010), as shown in Table 3.7. The hydration of GB cement generates a cementitious
paste which binds the granular material together and gives it strength upon hardening
and also provides an alkaline environment with which the fly ash can react. The nature
of the fly ash used with GB cement is discussed in the next section.

Table 3.7: Performance comparison of GB cement against AS 3972 requirements

Property GB cement AS 3972

Setting time
Initial 2-3 hours 45 minutes minimum
Final 3-4 hours 10 hours maximum
Compressive strength
18-25 MPa -
3-day

7-day 27-39 MPa 20 MPa minimum

28-day 50-60 MPa 35 MPa minimum

3.2.3.1.2 Fly Ash

Fly ash is a by-product of coal combustion in power stations during electric power
generation and offers environmental advantages when it replaces other cementitious
products. The hydration of cement produces a base medium (a supply of calcium
hydroxide) and activates the pozzolanic reaction as:

3 [Ca(OH)2] + 2 [SiO2] = [3(CaO) 2(SiO2)3(H2O)]

3 [Ca(OH)2] + Al2O3 = [3(CaO) Al2O3 6(H2O)]

A 15 kg bag of commercially available fly ash, a Blue Circle® product, which had the
chemical compounds shown in Table 3.8 was used with GB cement for stabilisation. As

69
Chapter 3 Research Methodology

the selected fly ash contains pozzolanic compounds, silica oxide alumina oxide, and
iron oxide more than 90%, it is classified as class F flyash based on ASTM C 618

Table 3.8: Chemical composition of fly ash used for experiments

Chemical
Percentage (%)
compound
SiO2 65.9
Al2O3 24.0
Fe2O3 2.87
CaO 1.59
K2O 1.44
Na2O 0.49
MgO 0.42

3.2.3.1.3 Slag–Lime (SL)

Slag, which is a by-product of iron and steel manufacturing processes, is a


supplementary cementitious material similar to fly ash that can be used to partially
replace clinker in cement production. Slag reacts in the presence of base media, such as
quicklime (CaO), lime slurry and hydrated lime (Ca(OH)2) which are three forms of
lime available for soil stabilisation. Lime provides the catalyst necessary to activate the
cementitious reaction of slag to produce a cementitious paste.

In this study, slag with hydrated lime in proportions of 85/15 (slag/lime), the same
proportions as commercially produced and distributed by Blue Circle Southern Cement
Pty Ltd. and marketed under the proprietary name ‘Stabilment’ in Australia, was used.
Typical oxide compositions of slag and Portland cement was presented in Table 2.4 in
which that slag possesses significant CaO that enables it to achieve greater strength than
a similar quantity of fly ash.

70
Chapter 3 Research Methodology

3.3 Laboratory Testing Methods

Durability and moisture susceptibility tests were performed as part of the mix design,
with the outcomes expected to satisfy the minimum requirements specified for each
element testing based on experience and field observations. The selected laboratory
testing methods comprehensively analysed a lightly stabilised granular material by
focussing on its different properties. The selected testing methodologies included simple
UCS to complicated RLT tests in order to obtain wide acceptability from different
interest groups. Different variables that could potentially affect durability were
investigated to find a better testing procedure that could be used in the laboratory to
study durability. A brief description of each test is presented here and details discussed
in the next four chapters where relevant.

3.3.1 Freeze-Thaw Durability Testing

Three element tests were used to evaluate the f-t durability of two lightly stabilised
granular materials. After curing, all the specimens were subjected to continuous f-t
cycles in an environmental chamber before being tested for their mechanical properties.
A brief description of each test is given in this section and details discussed in
subsequent chapters.

3.3.1.1 Assessing Freeze-Thaw Durability using Unconfined Compression


Strength Test

As the standard method for assessing the durability of stabilised materials, ASTM D560
(2003), has very limited use because of the ambiguity associated with its brushing
portion, many studies replaced UCS instead of measuring weight loss (Shihata and
Baghdadi 2001). Defining durability criteria by specifying minimum UCS values are
very common by road agencies and these values are often specified as some form of
UCS; for example 7-day or socked UCS. As the literature reveals only a limited number

71
Chapter 3 Research Methodology

of studies of lightly stabilised granular materials, in this study, a suitable procedure for
assessing their durability using UCS was performed through series of UCS tests.

The durability of a lightly stabilised granular material against f-t cycles was assessed
using cylindrical specimens of 105 mm in diameter and 115 mm height which were
prepared based on AS-5101.4 (2008) and cured for the required numbers of days. They
were then subjected to different numbers of f-t cycles and tested using displacement-
controlled loading. Four series of tests were performed with different variables to assess
their impacts and establish the best set of variables that could form a better laboratory
durability test method.

3.3.1.2 Assessing Freeze-Thaw Durability using Repeated Load Triaxial Test

A RLT test is preferred over other testing methods because of its capability to simulate
the in-situ stress and loading conditions in a laboratory. The stress-dependent
mechanical behaviours of materials are the main input parameters in a design and their
dependency on f-t, moisture and w-d cycles is paramount.

In this study, RLT tests were used to investigate both the resilient and permanent
deformation behaviours of the lightly stabilised granular base material after different
numbers of f-t cycles. Cylindrical specimens 195 mm height and 100 mm in diameter
were selected for the study and all were compacted with 1.5% of CF binder. Details of
the sampling procedure are provided in the next chapter.

Based on a preliminary investigation, the resilient modulus testing procedure specified


for an unbound granular material had to be amended to suit the lightly stabilised
granular material. Each RLT test was performed by keeping the confining pressure
constant and cycling the deviatoric stress by applying a cyclic frequency of 10 Hz with
a 0.9s rest period. On-sample deformations were used to obtain precise axial
deformation measurements with the collected deformation and stress measurements
used for calculations. The outcomes from the experiments are presented in both tabular
and graphical formats for better explanations.

72
Chapter 3 Research Methodology

3.3.1.3 Assessing Freeze-Thaw Durability using Flexural Beam Test

Fatigue failure is the main distress mode for cemented materials. Although lightly
stabilised granular material has a limited flexural capacity it is one quality that
distinguishes it from unbound materials. A few studies focussing on the durability of the
flexural properties of cemented materials are reported in the literature and there are
none relating to a lightly stabilized granular material.

Therefore, in this study, beam specimens of 76 mm high, 76 mm wide and 285 mm long
were tested for flexure after being subjected to different numbers of f-t cycles. In a 3-
point displacement-controlled test, the load and deflection at the middle were measured
to obtain the MoR and SSM. As a lightly stabilised granular material has limited
flexural strength, CF binder contents of 1.5% and 3% were selected for the experiments.
In Chapter 5, details of the experimental setups and results are presented.

3.3.1.4 Assessing Durability using Vacuum Saturation Test

The VST is used to evaluate the durability against f-t cycles of a material stabilised with
a pozzolan (for example, fly ash) activated with lime and lime-based activators.
According to ASTM-D593 (2011) which sets the outline for VST, after a VST, the
durability acceptance criterion for a specimen stabilised with a pozzolan should be
greater than 2.8 MPa. However, such a requirement is not applicable for a lightly
stabilised granular material and the VST’s suitability for investigating the lightly
stabilised material is unknown.

A VST involves compacting cylindrical specimens, curing them at 40oC in a sealed


container for 7days, subjecting them to a vacuum for 30 minutes and soaking them in
water for 4 hours before testing them for their UCS.

In this series of experiments, both CM and QM materials compacted at different


moisture contents, stabilised with 1.5% of CF were tested for their UCS after vacuum
saturation which were then compared against UCS obtained after 12 f-t cycles, which is
a direct durability test. Also, a 7-day UCS test with the same variables was conducted as

73
Chapter 3 Research Methodology

it is an easier test to perform. Results obtained from TST were also compared with these
tests to evaluate the acceptability of the DV as a potential durability measurement. All
these comparisons and a detailed analysis are presented in Chapter 4.

3.3.2 Wet-Dry Durability Test

Repetitive moisture fluctuations occur in pavements due to seasonal changes which


cause stabilised materials to lose their load-bearing capacities. W-d durability tests are
performed to assess a material’s capability to remain integrated after being subjected to
numerous w-d cycles.

In this study, the durability against w-d cycles of lightly stabilised granular materials
was assessed by testing the specimens after subjecting them to w-d cycles. Similar to
the f-t durability, UCS, flexural and RLT test specimens were prepared at their OMCs
and cured in the normal curing environment. After specified curing times, they were
subjected to w-d cycles and then tested for their UCS or flexural or RLT values. One w-
d cycle consisted of soaking a specimen in potable water at room temperature for a
period of 5 hours and then placing it in an oven at 160°F (71°C) for 42 hours, as
specified in ASTM D559. All the specimens were tested after different numbers of w-d
cycles and impairments of their mechanical properties compared with those of
specimens not subjected to w-d cycles. For the UCS test, CF binder percentages of 1.5%
and 3% were selected while the flexural and RTL test specimens were stabilised with
only 1.5%. Detailed descriptions of the tests and results are discussed in Chapter 5.

3.3.3 Moisture Susceptibility Test

3.3.3.1 Assessing Moisture Susceptibility Using TST

The TST determines the DVs of a soil which is a measure of the unbound water in the
soil’s matrix that is responsible for strength. It was first used in the mid1990s to identify
poorly performing unbound materials and is currently used to assess the durability of

74
Chapter 3 Research Methodology

stabilised materials. In this study, it’s used to determine the susceptibility to moisture of
a lightly stabilised material. Also, comparisons were made between other durability
studies and the TST to evaluate its potential as a laboratory durability test for studying
lightly stabilised materials.

Both the granular materials, CM and QM, were selected to investigate their
susceptibility to moisture after being stabilised with CF and SL. Cylindrical specimens
prepared with different moulding moisture contents were used for the TST and their
DVs measured using a Percometer® and surface probe. The binder content was also
varied to evaluate its impact on DVs. After the TST, the same specimens were tested for
their UCS to evaluate the correlation between their DVs and UCS. Also, two different
types of TST setup and three different types of porous disks were compared to
investigate their influence on DVs. Further details of the testing procedure and a
comprehensive discussion of the results are presented in the next chapter.

3.3.3.2 Assessing Moisture Susceptibility using Unconfined Compressive Strength


and Flexural Test

This was performed by changing the post compaction moisture content (PCMC) of the
cured UCS and Flexural specimens and tested them at different moisture content to
obtain their properties. Different moisture contents were achieved by drying the
specimens in an oven for different time period. The measured mechanical properties
were reported along with the obtained moisture content and its influence on the
properties were discussed.

3.3.4 Characterising Deformation Responses

To characterise the PD of lightly stabilised granular material, RLT and APM testings
were performed. In the RLT test, CM stabilised with 0.5% and 1.5% of CF and cured
for 7 days. Specimens stabilised with 0.5% of CF were tested for moisture contents of
8% and 6.5% for four and three different confining pressures respectively and three
moisture contents of 8%, 6.5% and 5% were selected for specimens stabilised with

75
Chapter 3 Research Methodology

1.5% of CF binder at one confining pressure of 50 kPa. The responses were analysed to
using shakedown theory for classifying the response and to obtain equations defining
PSL and PCL. Influence of moisture content on resilient modulus and relationship
between permanent and elastic deformations were also presented.

The APM test involved testing a model pavement which had a diameter of 1.1875 m
and was compacted with 600 mm of the subgrade material and 150 mm of CM
stabilised with 1.5% of CF. Four sequences of cyclic stresses at a 2 Hz frequency were
applied at different amplitudes and the deformation responses measured at different
depths, with pressures at different locations measured using earth pressure cells (EPC).
The deformation data obtained were used to back-calculate the moduli of both the
subgrade and base using a FE model developed using ABAQUS. The moisture contents
at the surface and interface were measured using four moisture probes throughout the
test. The results from the comprehensive instrumentation of the APM test and
comparisons of them with those from the FE model and repeated load testing were
expected to reveal some interesting findings

3.4 Summary

This chapter provided brief descriptions of each testing procedure used to assess the
durability and moisture susceptibility of lightly stabilised materials, including durability
methods using the UCS, RLT and flexural tests and another series consisting of the
VST, UCS after 12 f-t cycles and 7-day UCS. A brief discussion of characterising the
lightly stabilised granular material’s response using RLT and APM were also presented.
Apart from the testing procedures, the research methodology, granular materials
selected for the experiments and binders used for stabilisation were discussed. Details of
the physical and chemical properties of the materials and justification for the binders
selected and their amounts were also presented.

76
4 Chapter 4

Evaluating Suitability of Direct and Indirect Laboratory


Tests for Assessing Freeze-Thaw and Moisture Effects on
Lightly Stabilised Granular Materials

4.1 Introduction

The unbound granular base course of a flexible pavement with a thin asphalt layer acts
as a structural layer and distributes the traffic load to the subgrade. The aggregates in an
unbound granular base generally consist of crushed stone or quarry rock, sand, gravel
and/or any other durable materials of mineral origin. However, due to the scarcity of
high-quality aggregates, as a replacement, low-quality aggregates stabilised with
cementitious binders have also been in use for many years. Stabilisation not only
improves the mechanical properties of a material but also its durability. While there are
different laboratory tests available for assessing the durability against f-t cycles of
stabilised soils this chapter focusses on the durability of granular base materials
stabilised with low percentages of binders which is referred to as lightly stabilised
granular material.

The main objective of this investigation was to compare the results from different
laboratory tests and assess their acceptability for evaluating the durability of lightly
stabilised granular materials. Laboratory tests, such as the f-t durability, vacuum
saturation and 7-day unconfined compressive strength (UCS) tests and tube suction test
(TST), were selected to evaluate the durability of two granular materials from two
different sources. The vacuum saturation, f-t durability and 7-day UCS tests basically
represent the UCS after f-t cycles, vacuum saturation and normal curing respectively.
The TST was developed by the Finnish National Road Administration and Texas
Transportation Institute to identify poor performing unbound aggregates due to moisture
ingress. TST measures the dielectric value (DV) of a material which is a measure of the
material’s insulation capabilities. Because of its ease of use and relatively short
duration, the TST has gained wide acceptance as a potential alternative
Chapter 4 F-T and Moisture Effects

test for assessing moisture susceptibility of granular material and later extend for
studying durability against f-t of stabilised materials (Solanki et al. 2011; Zhang and
Tao 2008). Promising correlations have shown that TST could be a more suitable and
easy method than standard durability tests which are time consuming. However, as few
studies have been conducted on a lightly stabilised granular material and there is no
standard testing method for conducting the TST. Therefore, TST was included to assess
the durability of lightly stabilised granular materials.

For the durability study, moulding moisture content was used as a primary variable as
the moisture has significant effects on the properties of the lightly stabilised material.
Generally, construction specifications require a pavement’s granular base and subgrade
to be constructed at the optimum moisture content (OMC) and maximum dry density
(MDD) (Qian 2010). However, it has been found that the moisture content beneath a
pavement reaches an equilibrium condition several years after construction and operates
at this newly attained level for the rest of its life (Aitchison and Richards 1965; Basma
and Al-Suleiman 1991; Chu 1977). Therefore, it is important to evaluate the effects of
moisture variations in base/subgrade materials. In this study, assessments of moisture
variations were performed using all four types of tests by varying the moulding
moisture contents. In addition, two sets of UCS and beam specimens compacted at
OMC and dried at 60oC to vary their moisture contents were also used to assess the
influence of moisture contents.

In this study, granular materials from two sources were selected to study their durability
when stabilised with cement-flyash (CF) and slag-lime (SL). In the TST, the DV was
measured using a Percometer® and surface probe while the UCS was measured in the
other three tests. Also, the amount of binder was varied to evaluate its impact on the
DV. The specimens used for the TST were tested for their UCSs to evaluate the
correlation between their DV and UCS values. The results from each test are discussed
separately before discussing their correlations. In the next section, details of the
experimental procedures for all four types of tests are discussed in detail.

78
Chapter 4 F-T and Moisture Effects

4.2 Experimental Program

This section presents detailed descriptions of the experimental procedures adopted for
each test with Table 4.1 presenting a summary of the testing program. Additional TSTs
conducted on specimens stabilised with SL binders and 0.5% and 3% CF and testing
details that were used for assessing the moisture influence will be discussed in the latter
part of this section.

4.2.1 Materials and Binders

Granular materials from two different sources used for pavement base constructions in
Queensland and Canberra, called QM and CM respectively, were selected for the
experiments. A preliminary study suggested that the particle size distribution or grading
of parent granular material has a considerable effect on the mechanical properties of a
stabilised material (Theivakularatnam and Gnanendran 2014). To eliminate grading
effect, a reconstituted material with consistent grading (i.e., with a predefined amount of
each particle size) was used. It must be noted that the particle size distribution selected
satisfied the grading C requirement of ASTM-D1241-07 (2007). The cementitious
binders CF and SL were added to stabilise the granular base materials. The weight ratio
between cement:flyash and slag:lime was kept at 3:1 and 85:15 respectively. Detailed
descriptions of the physical properties of the materials were presented in Chapter 3,
section 3.2.

4.2.2 Freeze-Thaw Durability Test

The f-t durability test was based on ASTM-D560 (2003) which outlines procedures for
f-t durability studies of soil-cement mixes. According to this standard, the weight loss of
the cylindrical specimens subjected to f-t cycles has to be measured and compared
against acceptable limits. However, in this experiment, at the end of the f-t cycles, UCS
was measured, instead of weight loss.

79
Chapter 4 F-T and Moisture Effects

Table 4.1: Testing program for moisture and durability test

Type of
Moulding
Granular binder F-T Vacuum 7-day
moisture TST 1
base & durability saturation UCS
content &2
material binder test test test
(%)
%
8.0    
6.5    
CM 1.5% CF
5.0    
3.5    
9.0    
7.5    
QM 1.5% CF
6.0    
4.5    

To determine the UCS of the lightly stabilised granular material subjected to f-t cycles,
cylindrical specimens of 105 mm in diameter and 115 mm in height were prepared in a
standard compaction mould. For each specimen, approximately 2500 grams of oven-
dried granular material was proportioned based on its reconstituted particle size
distribution. Then, the required amount of CF binder was measured and dry-mixed with
the granular material for 2 to 3 minutes. Finally, an amount of water based on the
moulding moisture content (the selected moisture contents shown in Table 4.2) was
added to the dry-mix and mixed thoroughly to form a uniform mix.

A cylindrical split mould applied with a releasing oil to facilitate the easy removal of
specimen was chosen to cast 115 mm height specimen. The stabilised mixed was
poured in three equal layers, each of which was given the standard Proctor compaction
effort according to its moulding moisture content, as shown in Table 4.2. Each
compacted sample was kept in the mould for 24 hours to gain strength and avoid
disintegration. Then samples were removed from the mould after 24 hours and wrapped
with a polythene bag with vent holes and placed in a controlled environment at a 23 ±
2˚Celcius (C) temperature and 95 ± 5% relative humidity for curing. At the end of the

80
Chapter 4 F-T and Moisture Effects

designated curing period, samples were taken from the curing room and subjected to f-t
cycles.

To simulate f-t cycles, an environmental chamber with an automated capability to


operate both freezing and thawing temperatures was chosen. One f-t cycle comprised a
freezing phase of -23°C for 24 hours and a thawing phase of 21°C for 23 hours at a
relative humidity of 95%. In the thawing phase, moisture pads were kept underneath the
specimens to supply moisture. Further details of the f-t cycles are discussed in Chapter
5, section 5.2.4.1.

After 12 f-t cycles, UCS testing was carried out using monotonic loading at the rate of
0.5mm/min to obtain the failure load and determine UCS.

Table 4.2: Compaction efforts to achieve maximum dry density

Moulding moisture
Achieved relative
content Compaction effort
compaction
(CM or QM)
8.0% or 9.0% 25 blows 100%

6.5% or 7.5% 33 blows 99%

5.0% or 6.0 % 40 blows 98%

3.5 % or 4.5 % 50 blows 96%

4.2.3 7-day Unconfined Compressive Strength Testing

It is true that the 7-day UCS has a very minimal direct relationship with durability,
based on the concept of ‘higher the strength, higher the durability’ it has been widely
used by many road agencies to define durability requirements. Shihata and Baghdadi
(2001) summarised the minimum durability requirements defined by different agencies
based on the UCS. Apparently, the reasons for using a 7-day UCS test are the ease of
performing the test and there being a well-established UCS data base as well as the
incentive of obtaining an outcome in only 7 days. Moreover, 7-day UCS test results
have shown promising correlations with those from other durability tests (Shihata and

81
Chapter 4 F-T and Moisture Effects

Baghdadi 2001). Therefore, in this study, given its wide acceptability, the 7-day UCS
test was included with other durability tests.

The method for preparing the specimens was similar to that for the f-t durability test
explained in section 4.2.2, with the same specimen sizes used for 7-day UCS. However,
they were not subjected to f-t cycles but, instead, tested as soon as their curing periods
finished. Their measured UCSs were compared against those from the other durability
tests to determine the relationship between them.

4.2.4 Vacuum Saturation Test

The ASTM-D593 (2011) VST, which is used to assess the qualification of pozzolans,
has received attention as a potential test for assessing durability against f-t cycles. Its
testing procedure comprises subjecting cylindrical specimens to a vacuum and
saturating them with water before testing their UCSs. In the next section, details of the
experimental procedures for the VST are explained.

4.2.4.1 Experimental Procedure

The VST was conducted on specimens of a similar size to those of the UCS test, i.e.,
105 mm in diameter by 115 mm height. The following steps were undertaken for
sample preparation and testing.

 Cylindrical specimens were prepared similarly to those for the f-t durability and
7-day UCS tests discussed in section 4.2.2 with moulding moisture contents and
compaction efforts shown in Tables 4.2.

 Each compacted specimen was kept in the mould for 1 day and then cured at
400C in a sealed container for 7 days. The sealed container with a specimen is
shown in Figure 4.1. It is important to note that curing in a sealed container was
essential to prevent excessive moisture loss and it was found that this resulted in
losses of moisture of only approximately 15 to 20 grams. If the specimens had

82
Chapter 4 F-T and Moisture Effects

been cured in an open environment at 400C, the moisture retained would have
been only 1% to 2%.

 At the end of curing, the specimens were removed from the oven and kept at
room temperature for approximately 30 to 45 minutes to let them cool down.
Submerging them while they were warmer than the water could make fissures
within the specimens which needed to be avoided.

 Then, each specimen was placed in a vacuum chamber at a sufficient height to


ensure that it also received water from the bottom surface. As shown in Figure
4.2, a cylindrical mould with a mesh on top used to support the specimens. The
vacuum saturation chamber had three separate parts, a cylinder, and top and
bottom platens. The cylinder had an internal diameter of 187 mm, height of 275
mm and thickness of 6 mm. The top and bottom lids were steel circular platens
20 mm thick and 200 mm in diameter. There were six steel rods equally spaced
running from the bottom to top platen along the circumference of the cylindrical
surface.

 Both the top and bottom platens had a circular groove which could
accommodate O-rings of 195 mm – 200 mm. The cylindrical surface and top
and bottom platens were sealed using the O-rings and steel rods fastened at the
top and bottom with 1⁄4-inch threads. Failure to have an air-proof system could
result in water leakage and difficulty in holding the vacuum pressure.

 Then, a vacuum of 75 kPa was applied for 30 minutes through a vacuum line
connected to a suction pump with a capacity of -100 kPa. Although the standard
suggests that a vacuum of 80 kPa needs to be applied, the selected system could
only reach 75 kPa.

 After 30 minutes, while the vacuum was turned on, a de-aired water line
connected to a water chamber placed at a certain height from the vacuum
chamber supplied water which was allowed to flow through the other opening

83
Chapter 4 F-T and Moisture Effects

on top of the lid. It was important to apply the vacuum while filling the
container with water to increase the degree of saturation.

 Once the specimens were completely submerged in the water, the vacuum was
released and they were saturated for 5 hours. They were then removed from the
chamber and kept outside for 10 to 15 minutes for the water to drain off them
and then tested for their UCS in a similar way to the 7-day UCS test.

Figure 4.1: Sealed container used for curing specimens tested for vacuum
saturation test

84
Chapter 4 F-T and Moisture Effects

Cylinder

Vacuum pump

Air and water line

Bottom plate Cylindrical mould to


support specimen Top plate with opening for
air and vacuum

Figure 4.2: Pictorial view of vacuum saturation equipment

4.2.5 Tube Suction Test

Another laboratory test considered for investigating the durability of lightly stabilised
granular materials is the TST. Unlike the other three tests, TST is an indirect test, i.e., it
does not measure mechanical properties instead measures the DV of a material which
represents the unbound water content within a specimen, not simply the total volumetric
moisture content (Barbu and Scullion 2005; Scullion and Saarenketo 1997). As
unbound water is thought to be directly related to a material’s mechanical properties, the
DV is expected to provide an indication of those properties. Currently, it is popular for

85
Chapter 4 F-T and Moisture Effects

durability studies of stabilised materials because of the ease of conducting the test.
However, as its application to lightly stabilised granular materials has been very limited,
in this study, it was included with other f-t durability studies to assess its suitability as a
potential f-t durability test.

Table 4.1 presented part of the TST testing program and Table 4.3 shows the additional
tests performed on specimens stabilised with 0.5% CF and SL, using the testing
procedure explained in the next section.

Table 4.3: Additional TST program

Moulding
Granular base Type of binder
moisture content TST 1 & 2
material & binder %
(%)
8.0 
6.5 
CM 0.5% CF
5.0 
3.5 
9.0 
7.5 
QM 0.5% CF
6.0 
4.5 
CM 3% CF 8.0 
QM 3% CF 9.0 
3.0% SL
CM & QM 4.5% SL 8% or 9% 
6.0% SL

4.2.5.1 Experimental Procedure

A detailed description of the testing procedure adopted for TST is presented in this
section. It is important to note that TST does not have a standard testing procedure and
the methodology adopted for this experiment was based mainly on earlier studies
available in the literature. The basic concept of any TST test is that dried specimens are
allowed to absorb water through capillary action, with the DVs on top of their surfaces
measured.

86
Chapter 4 F-T and Moisture Effects

One of the earlier TST test setups suggested by Scullion and Saarenketo (1997) for
aggregates with maximum particle sizes of 25 mm was for specimen sizes of 305 x 152
mm. As a subsequent study by Barbu et al. (2004) of three different silts indicated that
the specimen diameter has no significant influence on the DV, the same size specimen
as that used in the standard UCS test (i.e., 105 mm in diameter and 115 mm height) was
chosen for TST as well. The maximum moulding moisture contents for CM and QM
were their OMCs, that is, 8.0% and 9.0% respectively. As shown in Table 4.1, three
moisture contents lower than the OMC were also chosen to study the impact of the
moulding moisture content on the DVs.

Although the standard Proctor hammer compactor was used for compaction, the
numbers of blows changed (increased with decreasing moisture content) to achieve a
relative density close to 100%, as discussed earlier in this section. The compacted
specimens were kept in moulds for a day to gain strength in order to avoid
disintegration while being un-moulded. Then, they were cured for 7 days in a curing
room maintained at a constant temperature of 23o ± 1o Celsius (C) and approximately
95% relative humidity. After 7 days, they were oven-dried at 40o C for 3 days to remove
any excess moisture. Although selection of drying temperature was based on findings
published in the literature, drying at 60o C or until the weight became constant has also
been reported (Barbu et al. 2004; Little 2000; Yeo et al. 2012). As it was found from the
preliminary study that 3 days of drying at 40oC removed a significant amount of
moisture from a specimen, as shown in Figure 4.3, this was selected for the
experiments. Thereafter, the specimens were kept at room temperature for an hour to
cool down and reach equilibrium with the environment, and then their DVs were
measured using two different experimental setups.

Table 4.4 presents a summary of previous TSTs performed mainly on stabilised


materials. As previously explained, significant differences in their testing procedures
can be noticed and their final testing arrangements also varied considerably. While most
TST setups allowed water infiltration through only the bottom surface, some studies
exposed part of the cylindrical surface (10 mm to 20 mm) from the bottom to water
(Barbu and Scullion 2005; Syed et al. 2000; Yeo et al. 2012). In this study, two
experimental setups, one for each type, were selected as explained in the next section.

87
Chapter 4 F-T and Moisture Effects

The TST experimental setups used for this investigation are shown schematically in
Figures 4.4 to 4.7 along with photographs of the specimens. In TST setup 1, the curved
peripheral surface was sealed with latex rubber to prevent any moisture loss or gain
through the cylindrical surface with water allowed to infiltrate through only the bottom
surface. In TST setup 2, apart from the bottom surface, water was allowed to pass
through the 20 mm high peripheral surface with the top surface covered with thin
polythene between readings to prevent moisture loss through the top surface.

The DV measuring devise consisted of an Adek Percometer™ and a 50 mm diameter


surface probe shown in Figure 4.8. The first reading was taken prior to placing the
specimens in the water and the second after 3 hours, after which the DVs were
measured every 24 hours for 10 days. Each DV reading represents an average value of
five readings taken on the top surface of a specimen. As shown in Figure 4.9, four
measurements were taken along the circumference of a specimen to represent four
separate quadrants and the fifth at the centre of the specimen.

It was observed that the consistency of the five measurements was notably affected by
the smoothness of the top surface as an uneven surface could create gaps which would
essentially be exposed to the air and, as the DV value of air is small, it is likely that the
DVs would be small. Barbu and Scullion (2005) recommended that, if DVs differ
considerably, the two highest readings should be used. However, in this study, to avoid
unevenness, the top surface of each specimen was smoothed using finer particles (< 0.3
mm). Two specimens were prepared for each combination of binder type and moulding
moisture, and the average of their values used for comparisons.

88
Chapter 4 F-T and Moisture Effects

Moisture Loss %

Figure 4.3: Percentages of moisture loss with number of days

105 mm diameter 115 mm


height specimen

Liquid rubber
membrane

Porous disk

Aluminium tray

Water

Figure 4.4: TST experimental setup 1

89
Chapter 4 F-T and Moisture Effects

Figure 4.5: Specimen prepared for TST experimental setup 1

105 mm diameter 115 mm


height specimen

Liquid rubber membrane

Porous disk Aluminium tray

Water

Exposing peripheral surface (20 mm)


for water ingress
Figure 4.6: TST experimental setup 2

90
Chapter 4 F-T and Moisture Effects

Figure 4.7: Specimen prepared for TST experimental setup 2

Figure 4.8: Percometer and surface probe used for DV measurement

91
Chapter 4 F-T and Moisture Effects

5
2
4

Figure 4.9: Five surface probe reading locations

4.2.6 Flexural and UCS Testing to study Effects of Moisture Content

For the flexural tests, two series of specimens 76 mm height, 76 mm width and 285 mm
length were compacted at the OMC (9%), stabilised with 1.5% of CF binder and cured
for 14 and 28 days. The detail description of the flexural beam test is presented in
section 5.2.3. The UCS test procedure was similar to those explained in this section
except for variations in the moisture content. Similarly, for UCS testing, only one set of
cylindrical specimens 115 mm height 105 mm in diameter were compacted at the OMC
(9%), stabilised with 1.5% of CF and cured for 14 days.

After curing, the specimens were taken from the curing room and placed in an oven at
60oC to vary their moisture contents. To achieve different moisture contents, they were
kept in the oven for different durations. Then, UCS and flexural tests were conducted
using the experimental setups described in Chapters 4.2.3 and 5.2.3 respectively.

92
Chapter 4 F-T and Moisture Effects

Table 4.4: Details of TST methodologies obtained from published literature

Curing and drying


Author Material Specimen size Testing
conditions
Unbound material compacted in plastic
Saarenketo and 11 different types of 152 mm diameter tubes with 1 mm diameter holes 10 mm
500C for 10 days
scullion (1996) aggregates by 305 mm height from bottom surface, with water depth
changed to different levels
Compacted in 2 types of plastic tubes, one
150 mm diameter 400C for 3-4 days with 2 mm diameter holes in bottom surface
Saarenketo Unbound base course
by 180 to 200 mm until weight became and other with 2 mm diameter holes 10 mm
(2000) aggregates
height constant from bottom surface, with water depth 10
to 20 mm
River gravel, crushed
limestone base material,
102 mm diameter 3-4 days at (400C)
Syed et al. recycled concrete with low Cylindrical surface exposed to 6 mm water
by 117 mm height room until weight
(2003) (1.5 %), medium (3%) and height and DVs measured for 10 days
No plastic mould became constant
high (4.5%) amounts of
cement
Plastic tube used to compact specimens and
305/180 mm
Barbu et al. Silt, silty sand and clay 500C until no more aluminium foil with holes to cover surface,
height by 152 mm
(2004) stabilised with Portland significant changes in with measurements taken until DV became
diameter with
cement weight constant
plastic tube

93
Chapter 4 F-T and Moisture Effects

Curing and drying


Author Material Specimen size Testing
conditions

203 mm height
Barbu and Specimens placed on top of porous disk and
Two base granular materials and 154 mm 600C for 2 days
Scullion (2005) DVs measured for 10 days
diameter
Dried specimens placed in plastic tubes with
Low plastic silt clay 1 day of curing and
100 mm diameter holes in bottoms and then placed in a large
Zhang and Tao stabilised with different kept at 400C for 14
by plastic container with porous stone
(2008) moisture contents and cement days until weight
180 mm height underneath and 20 mm water above bottoms
percentages almost constant
of specimens
Three specimen
Clay soils stabilised with 6%
sizes ranging from Cured for 7 days and Cylindrical surface coated with grease and
Solanki et al. hydrated lime (or lime), 10%
100-150 mm dried at 400C for 2 placed on porous stone with 10 mm water
(2011) class C fly ash (CFA) and
diameter by 100- days height and DVs measured for 10 days
10% cement kiln dust (CKD)
200 mm height
105mm diameter Cured for 7 days and Plastic membrane used to wrap specimens
Crushed rock aggregates
Yeo et al. by kept for to prevent moisture loss from cylindrical
stabilised with 1% to 6%
(2012) 115mm height 24 hours in oven at surface although 5 mm of cylindrical
cement
No plastic mould 60oC surface from bottom exposed to water
52.4 mm diameter
(Gandara et al. by Specially developed ice chest were used to
Unbound granular material 600C for 2 days 
2005) 203.2 mm height inject moisture within the specimens
 

         

94
Chapter 4 F-T and Moisture Effects

4.3 Results and Discussion

The results obtained from the experiments discussed in section 4.2 are presented here.
First section of the test result is focussed on individual test results whereas the latter part
of this section is allocated to compare the results between the tests.

4.3.1 Freeze-Thaw Durability Test Result

Initially, the first set of specimens was tested for f-t durability based on the procedure
explained in section 4.2.2, where the UCS of cylindrical specimens were measured after
12 f-t cycles. The specimens were stabilised with 1.5% CF binder and compacted at
four different moisture contents. Table 4.5 and Figures 4.10 and 4.11 show the
measured UCS of CM and QM respectively.

The effects of f-t cycles on UCS can be seen from the results comparison between UCS
obtained from the f-t durability test and the 7-day UCS test. UCS obtained from f-t
durability showed lower values than that of 7-day UCS for the same combination of
variables.

Furthermore, the effect of the moulding moisture content is clearly visible from the
UCS results for both CM and QM. The specimens stabilised with 5% showed the
highest UCS for CM, 23% higher than those of specimens compacted at the OMC,
while those stabilised with 6.5% also showed marginal increase over the OMC but
remained significantly lower than those with 5%. However, further reduction in the
moisture content to 3.5% did not allow the material to gain more strength. QM also
showed increase in UCS from 1.47 MPa to 1.67 MPa with decrease in the moisture
content from 9% to 7.5%. However, further reduction in the moisture content resulted in
declining strength.

The implication of the moisture content in terms of f-t effects is discussed in detail in
the next chapter. To give a brief in the context of this experiment, it is safe to assume
that the decrease in moisture content is likely to produce minimal adverse f-t effects.
Moreover, the effective stress, which relies on the degree of saturation, in this case the
moisture content, also needs to be considered when interpreting results.
95
Chapter 4 F-T and Moisture Effects

4.3.2 7-day Unconfined Compressive Strength Test Results

Although the UCS at the end of 7 days is not directly related to the durability of a
material, it is used by many road agencies to specify the minimum durability
requirements for a stabilised material (Solanki et al. 2011). A good correlation between
the results from 7-day UCS and other durability tests will be very useful for obtaining
good estimates in the absence of durability test results.

The results from the test demonstrated that the 7-day UCS was very dependent on the
moulding moisture content, as shown in Table 4.5 and Figures 4.12 and 4.13, with the
UCS of CM increasing from 1.9 MPa to 2.7 MPa for moisture content of 8% and 5%
respectively, and QM also indicating the same dependency.

4.3.3 Vacuum Saturation Test Result

The VST results are shown in Figures 4.14 and 4.15 for CM and QM respectively.
Curing the specimens at 400C accelerated the strength gain and resulted in higher
compressive strengths being obtained than normally cured specimens. Similar to the f-t
durability and 7-day UCS tests, the influence of the moulding moisture content can be
clearly seen in these two figures as CM and QM showed higher UCSs for specimens
stabilised with 5% and 7.5% moisture contents respectively.

One of the main advantages of UCS tests is the consistency of the experimental data
and, as shown in Table 4.5, the specimens’ CoVs remained in single digits, except for
one set of specimens. Compared with other mechanical properties, measuring the UCS
does not require a great deal of expertise and/or resources which is one reason why it
remains a popular testing method even today.

Moreover, a simple comparison of the absolute values indicated that all three tests are
likely to show good correlations. It will be interesting to see the outcomes of
comparisons with the DVs which are discussed in the next section.

96
Chapter 4 F-T and Moisture Effects

Table 4.5: Experimental results from different moisture and durability tests

Average
Binder UCS Average Average
Moisture UCS UCS 7-day
Granular and f-t UCS 7-day Moisture
content f-t CoV VST CoV UCS CoV DV UCS_TST
material binder durability VST UCS %
(%) durability (MPa) (MPa)
content (MPa) (MPa) (MPa)
(MPa)
1.69 1.93 1.83
11.1 1.52 7.2
8 (OMC) 1.51 1.62 6.1 1.88 1.90 1.4 1.82 1.74 9.1
10.6 1.57 7.3
1.67 1.89 1.55
2.07
1.64 1.83
10.1 1.81 6.1
6.5 1.64 1.65 1.0 2.12 5.1 1.83 1.80 2.9
2.24 11.5 1.65 6.2
1.5% 1.67 1.74
CM 2.04
CF
1.81 2.70 1.97
10.4 1.81 5.9
5.0 2.41 2.11 14.3 2.79 2.70 3.3 2.01 2.01 2.2
10.4 1.99 5.8
2.04 2.61 2.06
1.14 1.83 1.59
11.0 1.27 7.5
3.5 1.04 1.12 6.4 1.55 1.62 11.4 1.36 1.48 7.8
11.0 1.44 7.5
1.18 1.48 1.48
1.40 1.76 1.66
16.1 0.99 7.6
9 (OMC) 1.49 1.47 4.2 1.87 1.85 4.0 1.73 1.71 2.4
17.1 1.10 8.0
1.52 1.92 1.73
1.62 2.32 1.87
13.3 1.57 7.0
7.5 1.76 1.67 4.7 2.09 2.17 4.6 1.84 1.88 2.1
13.9 1.82 6.8
1.5% 1.63 2.10 1.92
QM
CF 1.56 1.73 1.62
18.2 1.92 7.1
6.0 1.59 1.59 2.2 1.82 1.73 5.2 1.71 1.66 2.7
18.0 1.73 8.6
1.63 1.64 1.66
1.19 1.48 1.49
18.2 0.76 7.8
4.5 1.40 1.32 8.6 1.51 1.49 1.0 1.39 1.42 4.1
19.7 0.76 8.3
1.37 1.49 1.39

97
Chapter 4 F-T and Moisture Effects

2.4

UCS (MPa) 1.6

1.2

0.8

0.4

0
8% 6.5% 5% 3.5%
Moisture content (%)

Figure 4.10: UCS of CM specimens stabilised with 1.5% CF and subjected to 12 f-t
cycles varying with different moulding moisture contents
2

1.6

1.2
UCS (MPa)

0.8

0.4

0
9% 7.5% 6% 4.5%
Moisture content (%)

Figure 4.11: UCS of QM specimens stabilised with 1.5% CF and subjected to 12 f-t
cycles varying with different moulding moisture contents

98
Chapter 4 F-T and Moisture Effects

2.8

2.4

UCS (MPa) 1.6

1.2

0.8

0.4

0
8% 6.5% 5% 3.5%
Moisture content (%)

Figure 4.12: UCS of CM specimens stabilised with 1.5% CF and subjected to


vacuum saturation varying with different moulding moisture contents

2.4

1.6
UCS (MPa)

1.2

0.8

0.4

0
9% 7.5% 6% 4.5%
Moisture content (%)

Figure 4.13: UCS of QM specimens stabilised with 1.5% CF and subjected to


vacuum saturation varying with different moulding moisture contents

99
Chapter 4 F-T and Moisture Effects

1.6

UCS (MPa) 1.2

0.8

0.4

0
8% 6.5% 5% 3.5%
Moisture content (%)

Figure 4.14: 7-day UCS of CM specimens stabilised with 1.5% CF varying with
different moulding moisture contents

1.6

1.2
UCS (MPa)

0.8

0.4

0
9% 7.5% 6% 4.5%
Moisture content (%)

Figure 4.15: 7-day UCS of QM specimens stabilised with 1.5% CF varying with
different moulding moisture contents

100
Chapter 4 F-T and Moisture Effects

4.3.4 Tube Suction Test Results

As noted in section 4.2.5, the TST was performed on 105 mm diameter and 115 mm
height cylindrical specimens. Two testing setups and variables such as the moulding
moisture content, binder content, type of binder and type of granular material were used
for the test. The stabilised specimens were cured for 7 days before being dried at 400C
for 3 days. Then, their DVs were measured for 10 days using an Adek percometer. In
the following section, the results obtained from the TST for evaluating the influence of
these factors on DVs are discussed in detail.

4.3.4.1 Effect of Moulding Moisture Content on Dielectric Value

To investigate the effect of the moulding moisture content on the DV, specimens of QM
and CM were prepared at different moulding moisture contents shown in Table 4.1. The
OMCs of both materials were selected as the highest moisture content, with three lower
moisture contents also added to the combination. Triplicate specimens were prepared
for each moisture contents, with two tested using TST setup 1 and the remaining one
using TST setup 2, and CF binders of 1.5% and 0.5% selected as stabilising agents. As
previously mentioned, the DVs were taken for 10 continuous days and their 10th day
DVs tabulated in Table 4.6. The measured DVs varying with the number of days for all
the selected combinations are shown in Figures 4.16 to 4.23. At the completion of the
TST, the specimens were tested to determine their UCSs although, as most of those
stabilised with 0.5% binder were damaged, particularly at the bottom, once they were
removed from the porous disks were not tested. The TST revealed the following results.

 As shown in Table 4.6, the impact of the moulding moisture content on the DVs
could be noticed in the final DV readings, the lowest of which were achieved at
5% for CM and 7.5% for QM and were 3% and 1.5% lower than their
respective OMCs. It should be noted that the dry densities of all the different
moisture contents were maintained at greater than 96% by varying the
compaction effort to eliminate the effect of density.

101
Chapter 4 F-T and Moisture Effects

 Overall, the lowest moisture contents of both materials were weak which shows
that reducing the moisture content (starting from the OMC) up to a certain level
increases the performance of the material. However, any further reduction is
likely to cause adverse results.

 QM showed higher DVs than CM (when comparing their OMCs and with their
respective lower moisture content levels) possibly due to the coarser particle
size distribution of CM.

 Specimens tested using TST setup 2 showed higher DVs than those in TST
setup 1. As previously mentioned, in TST setup 2, as well as from the bottom
surface, moisture was allowed to enter through the 2 mm peripheral surface
from the bottom would have probably added moisture which was reflected in
the DVs.

 The DVs also showed that the amount of binder content had an impact on them
as, for both CM and QM, the 1.5% binders had lower DVs than the 0.5% ones.
In section 4.3.4.2, additional results obtained using different binders and binder
contents are provided.

 To obtain an understanding of the absolute DVs presented in Table 4.6, the


following classification proposed by Guthrie and Scullion (2003) can be
referenced. However, as the classification is applicable to only to raw coarse
aggregate materials, it is expected that the limits for stabilised materials would
be lower. It must be noted that, as the testing procedure adopted by Guthrie and
Scullion (2003) was different from the method used in this study, comparing
their DVs would likely give erroneous impressions. However, the DVs of QM
in particular would be unlikely to satisfy durability requirements since they
were considerably higher than the maximum allowable limit of 16.

Final DV < 10 Good


10 < Final DV <16 Marginal
Final > 16 Poor

102
Chapter 4 F-T and Moisture Effects

Table 4.6: DVs measured with different moulding moisture and binder contents

TST Setup 1 TST Setup 2


Moulding
Granular Binder moisture UCS UCS
Moisture Moisture
material content content Final after Final after
content content
% DV TST DV TST
% %
(MPa) (MPa)
8 (OMC) 10.9 1.55 7.3 14.9 1.46 7.1
Canberra 6.5 10.8 1.73 6.2 11.2 1.71 6.9
1.5%
Material 5 10.4 1.91 5.9 10.9 1.79 6.6
3.5 11.0 1.27 7.5 14.1 0.91 7.2
8 (OMC) 14.8 _ 6.6 14.9 _ 7.3
Canberra 6.5 11.5 _ 6.6 12.6 _ 7.3
0.5%
Material 5 11.6 _ 6.8 12.4 _ 7.1
3.5 15.2 _ 8.1 15.8 _ 7.6
9 (OMC) 16.6 1.04 7.8 19.5 0.73 7.7
Queensland 7.5 13.6 1.70 6.9 14.4 1.76 6.9
1.5%
Material 6 18.1 1.82 7.9 16.4 1.37 8.2
4.5 19.0 0.62 8.1 16.4 1.25 8.1
9 (OMC) 20.8 _ 8.3 22.5 _ 8.4
Queensland 7.5 18.5 _ 7.6 21.1 _ 8.0
0.5%
Material 6 22.0 _ 8.0 23.8 _ 8.4
4.5 24.2 8.3 25.2 __ 8.6

1.5%CF_CM_P1
3.5% MC 6.5% MC
12
5% MC 8% MC

10
Dielectric Constant

0 50 100 150 200 250


Time (hours)

Figure 4.16: DVs of CM stabilised with 1.5% CF varying with time, measured
using TST setup 1

103
Chapter 4 F-T and Moisture Effects

0.5%CF_CM_P1
3.5% MC 6.5% MC
16 5% MC 8% MC

12

Dielectric constant
8

0 50 100 150 200 250


Time (Hours)

Figure 4.17: DVs of CM stabilised with 0.5% CF varying with time, measured
using TST setup 1

1.5%CF_QM_P1
3.5% MC 6.5% MC
5% MC 8% MC
20

16
Dielectric constant

12

0 50 100 150 200 250


Time (hours)

Figure 4.18: DVs of QM stabilised with 1.5% CF varying with time, measured
using TST setup 1

104
Chapter 4 F-T and Moisture Effects

0.5%CF_QM_P1
3.5% MC 6.5% MC
5% MC 8% MC
25

20

Dielectric constant
15

10

0 50 100 150 200 250


Time (hours)

Figure 4.19: DVs of QM stabilised with 0.5% CF varying with time, measured
using TST setup 1

1.5%CF_CM_P2
3.5% MC 6.5% MC
16
5% MC 8% MC

12
Dielectric Constant

0 50 100 150 200 250


Time (hours)

Figure 4.20: DVs of CM stabilised with 1.5% CF varying with time, measured
using TST setup 2

105
Chapter 4 F-T and Moisture Effects

0.5%CF_CM_P2
3.5% MC 6.5% MC
16 5% MC 8% MC

12

Dielectric constant

0 50 100 150 200 250


Time (Hours)

Figure 4.21: DVs of CM stabilised with 0.5% CF varying with time, measured
using TST setup 2

1.5%CF_QM_P2
3.5% MC 6.5% MC
20
5% MC 8% MC

16
Dielectric Constant

12

0 50 100 150 200 250


Time (hours)

Figure 4.22: DVs of QM stabilised with 1.5% CF varying with time, measured
using TST setup 2

106
Chapter 4 F-T and Moisture Effects

1.5%CF_QM_P2
3.5% MC 6.5% MC
30
5% MC 8% MC

25

Dielectric Constant 20

15

10

0 50 100 150 200 250


Time (hours)

Figure 4.23: DVs of QM stabilised with 0.5% CF varying with time, measured
using TST setup 2

4.3.4.2 Tube Suction Test with Different Binders and Binder Percentages

Another set of specimens of CM and QM stabilised with two binders, CF and SL with
three binder contents of 0.5%, 1.5% and 3% for CF, and 3%, 4.5% and 6% for SL were
selected. As shown in Table 4.7 and Figures 4.24 to 4.28, reductions in the DVs with
increasing binder contents were observed. Similar findings were reported in previous
studies in which, the DV reduced with increasing binder content (Syed et al. 2003; Yeo
et al. 2012; Zhang and Tao 2008). . However, the differences between specimens of
both materials stabilised with 1.5% and 3% CF were marginal. Moreover, specimens
stabilised with SL showed higher DVs irrespective of their higher binder percentages.
Unlike CF, as SL is a weak stabiliser, i.e., its capability to bind granular solid particles
and increase strength is minimal compared with that of CF, this could have resulted in
higher DVs which provide some indication regarding the possible relationship between
DVs and strength (UCS).

107
Chapter 4 F-T and Moisture Effects

The benefit of adding more binder is not only to improve the durability of the final
product but to delay the rate of increase in DVs (Zhang and Tao 2008). Yeo and Nikraz
(2012) also reported that the rate of increase in the DV (DV/day) reduced with an
increase in the binder content; for example, specimens stabilised with a cement content
of 1% exhibited a rate of 4.59 DV/day and for 6%, only 0.17 DV/day was observed.
According to Zhang and Tao (2008), this delay is helpful when the material is exposed
to free water in the initial period of construction.

Correlating the final DV with the other mechanical properties, such as the bearing
capacity or UCS obtained after the f-t test and TST, is beneficial and becoming common
in recent studies (Barbu and Scullion 2005; Zhang and Tao 2008). The relationships
between the maximum DVs and UCSs of CM and QM obtained from this series of TST
experiments are shown in Figure 4.29, with the UCS values presented in Table 4.1.
General observation is that an increase in the UCS resulted in a lower maximum DV.
Although CM showed a good linear correlation with the DV (R2 = 87%), the data were
scattered while the correlation obtained for QM was less reliable and further research
would be needed to obtain any solid evidence.

Table 4.7: DVs measured for different binder contents

Compaction MC UCS
Granular Type of Binder Final
moisture after after
material binder % DV
content % TST TST
0.5 14.8 6.6 _
Canberra
8 CF 1.5 10.9 7.5 13.4
Material
3 10.8 7.2 32.5
3 18.0 5.8 _
Canberra
8 SL 4.5 16.9 6.1 _
Material
6 15.1 5.8 _
0.5 20.8 6.7 _
Queensland
9 CF 1.5 16.6 8.0 9.0
Material
3 16.8 7.7 30.5
3 20.4 7.4 _
Queensland
9 SL 4.5 19.0 7.4 _
Material
6 13.0 7.7 _

108
Chapter 4 F-T and Moisture Effects

25

20

Dielectric Constant
15

10

0
3% 4.5% 6% 3% 4.5% 6% 0.5% 1.5% 3% 0.5% 1.5% 3%
QBM-SL CBM-SL QBM-CF CBM-CF

Figure 4.24: Comparison of DVs of different mixes of stabilised materials

CM_CF_MC(8%)_P1
16 0.5% CF 1.5% CF 3% CF

12
Dielectric constant

0 50 100 150 200 250


Time (hours)

Figure 4.25: DVs of CM stabilised with 0.5%, 1.5% and 3% CF varying with time

109
Chapter 4 F-T and Moisture Effects

CM_SL_MC(8%)_P1
20 3% SL 4.5% SL 6% SL

16

Dielectric constant

12

0 50 100 150 200 250


Time (hours)

Figure 4.26: DVs of CM stabilised with 3%, 4.5% and 6% SL varying with time

QM_CF_MC(9%)_P1
24 0.5% CF 1.5% CF 3% CF

20
Dielectric constant

16

12

0 50 100 150 200 250


Time (hours)

Figure 4.27: DVs of QM stabilised with 0.5%, 1.5% and 3% CF varying with time

110
Chapter 4 F-T and Moisture Effects

QM_SL_MC(9%)_P1
24 3% SL 4.5% CF 6% CF

20

16

Dielectric constant
12

0 50 100 150 200 250


Time (hours)

Figure 4.28: DVs of QM stabilised with 3%, 4.5% and 6% CF varying with time

CBM
CBM Regression Line
2 QBM
QBM Regression Line

1.6
UCS (MPa)

1.2

0.8

0.4

10 12 14 16 18 20
Dielectric Constant (DC)

Figure 4.29: Relationships between UCS and maximum DV for CM and QM

111
Chapter 4 F-T and Moisture Effects

4.3.4.3 Effect of Porous Disk on Dielectric Measurement

The repeatability of DV measurements can be markedly improved by using quality


porous disks which are important in helping specimens absorb water through capillary
action. Barbu and Scullion (2005) reported significant differences between the DVs
measured using new ceramic disks and old ones with impurities clogging their pores. In
this study, to expedite the experimental work, three different types of clean porous disks
of good quality and sufficiently large to fully cover the bottom surface of a specimen
were used (Figure 4.30).

Disk 1 Thickness 7 mm Disk 2 Thickness 3 mm

Disk 3 Thickness 8 mm

Figure 4.30: Different types of porous disks used for TST test

112
Chapter 4 F-T and Moisture Effects

To ensure that the type of disk did not influence the DVs, the results for two identical
specimens were compared, with typical ones obtained from the three different types of
disks shown in Figure 4.31. As, irrespective of the types of disks used, the specimens
showed comparable results, it can be concluded that, for consistent readings, the quality
of the disk rather than its thickness or shape is important.

20

16
Dielectric Constant

12

Disk 1_QM_CF_1.5%_9.0%
Disk 2_QM_CF_1.5%_9.0%
Disk 1_QM_CF_1.5%_7.5%
8 Disk 3_QM_CF_1.5%_7.5%
Disk 2_QM_CF_1.5%_6.0%
Disk 3_QM_CF_1.5%_6.0%

0 50 100 150 200 250


Time (hours)

Figure 4.31: DVs measured using different types of porous disks

4.3.4.4 Comparison of Dielectric Values measured from TST Setups 1 and 2

As noted in section 4.2.5, two TST setups were used to measure the DVs of dried
specimens which are presented in Table 4.7 and discussed in section 4.3.4.1. The
corresponding DVs from TST1 and TST2 were paired and are presented in Figure 4.32,
with a symmetrical line (y=x) drawn to facilitate a comparison of them. It can be seen
that, except for two measurements, the DVs from TST 2 had higher values which
demonstrates that a 10 mm cylindrical surface helped to add more moisture to samples
which resulted in higher DVs. However, these differences were not significant, as

113
Chapter 4 F-T and Moisture Effects

shown in the regression equation because, overall, the results for TST 1 were 93% of
those of TST 2. Therefore, it is safe to assume that the majority of the water ingress
took place through capillary action, with the exposed holes in the cylindrical surfaces
not greatly affecting the DVs. Placing porous disks beneath specimens was a better way
of improving the flow of water through the bottom surface than keeping them on top of
the water container. Compacting specimens in plastic tubes and placing holes in their
bottom surfaces to facilitate the movement of water would be likely to result in a very
low rate of increase in the DVs. Moreover, such approach may require a longer testing
time. In summary, water flow through the bottom surface needs more attention while
the other surfaces can be sealed to prevent moisture ingress

Moreover, the DVs measured using TST setups 1 and 2 and correspondingly paired
were statistically analysed to determine the relationship between them. The linear
regression showed that they were very well linearly correlated, with R2 of 0.86 and p
values of less than 0.0001.

T-test for ‘paired two samples for means’ is a good indicator to show that obtained
mean values are different. A null hypothesis in this analysis is that the mean values
from TST 1 and TST 2 are same. It- could be rejected based on the p values of the t-test.
For this analysis, an alpha of 0.05 was selected, which is common for this type of
experiment, and the analysis found out that the p values is 0.01. As the p values of the
one tail and two tails remained significantly less than 0.05, the null hypothesis could be
rejected. The means of the DVs of TST 1 and 2 were different which showed that they
did not read the same measurements but, as previously discussed, they were statistically
related.

114
Chapter 4 F-T and Moisture Effects

28

DVs (TST 1vs TST 2)


Symmetrical line
Regression line
24

Dielectric value TST 1


20

Y(DVTST1 )=0.93X(DVTST2)
R2=0.86
16

12

8 12 16 20 24 28
Dielectric value TST 2

Figure 4.32: Comparison of DVs measured from TST setups 1 and 2

4.3.4.5 Comparison of Dielectric Values and Moisture Content

As previously discussed, in this study, DVs up to 10 days were measured which is


common in TSTs. This was based on the assumption that the amount of water ingress
through capillary rise is negligible beyond 10 days. From a practical point of view,
obtaining an outcome in 10 days is more acceptable than continuing until a constant
reading is reached. However, continuing the test until the DV becomes constant, or up
to 500 hours, has been reported in earlier studies (Barbu et al. 2004; Little 2000). In this
experiment, the first readings were taken before placing specimens in the water, the
second after 3 hours, the third after 24 hours and then every 24 hours. Regarding the
rate of increases in DVs in this set of experiments, after 3 hours, they were significant
and, after 24 hours, the specimens reached considerable DVs compared with those on
the 10th day. The rates of variation in the DVs were discussed in section 4.3.4.1. Almost
all the samples tested showed a sudden rise in their DVs on the first day followed by
mild increase or constant reading up to 10 days. The cessation of water ingress or a
constant DV reading indicated that the infiltrated water reached the top of a sample and

115
Chapter 4 F-T and Moisture Effects

achieved a steady state of equilibrium which meant that 10 days was adequate to
terminate the test.

Moreover, the moisture distribution within the specimens also justified the selection of
10 days as a suitable time period. To discover the moisture distribution within
specimens and assess the adequacy of 10 days as an appropriate time to terminate the
test, two slices from the bottom and top halves of specimens were taken and their
moisture contents measured. The moisture distributions obtained, as shown in Figure
4.33, were not skewed towards either half which showed that, as the moisture had
reached the top surface, measuring the DVs for 10 days was sufficient for assessment.
Also, from a practical point of view, obtaining an outcome in 10 days is more
acceptable than continuing up to constant reading. However, as this may not be the case
for taller specimens or those with non-uniform pore size distributions, longer testing
times may be required.

Symmetrical line

8
Moisture content_bottom (%)

5 6 7 8 9
Moisture content_top (%)

Figure 4.33: Moisture contents measured from bottom and top of specimen

In an earlier section, it was explained that DVs are not representations of volumetric
moisture contents. To further validate this, the relationship between the DV and

116
Chapter 4 F-T and Moisture Effects

moisture content was established, as shown in Figure 4.34, using the average values of
the moisture contents were presented in Tables 4.1 and 4.2. This figure relates the
moisture content measured for each individual specimen at the end of the TST to the
corresponding DV, with the regression line shown in the figure is correlated with the
average moisture content. However, although the DVs were correlated with both the top
and bottom moisture contents, neither produced a good correlation. Therefore, it can be
concluded that a DV measurement is not a direct representation of the moisture content
as it represents only the unbound water within a soil specimen.

MC vs DV
10 Moisture content_Bottom
Moisture content_Top
Regression line
R2 = 0.54

8
Moisture content(%)

5 10 15 20 25 30
Dielctric value

Figure 4.34: Correlations between moisture contents measured after TST and
respective DVs measured during TST

117
Chapter 4 F-T and Moisture Effects

4.3.5 Comparison of Durability Test Results

In sections 4.3.1 to 4.3.4, individual test results were discussed separately and in this
section, all the tests results are compared against each other to evaluate their
correlations. As shown in Table 4.8 and Figures 4.35 to 4.40, all four tests were paired
up into six combinations and linear regression analysis was performed to obtain
regression coefficient.

As shown in Figure 4.35, VST and f-t durability tests showed a good correlation for
CM, with a regression coefficient of 0.84, while QM showed 0.67. A statistical analysis,
a paired t-test, was performed defining a null-hypothesis that UCS obtained from VST
and f-t durability tests are the same. As shown in Table 4.9, the null-hypothesis was
rejected for both CM and QM based on the p-values being less than 0.05 (common
alpha used in this type of analysis). This is shows that UCS obtained from VST is
significantly higher than f-t durability test. Similarly, the vacuum saturation and 7-day
UCS tests demonstrated a convincing relationship with R2 of 0.84 and 0.93 (Figure
4.36) for CM and QM respectively and the t-test showed that mean null-hypothesis can
be rejected as shown in Table 4.10. The t-test is important in both cases to show that
vacuum saturation, f-t and normal curing produced completely independent outcomes
even though all three of them moulded with the same moisture contents and specimen
size.

The relationship between f-t durability test and 7-day is shown in Figure 4.37 where
regression coefficient of 0.86 for CM and 0.76 QM was obtained. To statistically
compare the results of both materials, a paired t-test was performed with the null -
hypothesis that mean values of 7-day UCS and UCS after 12 f-t cycles were equal.
Outcomes of the analysis are shown in Table 4.11 and resulted p-values were less than
0.05 which lead to reject the null hypothesis and the mean values, 7-day UCS being
higher than f-t UCS showed that the f-t cycles cause material deterioration.

The linear correlations between the DVs and other three tests did not produce
convincing outcomes as shown in Figures 4.38 to 4.40, particularly for CM, with the
obtained R2 0.24, 0.23 and 0.13, while the R2 for QM of 0.88, 0.76 and 0.41 were much
more promising. It can be recalled that the DVs and the UCS obtained after TST were

118
Chapter 4 F-T and Moisture Effects

also showed minimal linear regression. Therefore, using TST to evaluate durability of
lightly stabilised granular materials requires further study.

Table 4.8: The relationship between different durability tests

Material Variable 1 (Test 1) Variable 2 (Test 2) R2


VST f-t durability 0.84
VST 7-day 0.84
VST DV 0.24
CM
f-t durability 7-day 0.86
f-t durability DV 0.23
7-day DV 0.13
VS f-t durability 0.67
VS 7-day 0.93
VS DV 0.88
QM
f-t durability 7-day 0.76
f-t durability DV 0.41
7-day DV 0.76

Table 4.9: Paired t-test for sample means of VST and F-t durability test

CM QM
F-t durability F-t durability
VST VST
UCS UCS
Mean 2.09 1.62 1.81 1.51
Variance 0.18 0.15 0.07 0.02
t-stat 9.30 6.24
t critical 1.80 1.80
P <0.0001 <0.0001

119
Chapter 4 F-T and Moisture Effects

Table 4.10: Paired t-test for sample means of VST and 7-day UCS test

CM QM
VST 7-day UCS VST 7-day UCS
Mean 2.09 1.76 1.81 1.67
Variance 0.18 0.05 0.07 0.03
t-stat 4.67 4.56
t critical 1.80 1.80
P 0.0003 0.0004

Table 4.11: Paired t-test for sample means of 7-day UCS and f-t durability test

CM QM
F-t durability F-t durability
7-day UCS 7-day UCS
UCS UCS
Mean 1.76 1.62 1.67 1.51
Variance 0.05 0.15 0.03 0.02
t-stat 2.40 6.30
t critical 1.80 1.80
P 0.02 <0.0001

VS vs f-t durability
2.5 CM Y(f-tCM )=0.82X(VSCM)-0.0926
QM
R2=0.84
UCS_f-t durability test (MPa)

Regression line_CM
Regression line_QM
2

Y(f-tQM)=0.4735X(VSQM)+0.6558
R2=0.67
1.5

0.5

1.2 1.6 2 2.4 2.8


UCS_Vacuum satruation test (MPa)

Figure 4.35: Correlations between UCS obtained VST and f-t durability tests

120
Chapter 4 F-T and Moisture Effects

VS vs 7-day UCS
CM
2.5 QM Y(7-dayQM)=0.629X(VSQM)+0.5283
Regression line_CM R2=0.93
Regression line_QM

7 day_UCS (MPa)
2

Y(7-dayCM)=0.4686X(VSCM)+0.7802
R 2=0.84

1.5

0.5

1.2 1.6 2 2.4 2.8


UCS_Vacuum satruation test (MPa)

Figure 4.36: Correlations between UCS obtained from vacuum saturation and 7-
day UCS tests

20

VS vs DV
18 CM
QM
Regression line_CM
Regression line_QM
Dielectric value

16
Y(7-dayQM)=-7.6646X(DVQM)+30.70
R2 =0.88

14

12
Y(DVCM)=-0.4524X(VSCM)+11.71
R2=0.24

10

1.2 1.6 2 2.4 2.8


UCS_Vacuum satruation test (MPa)

Figure 4.37: Correlations between UCS and final DVs obtained from vacuum
saturation test and TST respectively

121
Chapter 4 F-T and Moisture Effects

3
7-day UCS vs f-t UCS
CM
QM
Regression line_CM
2.5 Regression line_QM

Y(f-tQM)=0.99X(7-dayQM)+0.17
R2=0.76
7 day_UCS (MPa)
2

Y(f-tCM)=0.53X(7-dayCM)+0.9
R2=0.86

1.5

0.5

0.8 1.2 1.6 2 2.4


F-T durability test_UCS(MPa)

Figure 4.38: Correlations between UCSs and final DVs obtained from vacuum
saturation test and TST respectively

20
F-T UCS vs DV
CM
QM
Regression line_CM
Regression line_QM

16
Dielectric value

Y(7-dayQM)= -9.11X(DVQM)+30.60
R2=0.41

12
Y(DVCM)= -0.49X(f-tCM )+11.56
R2=0.23

1 1.5 2 2.5 3
F-T durability test_UCS (MPa)

Figure 4.39: Correlations between UCSs and final DVs obtained from vacuum
saturation test and TST respectively

122
Chapter 4 F-T and Moisture Effects

20
7-day UCS vs DV
CM
QM
Regression line_CM
Regression line_QM

Dielectric value 16
Y(DVQM)= -10.88X(7-dayQM)+34.96
R2=0.76

12
Y(DVCM)= -0.65X(7-dayCM )+11.90
R2=0.13

1 1.5 2 2.5 3
7-day_UCS (MPa)

Figure 4.40: Correlations between UCS and final DVs obtained from vacuum
saturation test and TST respectively

4.3.6 Flexural and UCS varying with moisture content

The outcome discussed so far showed that the UCS and the DV showed dependency to
moulding moisture content. In this section, the UCS and flexural results obtained by
changing the post compaction moisture content (PCMC) is presented. Figure 6.16
depicts the relationship between the PCMC and MoR of beam samples cured for 28 and
14 days, with their MoR varying from 0.156MPa to 0.538Mpa and 0.125 MPa to 0.342
MPa respectively and their moisture contents from approximately 7.70% to 1.90%. The
highest PCMC was obtained by the samples tested immediately after being removed
from the curing room and those of the other moisture contents by placing the samples in
an oven for different time durations. The trend lines for both 14 and 28 days showed
similar behaviours, whereby the MoR increased with an increasing moisture content,
reached a peak and then started to decline. Moreover, the flexural modulus and UCS
indicated similar trends, as shown in Figures 6.17 and 6.18 respectively. Nazarian and
Yuan (2008) found similar results for the seismic moduli of a fine-grained material

123
Chapter 4 F-T and Moisture Effects

compacted at different moisture contents. These results reinforce the concept of drying a
pavement prior to surfacing it to obtain better performance.

The effective stress of an unsaturated soil can be determined using Eq. (6.3) which was
proposed by Bishop (1959). In it, Xw is one for saturated soils and zero for dry ones
whilst matric suction is high at a low moisture content and decreases with increases in
the moisture content. Although Xw and matric suction have opposite trends, there is a
moisture content at which their combined effect is maximum and provides maximum
strength and stiffness. This occurs on the dry side of the OMC and explains the
behaviour obtained from this study.

 '  (  ua )  X w (ua  uw ) Eq. (6.3)

where  ' = effective stress,  = total stress, X w = Bishop’s effective stress parameter,
and Ua and Uw = pore air and pore-water pressure respectively.

0.6 28 Days
14 Days
28 Days MOR Trend Line
14 Days MOR Trend Line
0.5
Modulus of Rupture (MPa)

0.4

0.3

0.2

0.1

0 2 4 6 8
Moisture Content (%)

Figure 4.41: MoR varying with moisture content

124
Chapter 4 F-T and Moisture Effects

28 Days Flexural Modulus


5000
14 Days Flexural Modulus
28 Days Trend Line
14 Days Trend Line
4000

Flexural Modulus (MPa)

3000

2000

1000

0
0 2 4 6 8
Moisture Content (%)

Figure 4.42: Flexural modulus varying with moisture content

2.4

14 Days UCS
14 Days UCS Trend Line

2
UCS (MPa)

1.6

1.2

0.8

0 2 4 6 8 10
Moisture Content (%)

Figure 4.43: UCS varying with moisture content

125
Chapter 4 F-T and Moisture Effects

4.4 Summary

Chemical stabilisation of granular materials is in practice for many years to obtain


enhanced mechanical properties, but environmental factors posing threats to materials’
performance. Moisture and durability studies are performed to assess the materials’
ability for resisting environmental factors and remain integrated in its life. A
standardised and commonly accepted laboratory protocol for assessing durability of
stabilised granular materials is not available and for lightly stabilised granular material,
very limited studies focussed on moisture and durability aspects. This chapter
investigated four testing procedures all of them performed on standard cylindrical
specimens for evaluating the f-t durability while TST, USC and flexural beam test
investigated the moisture susceptibility.

Moulding moisture content was selected as a primary variable for assessing durability
of lightly stabilised granular materials. F-t durability, VST and 7-day UCS tests showed
significant dependency to the moulding moisture content in which the UCS increased
with decreasing moisture content before showing decline at the lowest moisture content.
The flexural properties and the UCS measured after varying their PCMC also revealed
similar outcome.

A comparison with f-t durability test and 7-day UCS showed that f-t cycle is less
destructive for materials compacted with lessor moisture content. Higher UCS at lower
moisture content leaves a thought in pavement engineers in controlling the moisture in
the base layers, in this case lightly stabilised granular material. Allowing certain amount
of moisture to evaporate and increase the bearing capacity of the pavements before
layering asphalt is certainly agree with the findings of this study. However, pavement
undergoes changes in moisture content during its service, therefore the measures to limit
excessive moisture during its function is essential for better pavements’ performance.

This study also showed that, limiting the moisture content is good for the controlling
destructive nature of f-t cycles. Irrespective of moisture content, f-t cycles caused
deterioration, but the effect was minimal for lower moisture contents. Therefore, as
presented in this study, interpretation of f-t durability test results with moisture content
gives more insight to the materials behaviour.

126
Chapter 4 F-T and Moisture Effects

TST measures DVs of cylindrical specimens was also included in the studies because it
is an indirect test method depends on moisture content as well. The tested variables and
its outcomes showed that TST has a potential evaluate durability and moisture aspects
of lightly stabilised granular material. One of the variables tested in TST was moulding
moisture content and similar to other three tests, TST also showed that the DVs are
dependent on moulding moisture content and reducing it from the OMC reduced the
DVs and increased the UCS which was measured at the end of the test. It also showed
the amount of binder content has an effect on DVs with increasing binder content
showed reduction in DVs. The results comparison with two different binders, CF and
SL, provided some indication regarding the possible relationship between DVs and
strength.

TST identified that the coarser material showed lessor DV, therefore it’s less susceptible
to moisture and f-t effects. Moreover, TST also revealed that the quality of the porous
disk and smoothness of the top surface of specimens were important to obtain precise
and repeatable measurements. A scattered relationship between DVs and moisture
content well explained that the latter is not a mere measure of moisture content. The rate
of increase in DVs and the moisture content measured in top and bottom of the
specimen showed that the reading up to10 days is enough terminate the test. Two
different TST setups showed the importance capillary rise of water through bottom
surface however specimens tested using TST setup 2 showed higher DVs than those in
TST setup 1 with because of some additional moisture entered through 20 mm
peripheral surface. These outcomes suggest that, TST can be used to assess the moisture
susceptibility of the material.

Finally, the comparison with all four test results indicated that the 7-day UCS, f-t
durability test and VST showed a good correlation among them. Vacuum saturation and
7-day UCS tests well correlated with f-t durability test for both CM and QM. Since the
testing time for f-t durability is nearly a month, Vacuum saturation or 7-day UCS tests
can be used to assess the durability of the material. The established relationship between
these three tests can be used in circumstances where one value is known and the other
value needs to be found. The linear correlation with DVs and the other three tests did

127
Chapter 4 F-T and Moisture Effects

not show promising outcomes. Therefore using TST for assessing durability of the
lightly stabilised material requires further evidence.

128
5 Chapter 5

Assessing Durability of Lightly Stabilised Granular Material


against Freeze-Thaw and Wet-Dry Cycles Using Different
Element Testings and Testing Procedures

5.1 Introduction

A pavement’s ability to withstand traffic loading depends on the strength and stiffness
of its layers which are considerably affected by environmentally driven variables such
as temperature, moisture and freeze-thaw (f-t) cycles. The influence of temperature is
more prominent in asphalt layers whereas the f-t effect is significant in granular bases
and subgrades in cold regions while, in arid and semi-arid regions, moisture variations
in granular bases and subgrades are more critical.

Stabilisation is the process of adding granular material or binder for improving the
mechanical properties of base and subgrade materials. In the mix-design process of
cementitious stabilisation, durability studies are carried out to determine the stabilized
materials’ behaviour when subjected to environmental variables. A typical design of a
soil-cement mixture considers strength and durability as the key engineering properties
for evaluating its quality. Unconfined compressive strength (UCS) testing helps to
ensure that the design of a mix meets the minimum strength requirement. Durability,
defined as the ability of a material to retain its stability and integrity over years of
exposure to the destructive aspects of weathering, show long-term performance of a
stabilized material. The most difficult aspect of chemical stabilisation is making sure
that it is sufficiently permanent to last for a pavement’s life (Dempsey and Thompson
1968).

At present, there is no commonly accepted testing procedure available for assessing the
durability of lightly stabilised granular materials. Of the related durability standards
Chapter 5 F-t, w-d durability study

available for soil-cement mixtures, ASTM D560 (2003) and ASTM D559 (2007)
outline procedures for assessing the durability of granular materials stabilised with
cement subjected to repeated f-t and w-d cycles respectively. According to these testing
procedures, at the completion of the first f-t/w-d cycle, each specimen has to be given
two firm strokes of approximately 13.3N by a wire scratch brush covering the entire
cylindrical surface. This procedure has to be repeated for 12 f-t/w-d cycles, after which
the percentage loss in the soil-cement is measured and compared with acceptable limits.
However, as questions have been raised regarding the uniformity and consistency of the
brushing process, many studies have replaced it with UCS testing after 12 f-t/w-d cycles
(Gutherie et al. 2008; Shihata and Baghdadi 2001; Zhang and Tao 2008).

Moreover, Shihata and Baghdadi (2001) tested three different soils stabilised with
cement and, as they found that results from the UCS measured after f-t cycles and
standard UCS test, were well correlated with those from the standard ASTM D 560
weight loss test, they suggested that the standard UCS test is a better indicator of
durability because of its shorter duration. Also, many highway agencies specify
minimum UCS as the durability criteria due to its acceptability and simplicity in terms
of sample preparation and testing (Shihata and Baghdadi 2001). However, given that the
field loadings caused by moving vehicles on pavement layers are better simulated by
cyclic load resilient modulus (Mr) test, using UCS test is not appealing. The main
drawback of the Mr test is its complexity and the level of skills required to obtain
consistent and repeatable results. The literature survey revealed that a limited number of
the durability studies available focussed on the Mr. Zaman (1999), Khoury and Brooks
(2010), and Zaman (2007) carried out series of f-t and w-d durability assessments of
aggregates stabilised with different stabilisers.

The ASTM-D593 (2011) vacuum saturation test (VST) is used for assessing the
qualification of pozzolans has received attention as a potential test for assessing
durability against f-t due to its shorter testing time. Gutherie et al. (2008) has reported
that the results from this test correlated well with those from the ASTM-D 560 (2003)
weight loss test and recommended the former due to its shorter duration. In recent years,
the tube suction test (TST), which was primarily developed to identify poorly

130
Chapter 5 F-t, w-d durability study

performing unbound aggregates, has also been used to assess the f-t susceptibility of a
granular material (Barbu and Scullion 2005; Solanki et al. 2011; Yeo et al. 2012).

Most of the durability studies carried out to date focussed mainly on stabilised
materials, with only a few related to lightly stabilised materials and the understanding of
its durability against f-t and w-d cycles have not been studied well so far. Moreover,
earlier studies used primarily one element testing whereas, in this study, three different
element testings were conducted in order to assess the durability of lightly stabilised
granular materials against f-t and w-d cycles while also considering the factors causing
changes in their mechanical properties. While monotonic displacement controlled
loading was selected for the UCS and flexural tests, repeated load triaxial (RLT) testing
was performed using cyclic axial loading with constant confining pressure. RLT testing
was used to study the resilient and permanent deformation (PD) behaviours and residual
unconfined compressive strength (RUCS). Although the characterisation of lightly
stabilised granular materials based on their tensile properties was carried using two
different elements testing (Paul and Gnanendran 2012b; Piratheepan J. 2010), the
sensitivities of their flexural properties to f-t and w-d cycles were never explored, which
formed the basis for one of the objectives. Along with the different element testings,
variables such as the type of granular material, number of curing days, binder
percentage, curing method and type of f-t procedures were considered for obtaining
better laboratory test to evaluate the durability of lightly stabilised granular materials.

5.2 Experimental Program

The details experimental methodology of each testing program is discussed in the


following section.

5.2.1 Unconfined Compressive Strength Test

5.2.1.1 Introduction

The UCS test is an index test that has gained wider acceptance in geotechnical and
pavement engineering because of its easy implementation. The UCS is defined as the

131
Chapter 5 F-t, w-d durability study

maximum axial stress a specimen can withstand in the absence of a confining pressure
and is used extensively in pavement and geotechnical engineering. In geotechnical
engineering, UCS test is used to obtain the undrained shear strength parameter (qu) of
cohesive soils and, in pavement engineering, to characterise unbound and stabilised
materials, and to define the minimum strength needed to meet durability requirements.

A UCS test is also used to find the optimum stabiliser content or optimum mix which is
mainly a relative comparison of different trial mixes; for example, Sireesh Saride et al.
(2014) evaluated the effectiveness of the flyash stabilisation of a reclaimed asphalt
pavement (RAP) using both the UCS and Mr. A similar study was carried out by Ganne
(2009) in which different percentages of a RAP, granular material and cement were
used to ascertain the optimum combination. Moreover, for a mix design of stabilised
granular materials, laboratory tests are conducted to determine the amount of binder that
needs to be added to achieve the minimum strength requirement which is often specified
as a 7-day UCS (Austroads 2010a). Although the classification used to differentiate
between modified and cemented materials is based on the UCS, there is no definitive
UCS value (Austroads 2010a). However, a clear distinction is essential in the design
stage because the material characterisations and failure distresses of modified and
cemented materials are different.

Most of the applications of UCS testing mentioned above use failure stresses to define
limits, however, it is not common to measure the strain and determine the stress-strain
behaviour. This is partly because of the difficulties of reasonably accurately measuring
strains. Piratheepan J. (2010) developed an experimental setup with internal strain
measurements for measuring strains and characterising lightly stabilised granular
materials using different elastic moduli. Paul and Gnanendran (2012a) used the same
experimental setup to examine the non-linearity of a lightly stabilised granular material
and presented a method for the predictions of modulus and stress-strain behaviour based
on modified and extended Ramberg-Osgood expressions. Evaluating the modulus using
established models which relate it to the UCS has been considered an alternative in the
absence of strain measurements. The ACI model shown in equation 5.1 which relates
the modulus of elasticity to the compressive strength has generally been used to
estimate the elastic modulus of a stabilised material (ACI 1997). Lim and Zollinger

132
Chapter 5 F-t, w-d durability study

(2003) reported that this model overestimated the modulus and proposed a different
regression coefficient for a stabilised material with strength in the range of 1.4 to 13.8
MPa.

E (t )  33  w1.5  f c (t ) 0.5 Eq. (5.1)

E (t )  modulus of elasticity in psi at time t


where:
w  mixture density in lb
f c  compressive strength in psi at time t

In the absence of commonly accepted durability test methods, the UCS is set as the
minimum strength requirement for satisfying the durability requirements of stabilised
base materials. As current test methods for stabilised materials are time consuming and
subject to uncertainties, it is common to relate or compare the results from different
traditional durability tests with those from UCS test with the objective to replace
traditional durability tests with UCS test. Zhang and Tao (2008) demonstrated that
durability, tube suction test (TST) and w-d tests showed better predictability with 7-day
UCS of a subgrade material stabilised with different binder contents. Shihata and
Baghdadi (2001) also reported that the standard f-t durability test demonstrated strong
correlation with either the RUCS or UCS. As a result, many road agencies have defined
a minimum UCS for satisfying durability requirements as an alternative to durability
testing (Shihata and Baghdadi 2001).

The main criticism of UCS test is its inability to simulate pavement loadings as it is
unconfined and subjected to monotonic loading. Therefore, it often requires validation
from field performances or correlation with other mechanical properties such as Mr.
However, it is still attractive because the database of its results established over the
years enables different limits to be defined and new findings to be compared with
previous ones, and it is relatively easy to implement.

As limited studies are available regarding the durability of lightly stabilised granular
materials in terms of w-d and f-t cycles, the UCS was chosen as one of three different
elements testing for carrying out f-t and w-d durability tests. Apart from assessing the

133
Chapter 5 F-t, w-d durability study

materials’ susceptibility to f-t and w-d, the influences of factors such as the f-t
procedure used, curing age of the samples, and types of material and binder content
were also extensively evaluated. Details of the experimental procedure are provided in
the next section.

5.2.1.2 Sample Preparation and Testing

To determine the UCS of a lightly stabilised granular material subjected to f-t and w-d
cycles, cylindrical specimens of 105 mm in diameter and 115 mm high were prepared in
a standard compaction mould. For each specimen, approximately 2500 grams of oven-
dried granular material was proportioned based on the reconstituted particle size
distribution. Then, the required amount of CF binder was measured and dry-mixed with
the granular material for 2-3 minutes and the amount of water determined from the
optimum moisture content (OMC) added and mixed thoroughly to form a uniform mix.

A cylindrical split mould applied with a releasing oil to facilitate easy removal of
specimens was chosen to cast 115 mm high specimens. The stabilised mix was poured
in three equal layers, with each was given 25 blows of Standard Proctor compaction
effort, as outlined in AS 5101.4 (2008). Each compacted sample was kept in the mould
for 24 hours to gain strength and avoid disintegration. Then, it was removed, wrapped
with a polyethene bag with vent holes and placed in a controlled environment at a
temperature of 23 ± 2˚Celcius (C) and 95 ± 5% relative humidity for curing. At the end
of the designated curing period, the samples were taken from the curing room and
subjected to f-t and w-d cycles according to the procedures discussed in section 5.2.4.
After the specified numbers of f-t and w-d cycles, the specimens were subjected to
monotonic displacement-controlled loading at a rate of 0.5mm/min, to obtain the failure
stress and for determining the UCS.

134
Chapter 5 F-t, w-d durability study

5.2.2 Repeated Load Triaxial Test

5.2.2.1 Introduction

It’s well known that pavement materials exhibit non-linear elastic responses and
experience PD during the initial application of repetitive loadings. Then, they tend to
show elastic responses after a certain number of loadings if the applied load is lesser
than the strength of the material. A RLT test is used mainly to investigate the resilient
(stiffness characteristics) and PD (rutting characteristics) behaviours of base and
subgrade materials. Studies related to resilient characteristics are more common due to
the time-consuming nature of PD tests. The main advantage of using a RLT test is the
ability to stimulate different in-situ stress conditions. While a static confining pressure
is used to simulate the lateral confining pressure from the soil, a cyclic deviatoric pulse
is applied to simulate wheel loadings.

The Mr, which relates the recoverable strain to the deviatoric stress, is a common
mechanical property used to characterise unbound and lightly stabilised materials. A
single resilient or any other elastic modulus was used in pavement analysis in the
empirical design era (AASHTO 1993). However, introduction of mechanical-empirical
(M-E) design approach demanded better characterisations of pavement materials,
including incorporating stress and the moisture-dependent Mr. A RLT test which
measures matric suction and models the suction-dependent Mr is gaining attention with
the aim of achieving better material characterisation and improved pavement design
(Craciun and Lo 2010; Ohiduzzaman et al. 2012; Salour et al. 2014).

In a RLT resilient modulus test, different stress sequences are applied to a cylindrical
specimen by varying the confining and deviatoric stresses. Typically, for a particular
stress sequence, the confining pressure is kept constant while the cyclic deviatoric stress
of constant magnitude is applied for a particular number of cycles at a defined
frequency. Then, the axial stress and resilient deformation are measured to determine
the Mr.

RLT test is also used to investigate the PD (rutting) responses of unbound materials.
Rutting, a phenomenon observed in subgrade and base layers, causes the material to fail

135
Chapter 5 F-t, w-d durability study

after thousands of load applications and is the primary failure criteria for granular base
and subgrade materials. Studies focussing on PD behaviour are generally performed
using RLT testing.

The equipment for a typical RLT test of an unbound granular base material consists of a
cylindrical cell, which can accommodate 100 mm by 200 mm cylindrical specimens and
operates at greater than 500 kPa lateral hydraulic or air pressure, a cyclic loading ramp
to apply an axial loading pulse and an arrangement for measuring deformations.

Also, RLT tests have been used in conjunction with accelerated pavement model (APM)
tests in an attempt to relate both methods and explore the possibility of correlating their
results (Hazell et al. 2010; Werkmeister 2006; Wilson et al. 2014). An APM test is
relatively time consuming, requires excessive manpower to conduct a single test and its
instrumentation has also been proven to be challenging and requires expertise to obtain
reproducible results. However, it holds the edge of being closer to the in-situ stress
conditions than RLT or any other laboratory tests. A successful linkage between RLT
and APM testing will help to gain confidence in RLT test results, using which influence
of different variables can be tested much more easily and at a lower cost. Another
important advantage of the RLT test is that, as it is not destructive, multiple tests can be
conducted on a single specimen.

Apart from characterising materials, the RLT test has been used to assess the durability
of stabilised materials against f-t and w-d actions and it is preferred over other
traditional durability tests because of its ability to closely represent in-situ stress
conditions. (Khoury 2005; Khoury and Zaman 2007; N and Zaman 2007). In a RLT
durability test, specimens subjected to different f-t or w-d cycles are tested for Mr and
limited studies are available focussing on PD.

In this study, durability of lightly stabilised granular materials subjected to different f-t
and w-d cycles was studied using RLT testing. Mr tests with selected stress
combinations (as explained in section 5.3.1) were conducted on specimens stabilised
with 1.5% of CF binder. Moreover, the resistance to rutting while undergoing
environmental phenomena, such as f-t and w-d, was studied in a series of RLT
permanent deformation tests, at the end of which, specimens were sheared at zero

136
Chapter 5 F-t, w-d durability study

confining pressure to obtain the failure stress referred to as the residual unconfined
compressive strength (RUCS).

5.2.2.2 Sample Preparation and Curing

For RLT testing, cylindrical specimens 195 mm high and 100 mm in diameter were
compacted using a split mould adopting the following procedure.

1. Nearly 4000 g of the material was obtained for each specimen based on the
reconstituted particle size distribution. It was oven-dried for 24 hours at 105o C
before being kept for 3-4 hours at room temperature to cool down (Figure 5.1).

2. The required amount of CF binder was added to the material and then dry-mixed
for 2-3 minutes to form a uniform dry mix.

3. The required amount of water was measured based on the OMC and divided into
two parts which were added to the dry mix one after the other and mixed
thoroughly for 4-5 minutes. The resultant uniform mix is shown in Figure 5.1.

4. As a result of the nature of the particle size distribution, the settlement of larger
particles along the circumference of a specimen formed an irregular surface. The
problem with the irregular surface is that it may hamper on-specimen
transducers mounted on the surface of the specimen and the sharp edges of
larger particles may puncture the membrane during the application of high
confining pressures. Therefore, 300 g of finer material (<2.36mm) was used
around the periphery during compaction to form a smooth surface.

5. Also, this finer material together with the CF binder and water kept the weight
ratio similar to that of the parent material.

6. Figure 5.2 shows the split mould used to mould a cylindrical specimen before
and after the assembly of its individual components.

7. A thin plastic sheet was placed around the circumference of the mould. 50 g of
finer material was poured into the mould and compacted with a 25 mm steel rod

137
Chapter 5 F-t, w-d durability study

to obtain a smooth bottom surface. Then, the required amount of parent material
was added using a funnel and 40 g of the finer material placed around the
circumference with the use of straight edge before compaction.

8. To achieve the maximum dry density (MDD), the materials were compacted in 7
layers, with each given 25 standard Proctor blows.

9. An extension collar was used for the final layer to reach a height above the top
edge of the cylindrical surface and then the top surface smoothed to reach the
required height.

10. The specimens were kept in the mould for a day to gain strength and avoid
disintegration while being unmoulded. Then, they were removed and wrapped
with polyethene bags to prevent damage from water drops in the curing room,
with vent holes made in the polyethene bags to allow sufficient circulation of
moist air for better curing. The curing room was controlled at a temperature of
23 ± 2˚Celcius (C) and 95 ± 5% relative humidity.

11. At the end of the specified curing period, the specimens were taken out of the
curing room and subjected to f-t and w-d cycles, as detailed in section 5.2.4. A
typical cured cylindrical specimen used for the f-t and w-d testing is shown in
Figure 5.3

12. The specimens were then installed in the triaxial system, before the RLT test.

Figure 5.1: Dry and wet mixes of Queensland granular base material (QM)

138
Chapter 5 F-t, w-d durability study

Figure 5.2: Split mould used to mould 100 mm diameter by 195 mm height
cylindrical specimens

Figure 5.3: Triaxial cylindrical specimen used for the f-t and w-d testings

139
Chapter 5 F-t, w-d durability study

5.2.2.3 Installation of Specimen

The following steps were undertaken to install the specimens in the triaxial base.

1. The bottom platen was attached to the triaxial base and the top surface cleaned
of any debris.

2. The specimens subjected to f-t and w-d cycles were taken from the
environmental chamber and oven, allowed to cool for 30 minutes to 1 hour to
equilibrate with the environment and then placed on top of the bottom platen.

3. A top platen with a diameter marginally larger than that of a specimen diameter
was chosen and, after ensuring that it was clean, was placed on top of the
specimen.

4. A membrane coated with liquid rubber was then placed around the specimen
and top and bottom platens. Liquid rubber was applied to the inner surface of
the membrane because Lo et al. (1989) showed that it minimises the risk of the
membrane being punctured due to high cell pressures and irregular surfaces of
the specimens.

5. The grooves of the top and bottom platens and O-rings were applied with
vacuum grease. Four O-rings were placed on the bottom platen and two on the
top, with one from each end aligned with the groove for proper sealing.

6. The deformation measurement arrangement discussed in the next section was


installed and the LVDTs adjusted to obtain the required range of readings.

7. The triaxial chamber was assembled using a Perspex cylinder and held together
with the steel top and bottom platens by six stainless steel rods.

8. Then, the chamber was filled with de-ionised and de-aired water from an
overhead tank which was then used to apply a confining pressure. A schematic
diagram and photographic view of the triaxial assembly are shown in Figures
5.4 and 5.5 respectively.

140
Chapter 5 F-t, w-d durability study

Figure 5.4: Schematic diagram of triaxial testing system

141
Chapter 5 F-t, w-d durability study

Figure 5.5: Photographic view of triaxial testing system

5.2.2.4 Deformation Measurement

The resilient deformation of a lightly stabilised granular material needs precise


measurement because it is very small as a result of the material’s high stiffness and it is
even more critical when the loading frequency is as high as 10 Hz. The resilient strain is
calculated based on the length of the specimen being considered for the deformation
measurement which does not necessarily represent its height.

Resilient deformation can be measured either externally or internally. Many works


related to triaxial tests were based on external deformation measurements as it is easy to
set up the measuring devices (Austin 2009; Khalid 2008; Khoury and Zaman 2004; Kim
and Drabkin 1994; Yang et al. 2005). AASHTO 307-99, the Mr standard for base
materials, also accepts deformation measurements from outside the triaxial cell which is
generally the deformation measurement of the loading ramp. The problem with external

142
Chapter 5 F-t, w-d durability study

measurements is that it is not as accurate as internal deformation measurement.


Furthermore, Kim and Drabkin (1994) suggested that an external deformation
measurement may not be suitable for high-stiffness materials as it could underestimate
the Mr.

Another approach for measuring deformation is to mount LVDTs on the top platen and
monitor this platen’s movement relative to that of the base of the cell with the measured
deformation assumed to be the sample deformation of the whole (Nazzal and
Mohammad 2010). Although this method is better than using external LVDTs, its main
drawback is the bedding error that occurs due to imperfect contact between the test
specimen and top and bottom platens. To minimise the bedding error, conditioning
cycles can be applied, as suggested in NCHRP (2004) and AASHTO-T307-99 (2007).
However, Pezo et al. (1991) criticized that the high confining pressure and deviatoric
stress of conditioning cycles could change the stress history and may affect the Mr, and
proposed “grouted ends”, whereby the top and bottom platens are grouted with the
specimen to provide proper contact. The problem with this technique is that the axial
deformation from the top platen cannot be accurately measured. Nazarian et al. (1996)
reported that deformation measurements of the whole height of a specimen grouted at
both ends showed erroneous results compared with those measured in the middle third
of a specimen The grouted ends technique also eliminates the possibility of measuring
matric suction during testing.

Ping et al. (2003) carried out deformation measurements using external and internal
arrangements and compared their results. Their setup is shown in Figure 5.6 in which
position 3 represents external deformation measurements and positions 2 and 1
represent internal full- and half-length measurement respectively. Different positions
(with position 1 moved upwards and downwards closer to the end platens) on internal
half-length measurements were also considered and they indicated the effect of the top
platen. Significant errors in external deformation measurements were also reported. The
authors suggested the internal full-length measurement due to the difficulty in alignment
of the internal half-length LVDT clamps.

Ideally, deformations should have a reference point within the sample and be measured
away from the top and bottom platens, as shown in Figure 5.7. Ekblad (2006) used 1000

143
Chapter 5 F-t, w-d durability study

mm high by 500 mm diameter cylindrical samples with anchoring devices buried during
compaction to measure deformations in the middle 60 cm of a specimen. This
arrangement is more suitable for samples with greater heights and diameters since
smaller ones may be disturbed by the burial of anchoring devices. kamal et al. (1993)
and Karasahin et al. (1993) adopted a similar technique by embedding studs in a
specimen.

Figure 5.6: Internal and external LVDT arrangements (Ping et al. 2003)

144
Chapter 5 F-t, w-d durability study

500 mm

1000 mm

LVDTs

Figure 5.7: On-specimen deformation measurement (Ekblad 2006)

Therefore, in this RLT test, the use of on-specimen transducers was essential for
accurate strain measurements and the arrangement used in this experiment measured
deformations at 75% of the length of a sample from the bottom.

LVDT Installation

As previously mentioned, deformation measurements were taken at 150 mm from the


bottom surface between the triaxial base and LW lug pasted on a specimen, with two
miniature LVDTs ranging from -5 mm to +5 mm used to measure the axial deformation.
Figure 5.8 shows an LVDT and its components installed on a specimen. The following
steps were undertaken to install the LVDTs.

1. The diametrically opposite locations of the LW lugs, 150 mm from the bottom
surface, were marked on the membrane.

145
Chapter 5 F-t, w-d durability study

2. At these locations, two perspex LW lugs were glued to the membrane using a
contact adhesive.

3. To make the bonding firm and steady, an elastic strap was used to hold the LW
lugs as soon as they were pasted onto the membrane to help them remain intact
with the specimen which was important for measuring any movement.

4. Two threaded rods were fixed into the triaxial base to connect the LVDT
holders.

5. Then, both LVDT bodies were inserted into the LVDT holders and tightened
sufficiently to prevent any movement.

6. The cores of the LVDTs were inserted into the LVDT bodies and placed nicely
to rest on top of the two LW lugs.

7. Then, bolts attached to the threaded rods were used to adjust the LVDT holders
to set the desired initial readings, upon reaching which, the LVDTs were fixed
by locking the bolts.

5.2.2.5 Stress Control Dynamic Test

An important consideration in any RLT test is to control the deviatoric and confining
stresses to measure deformations and determine the resilient and PD behaviours. The
dynamic triaxial system used for this study is a product of MTS which has a loading
frame (external load) with a force capacity of 25 kN and displacement of ± 100 mm.
Although the Australian calibration service periodically calibrated the external load cell,
the axial load and deformation measurements were not used for calculations because of
their lack of accuracy. Axial deformations were measured using the arrangement
explained in section 5.2.2.4 and an internal load cell, having a capacity of 25 kN and
calibrated using a SHIMADZU AUTOGRAPH AG-X 100 universal testing machine,
was used inside the triaxial chamber to measure the axial load. A hydraulic cell pressure
chamber could operate up to 500 kPa and was equipped with a transducer to control the

146
Chapter 5 F-t, w-d durability study

confining pressure. As shown in Figure 5.9, the frame’s axial loads and confining
pressures were controlled by the program TestStarTM.

LVDT body
LVDT body
LW core

LW Lug

Figure 5.8: On-specimen LVDT arrangement

The TestWare-SX software shown in Figure 5.10 was used to set up the input file
containing the stress sequences, with all the information required to define a haversine
load added for each stress combination. Because of the resting period, a loop had to be
created for each stress stage to define the number of loading cycles. The signal outputs
from the internal load cell, LVDTs, frame load and cell pressure transducers were

147
Chapter 5 F-t, w-d durability study

acquired in the .dat format and then converted to .xlsx for further analysis. Figures 5.11
and 5.12 show typical outputs from an internal load cell and a LVDT after been
converted to stress and strain respectively.

Figure 5.9: Control program TestStarTM

148
Chapter 5 F-t, w-d durability study

Figure 5.10: TestWare-SX program used to input stress sequences


Deviator Stress (kPa)

Figure 5.11: Typical load-time history

149
Chapter 5 F-t, w-d durability study

0.8

0.6
Strain ( 10-3)

0.4

0.2

0 400 800 1200 1600 2000 2400


Time (milliseconds )

Figure 5.12: Typical deformation-time history

5.2.2.6 Repeated Load Triaxial Resilient Modulus Test

Mr of base materials is the main input parameter required for pavement analysis and
design. Lightly stabilised granular base material is characterised similar to unbound
granular base materials using Mr (Austroads 2010a). Mr is defined as the ratio between
deviatoric stress and resilient deformation which is determined from a RLT test. There
are different testing standards available for determining Mr of a base and subgrade
materials. As different testing and sampling procedures can result in different Mr, it is
necessary to harmonise the testing procedure for better comparison across the globe.

5.2.2.6.1 Resilient Modulus Testing Standards

Resilient modulus testing standards for pavement materials have emerged from different
testing agencies from time to time in an attempt to improve the characterization of these
materials. The standards published by AASHTO for determining the resilient moduli of

150
Chapter 5 F-t, w-d durability study

subgrade and unbound granular materials have been widely used for resilient modulus
testing.

AASHTO introduced its first testing framework for subgrade soils in AASHTO T274-
82, the “Standard Method of Testing for Resilient Modulus of Subgrade Soils”, with the
AASHTO pavement design standard “Guide for Design of Pavement Structures” the
first to include it in 1986. The AASHTO T274-82 standard was criticized for reasons
including the many hours required for testing and severe loading conditions (Ping et al.
2003). To remedy these drawbacks, the modified testing procedure AASHTO T292-91I
was published in 1991 based on work presented by the Strategic Highway Research
Program (SHRP) and, in 1994, incorporating some changes, AASHTO T294-92
(AASHTO 1994) was released. The latest available resilient modulus guideline is
AASHTO T307-99 (2003)

The NCHRP published a guide for MEPD in 2004. As part of the project, NCHRP 1-28
set out guidelines for obtaining the resilient moduli of aggregate-based and subgrade
materials. Then, its extension, NCHRP 1-28A, provided another protocol which
attempts to harmonise the resilient modulus testing procedure for unbound materials.
The newly published M-E standard in AASHTO (2008) recommends a resilient
modulus procedure based on that of NCHRP 1-28A for an unbound granular material
along with/and AASHTO 307.

5.2.2.6.2 Comparisons of different testing standards


There are noticeable differences among the standards which may affect the outcomes of
experiments, for example, the sequences for applying different stress combinations may
have stiffening or softening effects. In T292-91, the confining pressure decreases from
higher values while, in T294-92 and 307, it increases from lower to higher ones. Ping et
al. (2003) argued that the the T292-91I procedure is more representative than the T294-
92 test procedure. However, no significant difference existed between the T292-91I
procedure and T294-92 procedure. The results were almost the same; either of them can
reach a reasonably accurate result for resilient modulus measurement.

151
Chapter 5 F-t, w-d durability study

In a Mr test, the magnitudes of the stresses selected (confining and deviatoric) have
always been different for subgrade and base materials in order to take account of the
different in-situ stress levels they experience under wheel loads (AASHTO-T307-99
2007). The recommended stress levels for subgrade materials are less than those for
base ones as they are the bottom-most layers and experience less stress than the base
layers.

AASHTO-T307-99 (2007) is the commonly used Mr testing standard for base and
subgrade materials while the AASHTO (2008) M-E design standard recommends the
newly developed NCHRP (2004) as a suitable testing method as well. There are no
specific guidelines available that suggest appropriate stress levels and stress sequences
for conducting a Mr test on a lightly stabilised granular material. Therefore, based on the
NCHRP (2004) and AASHTO T307-99 (2003) Mr testing guidelines, tests were
conducted on both unbound and stabilised materials to ascertain a suitable testing
procedure for determining the Mr.

Specimens of QM stabilised with 1.5% of CF binder and cured for 56 days and unbound
granular material were selected to determine the Mr based on the NCHRP (2004) testing
procedure. As shown in Tables 5.1 and 5.2, NCHRP (2004) suggests 30 stress
combinations with cyclic axial stresses ranging from 10 kPa to 966 kPa and five
different confining pressures, 20.7, 41.4, 61.9, 103.5 and 138.0 kPa, each with six
different deviatoric stresses.

The outcomes of the initial tests suggested that controlling lower axial load values was
difficult. The frame load showed higher values than the internal load cell due to friction
between the loading ramp and cell’s top platen. Although the Mr was based on
measurements from the internal load cell, as the input stress values of the axial loads in
the control program were directly linked to the frame load instead, the frame load’s
input values were increased to accommodate the loss of load due to friction so that the
internal load could achieve the target axial load. However, at lower axial stresses, the
differences between the expected and measured axial loads were significant, as shown
in Figure 5.13 whereas, at higher axial stresses (<200 kPa), they reduced to a large
extent.

152
Chapter 5 F-t, w-d durability study

The lower axial stresses also caused inconsistent deformation measurements, as can be
seen in Figure 5.14. As previously mentioned, two LVDTs were used to measure the
axial deformations, with the Mr calculation based on the average of their readings. The
discrepancies between the LVDT readings were noted to be significant (Figure 5.14),
with lower deviatoric stresses because of the small resilient deformation measurement.
It is also important to note that differences in LVDT readings were less for the unbound
granular material compared with lightly stabilised granular material due to the former’s
lower stiffness and it deforming markedly which could have resulted in consistent
measurements. Therefore, the small stress values suggested in NCHRP (2004) are not
only difficult to control but may also cause inconsistent measurements, particularly for a
lightly stabilised granular material.

Furthermore, to investigate the reproducibility of the Mr test and the influence of the
stress values, another specimen stabilised with 1.5% of CF binder and cured for 56 days
was tested similar to the previous test. The percentage differences between the
measured deviatoric stresses and Mr of both tests were calculated and tabulated, as
shown in Table 5.3, and plotted against the average deviatoric stress, as shown in Figure
5.15. Significant differences were observed in the lower deviatoric stress for both the
stress values and Mr. Therefore, to obtain reliable outcomes, small deviatoric stresses
needed to be eliminated from the Mr test. Moreover, controlling the confining pressure
had never been an issue as it was kept constant (not cyclic stress) for a particular stress
sequence.

As previously mentioned, the AASHTO-T307-99 (2007) testing guidelines are


commonly used for a Mr tests. Therefore, similar to the previous test, a Mr test based on
AASHTO-T307-99 (2007) testing protocol was conducted on a specimen stabilised
with 1.5% of CF binder and cured for 56 days. The Mr, percentage difference between
LVDT readings and target and achieved deviatoric stress readings are shown in Table
5.4. Notable differences between the NCHRP and AASHTO testing procedures are the
application of the stress sequence (the order of combining the deviatoric stress and
confining pressure), number of stress sequences and magnitude of the deviatoric stress.
The maximum deviatoric stress suggested in NCHRP for base materials is
approximately four times that in AASHTO. As far as the lightly stabilised material is

153
Chapter 5 F-t, w-d durability study

concerned, the amount of stress suggested by AASHTO for a base material was not
sufficient to generate a significant amount of strain as small resilient deformations
caused erroneous results, as can be seen in Table 5.4. Therefore, AASHTO testing
procedure may not be suitable for a Mr test of a lightly stabilised material.

Based on the preliminary assessment, the number of stress values suggested in NCHRP
was reduced from 30 to 16 by eliminating the lower stress values, and the order of the
stress sequence suggested in AASHTO was selected. Moreover, using the higher stress
values suggested by NCHRP is more relevant because the introduction of heavy
vehicles resulted in testing upgrades. Table 5.5 shows the selected stress combinations
for the remaining Mr tests. The Mr for each cycle was calculated based on the deviatoric
stress and resilient strain, as shown in equation 5.2, with the final Mr for each stress
sequence obtained by averaging the last five of its load cycles of each stress sequence.

Mr  d / r Eq. (5.2)

where:

Mr  resilient modulus
 d  axial deviatoric stress
 r =axial recoverable strain

154
Chapter 5 F-t, w-d durability study

Table 5.1: Mr test of lightly stabilised granular material based on NCHRP testing protocol

Target seating Measured Measured Average resilient


Target cyclic Confining % Err
pressure cyclic stress seating pressure Bulk stress (kPa) deformation Mr (MPa)
stress (kPa) pressure (kPa) deformation
(kPa) (kPa) (kPa) (mm)
10.4 4.1 1.2 29.3 20.7 92.6 0.002 22.6 118
20.7 8.3 2.9 13.9 41.4 141.0 0.002 49.1 199
34.5 13.8 18.8 37.1 69 262.9 0.005 63.2 542
51.8 20.7 33.7 44.7 103.5 388.9 0.007 64.4 734
69.0 27.6 53.2 51.4 138 518.7 0.007 63.5 1095
20.7 4.1 2.9 29.0 20.7 94.0 0.002 40.0 208
41.4 8.3 12.9 34.5 41.4 171.6 0.005 62.4 374
69.0 13.8 56.9 39.4 69 303.2 0.017 60.6 508
103.5 20.7 90.4 46.0 103.5 446.8 0.019 62.4 722
138.0 27.6 132.0 49.5 138 595.5 0.021 51.9 948
41.4 4.1 10.5 33.7 20.7 106.3 0.006 70.5 286
82.8 8.3 68.9 34.4 41.4 227.5 0.031 57.4 329
138.0 13.8 124.1 41.5 69 372.6 0.037 59.0 506
207.0 20.7 202.8 47.4 103.5 560.7 0.037 55.3 827
276.0 27.6 280.3 52.5 138 746.7 0.041 46.1 1030
62.1 4.1 43.5 32.4 20.7 138.0 0.024 65.3 274
124.2 8.3 108.2 35.8 41.4 268.2 0.046 61.4 350
207 13.8 198.9 43.3 69 449.2 0.053 61.2 565
310.5 20.7 313.2 49.2 103.5 672.9 0.053 46.9 886
414.0 27.6 422.3 52.7 138 889.1 0.059 32.3 1065
103.5 4.1 87.0 33.3 20.7 182.4 0.039 63.4 337
207.0 8.3 195.1 38.0 41.4 357.3 0.066 57.7 444
345.0 13.8 336.3 45.5 69 588.8 0.072 47.8 696
517.5 20.7 521.8 47.8 103.5 880.1 0.086 24.2 914
690.0 27.6 711.9 48.2 138 1174.2 0.111 13.5 963

155
Chapter 5 F-t, w-d durability study

Target seating Measured Measured Average resilient


Target cyclic Confining % Err
pressure cyclic stress seating pressure Bulk stress (kPa) deformation Mr (MPa)
stress (kPa) pressure (kPa) deformation
(kPa) (kPa) (kPa) (mm)
144.9 4.1 144.3 32.7 20.7 239.1 0.056 40.7 384
289.8 8.3 292.7 36.1 41.4 453.0 0.096 33.2 459
483. 13.8 499.6 39.0 69 745.6 0.117 22.2 643
724.5 20.7 748.3 44.6 103.5 1103.4 0.134 10.8 835
966.0 27.6 1016.5 47.3 138 1477.8 0.172 5.6 886

Table 5.2: Mr test of unbound granular material based on NCHRP testing protocol

Target seating Measured Measured Average resilient


Target cyclic Confining % Err
pressure cyclic stress seating pressure Bulk stress (kPa) deformation Mr (MPa)
stress (kPa) pressure (kPa) deformation
(kPa) (kPa) (kPa) (mm)
10.4 4.1 0.4 49.8 20.7 112.3 0.0015 25.0 50
20.7 8.3 1.2 45.0 41.4 170.3 0.0022 41.5 82
34.5 13.8 8.5 38.7 69 254.1 0.0059 78.3 225
51.8 20.7 15.5 49.9 103.5 375.9 0.0079 58.6 300
69.0 27.6 23.8 60.1 138 497.8 0.0089 58.6 405
20.7 4.1 1.5 47.5 20.7 111.1 0.0017 61.7 160
41.4 8.3 8.2 43.1 41.4 175.6 0.0075 66.6 171
69.0 13.8 26.5 51.4 69 284.8 0.0162 45.0 246
103.5 20.7 70.2 52.8 103.5 433.4 0.0324 30.6 325
138.0 27.6 111.3 56.0 138 581.3 0.0416 24.0 402
41.4 4.1 7.2 44.8 20.7 114.0 0.0059 76.2 184
82.8 8.3 40.8 45.9 41.4 210.9 0.0284 45.6 216
138.0 13.8 105.0 47.8 69 359.8 0.0563 34.2 280
207.0 20.7 179.4 52.8 103.5 542.8 0.0713 25.3 377

156
Chapter 5 F-t, w-d durability study

Target seating Measured Measured Average resilient


Target cyclic Confining % Err
pressure cyclic stress seating pressure Bulk stress (kPa) deformation Mr (MPa)
stress (kPa) pressure (kPa) deformation
(kPa) (kPa) (kPa) (mm)
276.0 27.6 257.1 58.1 138 729.2 0.0861 20.9 448
62.1 4.1 11.4 47.0 20.7 120.5 0.0081 81.0 215
124.2 8.3 84.0 47.0 41.4 255.1 0.0518 41.5 243
207 13.8 175.0 50.0 69 432.0 0.0833 30.1 315
310.5 20.7 288.7 54.9 103.5 654.1 0.1095 23.9 395
414.0 27.6 397.4 58.9 138 870.3 0.1301 19.3 458
103.5 4.1 56.7 47.1 20.7 165.9 0.0374 51.5 227
207.0 8.3 170.3 49.4 41.4 343.8 0.0904 31.9 283
345.0 13.8 314.0 52.9 69 573.9 0.1320 20.5 357
517.5 20.7 497.8 55.7 103.5 864.0 0.1796 6.8 416
690.0 27.6 679.9 58.8 138 1152.7 0.2174 3.9 469
144.9 4.1 114.5 44.9 20.7 221.5 0.0767 29.3 224
289.8 8.3 263.9 47.1 41.4 435.2 0.1355 16.0 292
483. 13.8 467.1 49.3 69 723.4 0.1881 7.2 372
724.5 20.7 713.6 56.3 103.5 1080.4 0.2403 1.5 446
966.0 27.6 973.2 66.0 138 1453.1 0.2588 5.7 564

157
Chapter 5 F-t, w-d durability study

100

% Difference between measured and applied cyclic stress


80
Unbound
1.5% Stabilised

60

40

20

0 200 400 600 800 1000


Maximum Axial Stress (kPa)
Figure 5.13: Percentage differences between measured and applied cyclic stresses
of lightly stabilised material and unbound material tested based on NCHRP (2008)
Mr standard

100

80
% Difference between LVDTs readings

Unbound Material
60 Stabilised with 1.5% binder

40

20

0 200 400 600 800 1000 1200


Maximum Axial Stress (kPa)

Figure 5.14: Percentage differences between LVDT readings of lightly stabilised


material and unbound material tested based on NCHRP (2008) Mr standard

158
Chapter 5 F-t, w-d durability study

Table 5.3: Two different Mr tests of lightly stabilised material based on NCHRP
(2004)

Deviator Deviator
% Mr Test 1 Mr Test 2 %
stress test 1 stress test 2
difference (MPa) (MPa) difference
(kPa) (kPa)
1.2 1.7 29.4 118 193 38.7
2.9 3.5 17.1 199 282 29.3
18.8 25.1 25.1 542 658 17.6
33.7 41.6 19.0 734 870 15.7
53.2 62.6 15.0 1095 1043 4.8
2.9 3.9 25.6 208 379 45.1
12.9 17.7 27.1 374 508 26.3
56.9 61.1 6.9 508 642 20.9
90.4 97.7 7.5 722 738 2.2
132.0 136.2 3.1 948 865 8.8
10.5 16.3 35.6 286 385 25.8
68.9 71.9 4.2 329 500 34.2
124.1 130.8 5.1 506 590 14.2
202.8 206.1 1.6 827 754 8.9
280.3 281.4 0.4 1030 887 13.9
43.5 45.1 3.5 274 402 31.8
108.2 111.5 3.0 350 474 26.0
198.9 201.9 1.5 565 612 7.6
313.2 314.2 0.3 886 769 13.3
422.3 418.7 0.9 1065 941 11.7
87.0 87.9 1.0 337 407 17.2
195.1 199.7 2.3 444 481 7.7
336.3 343.2 2.0 696 625 10.3
521.8 522.4 0.1 914 860 5.9
711.9 696.4 2.2 963 869 9.7
144.3 130.1 9.8 384 409 6.2
292.7 275.9 5.7 459 478 3.9
499.6 479.1 4.1 643 620 3.6
748.3 730.0 2.4 835 807 3.3
1016.5 994.5 2.2 886 940 5.7

159
Chapter 5 F-t, w-d durability study

50

% Difference between two resilient modulus test results


40

30

20

10

0 200 400 600 800 1000


Maximum average deviator stress (kPa)

Figure 5.15: Stress dependency of Mr

160
Chapter 5 F-t, w-d durability study

Table 5.4: Mr tests of lightly stabilised material based on AASHTO testing


standard

Target Measured % % Average


Confining
deviator deviator difference difference resilient Mr
pressure
stress stress deviator LVDTs deformation (MPa)
(kPa)
(kPa) (kPa) stress readings (mm)
20.7 18.6 10.5 43.5 13.4 0.0042 392
20.7 37.3 36.2 2.9 33.4 0.0153 355
20.7 55.9 56.2 0.5 31.3 0.0253 334
34.5 31.0 31.2 0.6 18.5 0.0124 377
34.5 62.0 63.5 2.4 30.3 0.0271 351
34.5 93.1 93.5 0.4 29.6 0.0397 354
68.9 62.0 67.9 8.7 19.0 0.0225 452
68.9 124.1 129.7 4.3 26.6 0.0446 436
68.9 186.1 198.1 6.1 27.8 0.0681 436
103.4 62.0 71.7 13.5 18.6 0.0175 617
103.4 93.1 102.5 9.2 19.3 0.0289 533
103.4 186.1 203.5 8.6 24.0 0.0590 518
137.9 93.1 108.2 14.0 14.4 0.0247 658
137.9 124.1 143.2 13.3 19.1 0.0329 652
137.9 248.2 272.6 9.0 24.1 0.0649 630

161
Chapter 5 F-t, w-d durability study

Table 5.5: Selected stress combinations for Mr testing program

Confining Maximum cyclic Bulk


Number
Stress sequence pressure deviatoric stress stress ()
of cycles
(3) (kPa) (q) (kPa) (kPa)
0 (Conditioning
103.5 207.0 517.5 500
cycle)
1 41.4 91.1 215.3 100
2 41.4 132.5 256.7 100
3 41.4 215.3 339.5 100
4 41.4 298.1 422.3 100
5 69.0 151.8 358.8 100
6 69.0 220.8 427.8 100
7 69.0 358.8 565.8 100
8 69.0 496.8 703.8 100
9 103.5 227.7 538.2 100
10 103.5 331.2 641.7 100
11 103.5 538.2 848.7 100
12 103.5 745.2 1055.7 100
13 138.0 303.6 717.6 100
14 138.0 441.6 855.6 100
15 138.0 717.6 1131.6 100
16 138.0 993.6 1407.6 100

5.2.2.7 Repeated Load Triaxial Permanent Deformation Test

As a Mr test evaluates the short-term behaviour of a material but not its long-term
rutting behaviour, a PD test was carried out to study the rutting behaviours of specimens
subjected to f-t and w-d cycles. As, at the end of the Mr test, specimens deformed less
than 1% strain and did not fail, a PD test was carried out on the same specimens.
Repeated load PD test is expected to evaluate the impact of f-t cycles on rutting

162
Chapter 5 F-t, w-d durability study

behaviour and the comparison with other mechanical properties would be useful for
selecting the appropriate testing method.

A typical PD test procedure consists of applying a single level of stress to each


specimen. Usually, the number of repeated loading cycles greater than 100,000 enables
the study of long-term rutting behaviour. A main drawback of this procedure is that high
number of tests required in completing a series of test varying different stress values. As
a solution, it is common to conduct multiple stress stages on the same specimen with a
limited number of loading cycles (usually 10,000). This test is referred to as the multi-
stage RLT test and requires fewer numbers of specimens than a single-stage test.

Table 5.6 shows the four stress stages used for the multi-stage PD test, each of which
was applied for 10,000 load cycles. Confining pressures of 75, 100, 150 and 250 kPa
were selected to correspond to deviatoric stresses of 450, 600, 900 and 1500 kPa
respectively, and were kept constant for each deviatoric stress. All the tests were
conducted keeping q/p constant (= 2), as shown in Figure 5.16. Loading frequency of 10
Hz with 0.9s resting period, similar to the Mr test, was selected for this study as well.

At the end of the PD tests, all the specimens were tested under monotonic loading to
determine their failure loads at a zero confining pressure. Since all specimens
experienced almost identical stress histories and did not fail at the end of the PD tests,
relative comparisons were possible. The failure stress obtained is termed as ‘residual
unconfined compressive strength (RUCS)’.

Table 5.6: Stress combinations for PD testing

Cyclic
Confining Bulk
deviatoric Number of
Stress stage pressure stress (p) q/p
stress (q) cycles
(kPa) (kPa)
(kPa)
1 75 450 225 10,000 2
2 100 600 300 10,000 2
3 150 900 450 10,000 2
4 250 1500 750 10,000 2

163
Chapter 5 F-t, w-d durability study

Figure 5.16: Stress paths followed in PD testing

5.2.3 Flexural Test

5.2.3.1 Introduction

As there is a clear distinction between an unbound and cemented material in terms of


their distress mechanisms, their design approaches are different. A stabilised material
with a 28-day UCS greater than 1.5MPa is defined as a cemented material by Austroads
(2010a). Earlier characterisations of cemented materials were based on elastic
properties, such as the modulus of elasticity and Poisson’s ratio determined from
compressive loading tests (AASHTO 1993). However, characterisations based on
flexural properties are now favoured over compressive elastic properties since in-situ

164
Chapter 5 F-t, w-d durability study

wheel loadings can be better represented using a bending test as flexural strength is a
characteristic of a cemented material and tensile strength the governing factor in
pavement analysis.

The Austroads M-E design approach simplifies cemented layers as elastic, isotropic and
homogeneous which it uses in the multi-layer structural analysis program CIRCLY to
determine the tensile strain at the bottom of the layer, with the flexural modulus used as
the main input parameter. The obtained tensile strain is used in a tensile fatigue criterion
to determine the number of allowable equivalent standard axles (ESA).

Austroads (2010a) prefers a three-point flexural loading test to other laboratory tensile
tests, such as the direct tensile, IDT, longitudinal vibration and direct compression, for
measuring the flexural modulus while longitudinal vibration and direct compression
tests are not recommended for determining fatigue properties Austroads (2010a)

As, in a flexural loading test, a cemented material is considered a beam with loading on
its top, this test is a close simulation of actual wheel loading. However, the
unavailability of a standard and the requirement for special experimental arrangements
for conducting this test make it unattractive. To obtain approximate values, relationships
have been established between the UCS and other tensile properties, with one presented
by Austroads (2010a) given in equation 5.3. Austroads (2010a) also suggests
presumptive values for the flexural modulus which could be used directly from the table
in the absence of test results.

E Flex, 28  kUCS Eq. (5.3)

where:

E Flex, 28  the flexural modulus of field beams at 28 days of moist curing

UCS = the UCS of laboratory specimens at 28 days


k  the values of 1000 to 1250 used for GP cement materials depending on the
laboratory procedure.

165
Chapter 5 F-t, w-d durability study

Equation 5.4 shows the relationship developed between the flexural strength and
flexural modulus for a binder content of between 0.5% and 3% CF binder (Paul and
Gnanendran 2012b).

E Flex  4840 * MoR (R2 = 0.96) Eq.


(5.4)

where:

MoR = the flexural strength

EFlex = the flexural modulus

Fatigue failure, which is defined as the reduction in stiffness by half caused by


repetitive loadings, is the primary failure distress mechanism for a cemented material.
As the peak load in a fatigue test is selected so that the specimen remains in the elastic
range, it is generally limited to less than 60-90% of the flexural strength (Austroads
2010c). Therefore, a fatigue test generally follows a static breakage test which
determines the flexural strength. Characterising fatigue and determining the fatigue life
of a cemented material has attracted more attention over the past two decades (Arnold
and Morkel 2012; Austroads 2010c; Yeo et al. 2011). The outcomes of such studies
produced numerous equations relating the fatigue life of a cemented layer to either the
applied stress or strain of the material. The fatigue criteria defined by Austroads (2010a)
is:

0.804
113000 / Eini  191
N  RF[ ] Eq. (5.5)


Where:

N = allowable number of repetitions of the load

µe=tensile strain produced by the load (micro strain)

Eini = initial flexural modulus (MPa)

166
Chapter 5 F-t, w-d durability study

RF = reliability factor
As can be seen from the equation 5.5, fatigue life is dependent on the initial modulus of
the material. However, previous studies showed that fatigue life is dependent more on
braking strain than flexural modulus (González et al. 2013; Jameson et al. 1992;
Litwinowicz 1986). Austroads (2010c) suggested that the ratio between the applied
initial strain and breaking strain at 95% of the breaking load as a suitable indicator than
flexural modulus. Moreover equation 5.5, considers only pre-cracking behaviour which
is likely to lead towards a conservative design. Apart from this, a reliable test method
relating in-situ modulus to laboratory modulus is essential for better pavement life
prediction. Therefore, more fatigue studies needed considering all the factors discussed
above to define a reliable fatigue criteria.

Regarding a lightly stabilised granular material, as characterising it using the flexural


modulus was thought to be difficult due to its low tensile strength, IDT test was chosen
to characterise its tensile properties (Piratheepan et al. 2009). However, Paul and
Gnanendran (2012b) showed the possibility of characterising a lightly stabilised
material using a flexural beam test under monotonic loading, which measured the
flexural strength, and cyclic loading, which studied the fatigue characteristics of the
material.

There are only a few studies that have investigated environmental effects on flexural
properties. A study presented by Khoury and Zaman (2006) on the flexural properties of
granular materials stabilised with 10% class C fly ash subjected to f-t cycles showed
that these cycles had significant effects. As the durability of a lightly stabilised material
against f-t and w-d cycles is not a well-known concept in pavement engineering, in this
study, a series of beam specimens were prepared to study the f-t and w-d effects on the
flexural properties of specimens stabilised with 1.5% CF binders. The sample
preparation and test setup are discussed in the next sections.

167
Chapter 5 F-t, w-d durability study

5.2.3.2 Sample Preparation

The processes for the making and curing of soil-cement compression and flexural test
specimens in the laboratory recommended in ASTM D 1632 (2007) were followed for
the sample preparation. A beam specimen 76 mm high, 76 mm wide and 285 mm long
was prepared using a steel mould. The steel mould and compacted specimen are shown
in Figures 5.17 and 5.18 respectively. Initially, an oven-dried granular material was
proportionally divided based on the reconstituted particle size distribution and then dry-
mixed with the binder before being thoroughly mixed with water. The beam mould was
set up with its top platen open and moulding oil applied for the easy removal of
specimens. The mixed sample was compacted in the beam mould in three equal layers,
each of which was manually compacted with 90 blows of a square-ended tamping rod.
After levelling the mix, the top plate was placed on top of it and fitted with it. Final
compaction was achieved by applying a static load on the top platen using a spherical
seat to bring the sample to the desired height of 76 mm. Each compacted sample was
kept in the mould for 24 hours to gain strength and avoid disintegrating when
unmoulded. It was then removed and its weight measured to determine its dry density
with all specimens showing relative dry densities of 99%-100%. They were then
wrapped in polythene bags with breathing holes and placed in a controlled environment
at a temperature of 23 ± 2°C and 95 ± 5% relative humidity for curing.

168
Chapter 5 F-t, w-d durability study

Figure 5.17: Photographic view of steel mould, 76 mm by 76 mm by 284 mm used


for preparing flexural beam specimens

Figure 5.18: Compacted 76 mm by 76 mm by 284 mm flexural beam specimen

5.2.3.3 Experimental Setup

The material’s flexural properties, such as its modulus of rupture (MoR) and flexural
modulus, were obtained from a three-point loading test. The experimental setup
included measuring the applied load and net-span deflection based on the ASTM-C1609
(2010) “Standard Test Method for Flexural Performance of Fiber-Reinforced Concrete”.
A schematic diagram of the experimental setup is shown in Figure 5.19 and a
photographic view in Figure 5.20.

The net deflection at the mid-span was measured using two LVDTs, one mounted on
each of the two sides of a C-shaped bracket fitted with the beam aligned on its neutral
axis and the LVDTs fixed at its centre. The LVDT tips were pointed towards a steel
block which was placed on top of the surface of the beam. As Paul and Gnanendran
(2012b) found that the loading rate had a profound effect on the strength and stiffness of
a lightly stabilised material, the recommended rate of 0.5 mm/min was applied using a

169
Chapter 5 F-t, w-d durability study

Moogs testing machine. The load cell reading and LVDT measurements were obtained
during the test using a data acquisition program.

The MoR defined by equation 5.6 is the maximum tensile stress produced at the
outermost bottom fibre of a specimen subjected to flexural loading.

Pmax L
MoR  Eq.
bh 2
(5.6)where:

MoR = the flexural strength in MPa

L = the span length in mm

b = the average width of the specimens in mm

h = the average depth of the specimens in mm

Pmax = the maximum applied load in N

The flexural modulus at the peak load was calculated using,

23PL3  216h 2 (1  v) 
E Flex  1   Eq. (5.7)
108bh 3  115 L2 

where EFlex is the flexural modulus in MPa; P the load in N corresponding to the
deflection (δ) in mm; and ν the Poisson’s ratio.

170
Chapter 5 F-t, w-d durability study

Figure 5.19: Experimental setup for flexural testing

Figure 5.20: Photographic view of flexural testing arrangement (Paul et al. 2015)

5.2.4 Freeze-Thaw and Wet-Dry Cycles

At the end of the specified number of curing days, the UCS, triaxial and beam
specimens were subjected to f-t and w-d cycles, each of which took approximately two
days to complete. The selection of f-t and w-d procedures for the laboratory experiments
should closely simulate in-situ conditions. However, given the differences of scale
(element testing versus actual field conditions) and practical difficulties, compromises
are being made in methodologies for laboratory experiments. It is also important that

171
Chapter 5 F-t, w-d durability study

the selected procedures are within a certain timeframe to make them suitable for
industry applications. The selected f-t and w-d procedures for this series of experiments
are discussed in the following sections.

5.2.4.1 F-T Cycles

There is no standard f-t testing methodology available for lightly stabilised granular
materials. Of the relevant standards, ASTM-D560 (2003) recommends procedures for
assessing the durability of granular materials stabilised with cement. Most previous
studies selected the environmental parameters for freeze and thaw phases based on
ASTM D 560 and considered a freezing phase of −23°C for 24 hours and thawing phase
of 21°C for 23/24 hours as one f-t cycle (Khoury and Zaman 2007; N and R 2010).
However, to make the f-t cycle closer to in-situ conditions, supplying water through
moisture pads is recommended by ASTM-D560 (2003). According to this standard, a
water-saturated pad 6 mm thick and continuously supplied with water can be placed
underneath specimens to enable them to absorb water through capillary forces during
the thaw cycle. A similar procedure was used by Shihata and Baghdadi (2001) whereby
moisture pads were kept underneath the specimens during both f-t cycles and ensured
that they remained moist during the thaw cycle. Similarly, Khoury and Zaman (2006)
used a water chamber to supply water during the thaw cycle. However, the majority of
f-t durability tests on stabilised materials did not use moisture pads or any form of water
supply during either the freeze or thaw cycle (Gutherie et al. 2008; Khoury and Zaman
2007; Miller and Zaman 2000; N and R 2010)

In this study, to simulate f-t cycles, an environmental chamber as shown in Figure 5.21
having automated capability to operate both freezing and thawing temperatures was
chosen, with two different f-t procedures used for testing. In f-t procedure 1, one f-t
cycle comprised a freezing phase of -23°C for 24 hours and thawing phase of 21°C for
23 hours at a relative humidity of 95%. In f-t procedure 2, similar environmental
parameters were selected except that, in the thawing phase, moisture pads were placed
underneath the specimens to supply moisture, as shown in Figure 5.22. At the end of the
designated numbers of f-t cycles, the specimens were taken from the environmental
chamber for testing.

172
Chapter 5 F-t, w-d durability study

5.2.4.2 W-D Cycles

There is no standard w-d testing procedure available for lightly stabilised granular
materials. According to ASTM-D559 (2007), which is for compacted soil-cement, one
w-d cycle comprises submerging a specimen in water for 5 hours at room temperature
before placing it in an oven at a 71.8°C temperature for 42 hours. Parsons and Milburn
(2002) tested the durability of 7 different types of silt and clay against w-d based on
ASTM D 599 to assess the effectiveness of different stabilisers. Zhang and Tao (2008)
followed the same testing procedures to evaluate the effects of w-d cycles on silt clay
stabilised with cement at different moisture contents.

Figure 5.21: Environmental chamber used for f-t cycles

173
Chapter 5 F-t, w-d durability study

Figure 5.22: Triaxial specimen placed on top of moisture pad

In a series of studies, Zaman. et al. (1999), Khoury and Zaman (2002) and N and Zaman
(2007) tested the durability of aggregates stabilised with cement kiln dust, fly ash and
fluidised bed ash respectively against w-d cycles using similar environmental
parameters to those in ASTM D 559. However, they reversed the process of wetting
followed by drying in ASTM D 559, that is, one w-d cycle consisted of placing a
specimen in an oven at a temperature of 71.8oC for 24 hours and then submerging it in
potable water for 24 hours at room temperature. As an extensive literature study did not
reveal any other w-d testing procedures, the procedure suggested in ASTM D 559 was
closely followed in this series of tests. Each test was continued for 12 w-d cycles or
until the sample failed, with the latter found to be common. Figure 5.23 shows
specimens being submerged in water for 5 hours in every w-d cycle. After specific
numbers of w-d cycles, the specimens were subjected to UCS, flexural and RLT testing,
as explained in sections 5.2.1, 5.2.2 and 5.2.3 respectively.

174
Chapter 5 F-t, w-d durability study

Samples cured under normal curing conditions were also prepared and tested at different
curing ages which matched the total numbers of days spent by the f-t and w-d samples
in different cycles

Figure 5.23: Testing set-up for w-d process

175
Chapter 5 F-t, w-d durability study

5.3 Results and Discussion

5.3.1 Effect of Stress and Loading Frequency on Resilient Modulus

It is important to select suitable loading parameters when simulating actual wheel


loadings in a triaxial system or accelerated pavement model testing. They include the
shape, frequency and magnitude of the vertical stress pulse. Studies of the in-situ stress
distribution under a wheel loading have suggested that an actual loading wave form in
the field takes the form of either haversine or triangular loading (Huang 2004). Loulizi
et al. (2002) measured the stresses of flexible pavements having different designs and
vehicles travelling at different speeds, also found that the compressive pulse followed
either a haversine or bell-shaped equation. As AASHTO-T307-99 (2007), the laboratory
Mr testing standard for unbound granular materials, also suggests using a haversine load
pulse, all the RLT tests in this study were carried out using a haversine load pulse.

Regarding the loading frequency, depending on the speed of the vehicle and depth of
the pavement, the duration of the loading pulse varies. Barksdale (1971) used a finite
element model to develop a chart for calculating the loading duration as a function of
the speed of a vehicle and depth of a pavement layer. He also reported that the top
surface of a pavement experiences both sinusoidal and triangular wave forms at greater
depths. Loulizi et al. (2002) measured the stress pulses of vehicles moving at different
speeds at different depths using pressure cells and reported that the duration of loading
depended on the depth and vehicle speed. Vehicles moving at 70 km/h produced a 0.02
second load pulse at a depth of 40 mm and a 1 second duration was recorded at 597 mm
for a vehicle speed of 10 km/h. The current RLT test procedures in AASHTO-T307-99
(2007) and NCHRP (2004) recommend a haversine load pulse with a load duration of
0.1 second and cycle duration of 1 second to simulate a typical highway loading (a
resting period of 0.9 second). However, it is important to study the dependency of the
Mr of a lightly stabilised granular material on the loading frequency, which is not a
constant in field conditions, before selecting a particular loading frequency for
durability tests.

176
Chapter 5 F-t, w-d durability study

Moreover, there are ample studies available which show that the Mr of an unbound
granular material depends greatly on the material’s stress levels (Hicks 1970; Kolisoja
1997; Morgan 1966; Sweere 1990b). Smith and Nair (1973) noted a 50% increase in the
Mr for a bulk stress increase from 70 kPa to 140 kPa, and a 500% increase was shown to
result from an increase in the confining pressure from 20 kPa to 200 kPa (Monismith et
al. 1967). Moreover, the confining stress and sum of the principal stresses (bulk stress)
have been seen to be much greater influences than the deviatoric stress (Hicks 1970;
Monismith et al. 1967; Morgan 1966; Sweere 1990b). However, there is limited
knowledge available regarding the stress dependency of a lightly stabilised granular
material. Therefore, three RLT tests with three different loading frequencies were
conducted to assess the effects of loading frequencies and stresses on the Mr.

A specimen of Canberra material (CM) stabilised with 1.5% of CF binder and cured for
56 days was selected for RLT resilient modulus testing using the sequence established
in section 5.2.2.6. Three loading frequencies, 10 Hz, 5Hz and 1Hz, with a haversine-
shaped loading were selected and, as shown in Figure 5.24, a resting period of 0.9
second was kept constant irrespective of the loading frequency.

Table 5.7 shows the Mr obtained for the sixteen stress sequences along with their
measured stresses, confining pressures, and maximum axial and bulk stresses. As shown
in Figure 5.25, the k-θ (Mr versus bulk stress) model was used to compare the Mr
obtained for different frequencies. Irrespective of the loading frequencies, a good
correlation was found between the Mr and bulk stress with the regression coefficient
(R2) greater than 0.95. The regression equation was then used to determine the Mr for
different stress sequences and typical calculated variations in Mr for different numbers
of w-d cycles shown in Table 5.8 for stress sequences 4, 8, 12 and 16 (this is to make
comparison with specific bulk stresses which differ by small margins between the tests
largely because of variation in the deviatoric stresses). It can be seen in Table 5.8 that
the highest Mr was recorded for the lowest frequency. However, the difference in the Mr
between 5 Hz and 10 Hz was higher than that between 1 Hz and 5 Hz; for example, for
stress sequence 16, there was only a 10 MPa difference between 10 Hz and 5 Hz but
144 MPa between 5 Hz and 1 Hz. Similar observations can be made for the other stress
sequences as well. Therefore, it can be concluded that the loading frequency has a
minimal effect on the Mr of the lightly stabilised granular material.

177
Chapter 5 F-t, w-d durability study

1200 1200

800 800
Deviator Stress (kPa)

Deviator Stress (kPa)


400 400

0 0

0 150 300 450 600 750 900 1050 1200 1350 1500 0 150 300 450 600 750 900 1050 1200 1350 1500 1650
Time (milliseconds ) Time (milliseconds )

(a) (b)

1200

800
Deviator Stress (kPa)

400

0 400 800 1200 1600 2000 2400 2800 3200 3600


Time (milliseconds )

(c)

Figure 5.24 Load frequencies (a) 10 Hz (b) 5 Hz (c) 1 Hz

178
Chapter 5 F-t, w-d durability study

According to Fredrick Lekarp (2000), the Mr of unbound granular materials are also less
affected by the loading duration. However, Kim and Tutumluer (2006) reported that the
loading duration had a considerable effect on the PD behaviour of unbound granular
materials. Two different axial loading rates of 0.1 and 0.5 second were selected for the
study, axial pressures of 63 and 336 kPa applied with a constant confining pressure of
21 kPa and each load combination subjected to 72,000 load cycles. It was reported that
the 0.5 second loading accumulated 40% more PD than load duration for 0.1 second. It
is important to note that the Mr may not be a good indicator of PD behaviour and the
loading frequency may influence the rutting behaviour of a lightly stabilised granular
material. Therefore, further PD studies are needed to determine the effect of the loading
frequency on a lightly stabilised granular material.

Table 5.7: Mr versus loading frequencies

Maximum Maximum Maximum


Conf. Bulk Bulk Bulk
axial Mr axial Mr axial Mr
pressure stress Stress Stress
Stress (MPa) Stress (MPa) Stress (MPa)
(kPa) (kPa) (kPa) (kPa)
(kPa) (kPa) (kPa)
10 Hz 5 Hz 1 Hz

41.4 110 235 442 118 242 512 122 246 577

41.4 150 275 444 162 286 521 168 292 570

41.4 237 361 473 258 383 571 267 391 613

41.4 322 447 540 352 476 623 364 489 659

69.0 170 377 524 180 387 638 188 395 670

69.0 240 447 548 258 465 665 269 476 692

69.0 381 588 645 416 623 763 430 637 805

69.0 531 738 753 582 789 876 596 803 921

103.5 248 558 679 264 575 810 273 583 828

103.5 357 668 750 384 695 874 398 708 906

103.5 570 881 899 622 932 1043 636 946 1092

103.5* 569 879 896 620 931 1040 633 944 1104

138.0 329 743 889 351 765 992 359 773 1049

138.0 467 881 963 505 919 1105 518 932 1168

138.0 752 1166 1160 823 1237 1336 840 1254 1417

138.0 1059 1473 1388 1162 1576 1538 1193 1607 1606

179
Chapter 5 F-t, w-d durability study

2000

Frequency
10 Hz 0.5995 2
1600 5 Hz Mr=18.234() (R =0.95)
1 Hz 0.5944 2
Mr =18.108() (R =0.97)
10 Hz 0.6638 2
5 Hz Mr=10.184() (R =0.96)
Resilient Modulus (MPa)

1 Hz
1200

800

400

0
200 300 400 500 600 700 800 900
1000
Bulk Stress (kPa)

Figure 5.25: Variations in Mr with different frequencies and bulk stresses

Table 5.8: Calculated Mr based on regression equation

Mr (MPa) Mr (MPa)
Mr (MPa) Mr (MPa)
Load Load
Loading Regression Load Load
Sequence 12 Sequence 16
Frequency Parameters Sequence 4 Sequence 8
θ =1055.7 θ =1407.6
θ =422.3 kPa θ =703.8 kPa
kPa kPa
k1=18.234
1 Hz k2=0.5995 684 929 1184 1407
R2=0.95
k1=18.108
5 Hz k2=0.5944 679 922 1175 1397
R2=0.97
k1=10.184
10 Hz k2=0.6638 563 791 1035 1253
R2=0.96

180
Chapter 5 F-t, w-d durability study

As illustrated earlier in Figure 5.25, Mr is very well related with bulk stresses. The
results were further analysed to study the stress dependency of the Mr of a lightly
stabilised granular material. As shown in Figure 5.26, for a particular confining
pressure, an increase in the deviatoric stress resulted in an almost linear increase in the
Mr. However, for lower deviatoric stresses and confining pressures, such dependency
was not noticed and, for a particular deviatoric stress, an increase in the confining
pressure resulted in an increase in the Mr. The combined effects of deviatoric stresses
and confining pressures are best described using the bulk stress, as previously
explained, with the k-θ model able to be used to describe the stress dependency of a
lightly stabilised material.

181
Chapter 5 F-t, w-d durability study

10 Hz 1600 5 Hz
1600
20.7 kPa 20.7 kPa
41.4 kPa 41.4 kPa
69.0 kPa 69.0 kPa
103.5 kPa 103.5 kPa
138.0 kPa 138.0 kPa
1200

Resilient Modulus (MPa)


Resilient Modulus (MPa)

1200

800

800

400

0 400

0 250 500 750 1000 0 250 500 750 1000 1250


Deviator Stress (kPa) Deviator Stress (kPa)

(a) (b)

2000 1 Hz
20.7 kPa
41.4 kPa
69.0 kPa
103.5 kPa
138.0 kPa
1600
Resilient Modulus (MPa)

1200

800

400

0 250 500 750 1000 1250


Deviator Stress (kPa)

(c)

Figure 5.26: Variations in Mr with different deviatoric stresses and confining


pressures (a) 10 Hz (b) 5 Hz (c) 1 Hz

182
Chapter 5 F-t, w-d durability study

5.3.2 F-T Durability Test

In cold regions, f-t is a common environmental threat to stabilised materials as their


mechanical properties are considerably affected by temperature and seasonal reversals.
Assessing durability against f-t is carried out in laboratory tests as part of a mix design
to ensure that the material has the capability to resist f-t and remains integrated for the
entire life of a pavement. However, no established standard for assessing the durability
of a stabilised granular material against f-t cycles is currently available. Of the relevant
standards, ASTM-D560 (2003) suggests a testing procedure for a soil-cement mix based
on weight loss after 12 f-t cycles. The weight losses measured in laboratory experiments
are compared with the allowable limit which was established based on field studies.

The ASTM-D560 (2003) method for assessing the durability of stabilised materials
against f-t has many drawbacks. The absence of a proper testing methodology for
stabilised materials has opened avenues for several different testing methodologies.
Literature studies revealed that there are limited studies available that focussed on the f-
t effects on a lightly stabilised material. Therefore, the main objective of this series of
experiments is to assess the durability of a lightly stabilised material against f-t cycles.
As previously mentioned, three different element tests were conducted to assess the
effects of f-t on different mechanical properties. UCS testing was carried out under
monotonic loading and RLT testing under dynamic loading. RLT test studied the
resilient and PD characteristics of the materials subjected to f-t cycles. Moreover, the
MoR and SSM were also determined from flexural testing, as outlined in section 5.2.3.
The outcomes from these three different element tests are expected to provide a
complete picture of the effects of f-t cycles. In addition, the study was extended to
evaluate the impacts of different variables, including the number of curing days, f-t
procedure, curing method and binder content used, on the selected mechanical
properties to determine a better f-t testing procedure for assessing the durability of a
lightly stabilised material against f-t cycles. In the following section, the experimental
results obtained from each test are discussed in detail.

183
Chapter 5 F-t, w-d durability study

5.3.2.1 F-T Durability Study using Unconfined Compression Strength Test

F-t studies of lightly stabilised granular base materials were initiated with a series of
UCS tests. The UCS of cylindrical specimens 105 mm in diameter and 115 mm height
were determined from displacement controlled monotonic testing described in section
5.2.1. To assess the effects of f-t cycles on the UCS and determine the influence of
different variables and procedures on UCS, testing program shown in Table 5.9 was
carried out and depending on the variables, the testing program was divided into four
different series. A comparison between these different series is expected to suggest the
most appropriate testing method for assessing the durability of lightly stabilised
granular materials against f-t cycles. Two different granular materials, CM and QM,
stabilised with CF binders and two different f-t procedures were selected. The influence
of variables such as the curing age, binder content and method of curing were also
studied using triplicate samples moulded for each f-t cycle.

The first series of QM specimens stabilised with 1.5% CF were cured for 7 days and
subjected to f-t procedure 1. Table 5.10 shows the UCS measured after different
numbers of f-t cycles which can also be referred to as the RUCS as they were obtained
after being subjected to f-t cycles. The UCS at 0 f-t cycle represents specimens tested
after 7 days of curing without being subjected to any f-t cycles. As shown in Figure
5.27, the UCS increases with an increasing number of f-t cycles up to 12 before
showing a decline after 24 f-t cycles. Compared to 0 f-t cycle, the strength gain after 12
f-t cycles is approximately 50%. Although significant reductions in strength were
observed in the next 12 f-t cycles, after 24, the UCS remained almost equal to that for 0
f-t cycles. Another set of specimens stabilised with 3% CF also showed a similar trend
to that of specimens stabilised with 1.5% CF as shown in Figure 5.28. After 12 f-t
cycles, a strength gain of 33% 0 f-t cycles was obtained. Although the next 12 f-t cycles
showed deleterious f-t effects, the UCS after 24 f-t cycles remained marginally higher
than UCS of 0 f-t specimens. The other granular material, CM, stabilised with 1.5% CF
and cured for 7 days also experienced a strength gain in the first 12 f-t cycles before
declining in the next 12 f-t cycles, as shown in Table 5.11 and Figure 5.27.

184
Chapter 5 F-t, w-d durability study

A possible reason for increases in the UCS during the first 12 f-t cycles could be that the
strength gain due to hydration or pozzolanic reactions outweighed the adverse f-t
effects. It is important to note that the stabilised materials continuously gained strength
during the f-t cycles. In the early days of curing, due to the high cementitious reaction
rate, stabilised materials gain strength a rapid rate. As the series 1 specimens were cured
for only 7 days, their rapid strength gains could have overcome deleterious f-t effects.
However, when strength development slowed, the f-t effects could have begun to
dominate and resulted in decline in strength at the end of 24 f-t cycles.

Table 5.9: Testing program for UCS testing of materials stabilised with CF binder

Granular Binder Number of f-


Series Curing Procedure
Material Content t Cycles
QM 7 days – normal 1.5% 1 0, 1, 4, 12, 24

1 QM 7 days – normal 3.0% 1 0, 1, 4, 12 ,24

CM 7 days – normal 1.5% 1 0, 12, 24


28 days – 0, 4, 12
QM 1.5% 1
normal
2
28 days –
CM 1.5% 1 0, 4, 12
normal
QM 7 days at 40oC 1.5% 2 0, 4, 12
3
CM 7 days at 40oC 1.5% 2 0, 4, 12

QM 7 days – normal 1.5% 2 0, 4, 12


4
CM 7 days – normal 1.5% 2 0, 4, 12

The effect of hydration/pozzolanic reaction is further noticed in the next series of


specimens which were cured for 28 days subjected to f-t procedure 1. As shown in table
5.12, specimens of QM showed 13% of strength loss even after 12 f-t cycles and at the
end of 24 f-t cycles, further 19% strength loss was observed. Similar trend was observed
for specimens on specimens stabilised with CM as well in which strength reduced by
9% after 24 f-t cycles.

185
Chapter 5 F-t, w-d durability study

Comparing these results with those from Series 1, it can be seen the specimens cured for
28 days showed decline in strength after fewer f-t cycles than those cured for 7 days.
The rate of increase in strength reduced over time, as can be observed in Figure 5.29,
where the UCS gained gradually with the age (number of cycles) of a sample.
Therefore, the rate of strength gain of the 28-day cured specimens were less than those
of the 7-day cured specimens which were not sufficiently high to resist adverse f-t
effects and were noticeable as early as 4 f-t cycles. When increases in strength due to
cementitious reactions are insignificant or less sufficient, the deleterious effect of f-t
could become noticeable. Khoury and Zaman (2006) also reported the effect of the
curing age on durability against f-t. Specimens cured for 28 days showed increases in
the Mr up to 12 cycles which reduced after 30 while those cured for 3 days did not show
any reduction in the Mr up to 30 f-t cycles.

According to Fredrick Lekarp (2000), the Mr of unbound granular materials are also less
affected by the loading duration. However, Kim and Tutumluer (2006) reported that the
loading duration had a considerable effect on the PD behaviour of unbound granular
materials. Two different axial loading rates of 0.1 and 0.5 second were selected for the
study, axial pressures of 63 and 336 kPa applied with a constant confining pressure of
21 kPa and each load combination subjected to 72,000 load cycles. It was reported that
the 0.5 second loading accumulated 40% more PD than load duration for 0.1 second. It
is important to note that the Mr may not be a good indicator of PD behaviour and the
loading frequency may influence the rutting behaviour of a lightly stabilised granular
material. Therefore, further PD studies are needed to determine the effect of the loading
frequency on a lightly stabilised granular material.

186
Chapter 5 F-t, w-d durability study

Table 5.10 Variations in UCS with number of f-t cycles for QM stabilised with
1.5% and 3% CF binders

No. of
Parent
Curing Number of f-t
Material & UCS Average CoV
Days & f-t Cycles
Binder
Procedure
0 1.66
0 1.73 1.71 2
0 1.73
1 2.06
1 1.84 1.99 6
1 2.06
7-day 4 2.30
QM
f-t procedure 4 2.67 2.37 11
1.5% CF
1 4 2.15
12 2.67
12 2.86 2.56 14
12 2.15
24 1.67
24 1.66 1.68 2
24 1.71
0 4.18
0 4.09 4.38 10
0 4.86
1 4.74
1 4.26 4.44 6
1 4.30
7-day 4 4.22
QM
f-t procedure 4 4.66 4.56 7
3% CF
1 4 4.81
12 6.02
12 5.83 5.81 4
12 5.58
24 5.01
24 4.45 4.77 6
24 4.85

As previously noted, one of the objectives of this study was to find a suitable testing
method for assessing the durability of lightly stabilised materials against f-t within a
reasonable amount of time. Therefore, it is important that the selected testing procedure
should be able to capture the adverse f-t effects (decline in UCS) within the selected
number of f-t cycles. The variables in Series 1 or 2 may not have been the best choices
for assessing f-t effects because, to notice them, 7-day curing (Series 1) required 24 f-t

187
Chapter 5 F-t, w-d durability study

cycles whereas 28-day curing (Series 2) needed 12 which would take more than 50
days. Therefore, it is important to make changes in the existing procedure and variables
to find a better laboratory durability procedure for assessing f-t effects.

Table 5.11: Variations in UCS with number of f-t cycles for specimens cured for 7
days

No. of
Parent
Curing Number of f-t
Material & UCS Average CoV
Days & f-t Cycles
Binder
Procedure
1.83
0 1.55 1.73 9
1.82

7-day 2.12
CM
f-t 12 1.93 2.14 11
1.5% CF
procedure 1 2.38

1.80
24 2.05 1.91 7
1.87

Based on the previous findings, f-t effects could become significant with fewer cycles if
the strength gains due to hydration/pozzolanic reaction were controlled during f-t
cycles. In other words, specimens should gain significant amounts of strength before
being subjected to f-t cycles. This will enable a specimen to possess fewer unreacted
cementitious compounds and, as a result, during f-t cycles, its strength gain due to
hydration/pozzolanic reaction can be minimised. As one possible option is to cure
specimens at an elevated temperature, it was decided to cure them at 40oC rather than
the conventional 23oC curing temperature and carry out testing on another series of
specimens.

188
Chapter 5 F-t, w-d durability study

3.2

2.8

2.4
UCS (MPa)

1.6

1.2

0 5 10 15 20 25
Number of f-t cycles

Figure 5.27: Variations in UCS with number of f-t cycles for QM and CM with
1.5% CF binder

6
UCS (MPa)

0 5 10 15 20 25
Number of f-t cycles

Figure 5.28: Variations in UCS variations with number of f-t cycles for QM and
CM with 1.5% CF binder

189
Chapter 5 F-t, w-d durability study

Table 5.12: UCS after different numbers of f-t cycles for QM and CM specimens
cured for 28 days

Parent No. of Curing


Number of f-t
Material & Days & f-t UCS Average Cof
Cycles
Binder Procedure
2.15
0 2.15 2.20 4
2.30
1.91
QM 28-day
4 1.92 1.92 0.2
1.5% CF f-t procedure 1
1.92
1.69
12 1.44 1.49 12
1.34
2.21
0 2.34 2.29 3
2.32
2.22
CM 28-day
4 2.33 2.22 5
1.5% CF f-t procedure 1
2.10
2.10
12 2.14 2.10 2
2.07

Moreover, another f-t procedure for studying the f-t effect and evaluating its impact on
the UCS was considered. In f-t procedure 2, moisture pads were kept underneath the
specimens during the thaw cycle, as explained in section 5.2.4.1. The aim of this
approach was to supply water during the thaw phase which was expected to produce
more deterioration in the specimens. ASTM D560 (2003) also suggests placing
absorbent pads beneath the samples to supply free water during the f-t cycle as this
better simulates field conditions.

These two changes were adopted for a new set of specimens (Series 3) which were
cured at 400C and subjected to f-t procedure 2. Table 5.13 shows the UCS measured
after 0, 4 and 12 f-t cycles. After 7 days of curing at 400C, both QM and CM achieved
strengths 27% and 36% higher respectively than those of a 7-day normally cured
specimen and approximately equal to 28-day strength. As previously mentioned, curing
at 400C proved to be a better approach for accelerating the early development of

190
Chapter 5 F-t, w-d durability study

strength. As a result, strength losses were observed even after 4 f-t cycles and, after 12,
QM and CM showed decline in UCS of 31% and 24% respectively.

2.6

6
QM_1.5% CF_normally cured
QM_3.0% CF_normally cured
Line connecting average UCS
Line connecting average UCS
2.4
5.6

2.2
UCS(MPa)

5.2

UCS(MPa)
2
4.8

1.8
4.4

1.6 4

0 5 10 15 20 25 0 5 10 15 20 25
Number of Cycles (one cycle = 2 days) Number of cycles (one cycle = two days)
(a) (b)

Figure 5.29: Strength-gaining patterns of QM lightly stabilised granular material


with (a) 1.5% CF binder (b) 3.0% CF binder

Finally, another set of specimens (Series 4) was tested similar to Series 1 except the f-t
procedure in order to evaluate the effects of f-t procedure. This set of specimens were
cured for 7 days, subjected to f-t procedure 2 and tested for UCS. Table 5.14 shows the
results after 0, 4 and 12 f-t cycles in which after 4 f-t cycles, the QM specimens did not
show much decrease in strength but, after 12 f-t cycles , lost 14% while the CM
specimens showed 6% reductions in UCS after 12 f-t cycles. It should be noted that the
strength losses observed after 12 f-t cycles were higher than those of the 7-day cured
specimens subjected to f-t procedure 1, i.e., Series 1. The destructive nature of f-t
procedure 2 was noticeable from the UCS of the Series 1 and 4 specimens. Series 2 and
3 also exhibited similar observation which can be seen from table 5.11 and 5.13. The
specimens cured for 28 days (Series 2) and 400C (Series 3) achieved approximately
equal UCS at the end of their respective curing periods but the f-t procedures adopted
for them were different, with f-t procedure 1 used for the former and f-t procedure 2 for
the latter. After 12 f-t cycles, QM retained almost the same UCS (1.5 MPa) but a

191
Chapter 5 F-t, w-d durability study

significant difference could be observed for CM. Specimens subjected to procedure 2


attained average UCS of 1.8 MPa whereas specimens subjected to procedure 1 showed
UCS of 2.1 MPa. Therefore, it was apparent that the f-t procedure 2 had more
destructive effects than f-t procedure 1.

Table 5.13: UCS after different numbers of f-t cycles for QM and CM specimens
tested using f-t procedure 2 and cured at 400C

Average
Number of Number of f-t UCS
Material Binder UCS CoV
Curing Days Cycles (MPa)
(MPa)
2.18
0 2.10 2.16 3
2.21
1.79
1.5%
QM 7 days at 400C 4 1.76 1.81 3
CF
1.87
1.55
12 1.50 1.50 3
1.46
2.36
0 2.31 2.38 3.3
2.46
2.03
1.5%
CM 7 days at 400C 4 2.31 2.21 6.9
CF
2.28
1.70
12 1.82 1.80 5.1
1.87

192
Chapter 5 F-t, w-d durability study

Table 5.14: UCS after different numbers of f-t cycles for QM and CM specimens
tested using f-t procedure 2

Number of Number of f-
Material Binder UCS Average CoV
Curing Days t Cycles
1.66
0 1.73 1.71 2
1.73
1.56
1.5%
QM 7 4 1.62 1.62 4
CF
1.68
1.40
12 1.49 1.47 4
1.52
1.83
0 1.55 1.73 9
1.82
1.73
1.5%
CM 7 4 1.76 1.77 3
CF
1.83
1.69
12 1.51 1.62 6
1.67

Comparison with series 3 and 4 demonstrates that, series 4 is more destructive than
series 3. In absolute values after 12 f-t cycles, QM and CM in series 3, reduced by 0.66
MPa and 0.48 MPa respectively. In contrast, series 4 showed only 0.24 MPa and 0.11
MPa reductions for QM and CM respectively. It is possibly because of the different rate
of strength gain pattern of both series which is discussed in detail later in this chapter.

All four series of tests showed that the number of f-t cycles required to notice the
adverse effect of f-t depended on the combinations of the selected variables. Series 1
required up to 24 f-t cycles whereas Series 2, 3 and 4 showed significant losses of
strength after 12 f-t cycles. However, Series 2 may not have been a suitable laboratory
test because of its longer testing time. Therefore, based on the UCS test results, Series 3
and 4 could be used to assess the f-t effects on lightly stabilised granular materials.

193
Chapter 5 F-t, w-d durability study

The discussion so far has focussed on comparing the UCS after a certain number of f-t
cycles with those with 0 f-t cycles to assess the durability of a lightly stabilised material
against f-t effects. However, comparing specimens subjected to f-t cycles against
specimens of similar ages cured in a normal curing environment could reveal factors
that may affect UCS values; for example, specimens cured for 7 days and subjected to
12 f-t cycles aged 31 days are comparable to those cured for 31 days in a normal curing
environment. As previously mentioned, the rate of cementitious or pozzolanic reaction
due to the freezing phase is one of the main differences between normally cured
specimens and those subjected to f-t cycles because the environmental parameters of the
thawing phase and normal curing environment are similar (23oC and 90-95% RH).

The temperature has an influence on hydration and pozzolanic reaction where it had
been shown that higher temperature accelerates the reaction (Neville and Brooks 2010).
Figure 5.30 shows the influence of temperature on a concrete strength-gaining pattern.
Although hydration still takes place in a specimen at a curing temperature of -50C, the
strength gain is significantly lower than that of a specimen cured at higher positive
temperatures. It is also important to note that there is an ‘optimum temperature’ beyond
which an increase in temperature does not benefit strength development. Therefore, it is
likely that normally cured specimens gained more strength than those subjected to f-t
cycles. Moreover, the adverse effect of f-t also impairs the strength of specimens
subjected to f-t cycles. Therefore, it can be expected that normally cured specimens
show higher UCS than similarly aged specimens subjected to f-t cycles. Khoury and
Zaman (2007) presented a similar comparison of Mr of 28-day cured specimens
subjected to 30 f-t cycles and 90-day normally cured specimens and showed that the
above theory is true. Therefore, another set of specimens cured in a normal curing
environment was tested for different curing periods matching the number of days
specimens were subjected to f-t cycles. As shown in Figure 5.31, after 24 f-t cycles (f-t
procedures 1 and 2), similar results were obtained, whereby normally cured specimens
showed higher UCS than normally cured specimens subjected to f-t cycles of similar
age.

194
Chapter 5 F-t, w-d durability study

Effects of curing temperature


50
28-day curing
7-day curing
3-day curing
1-day curing
40

Compressive strength (MPa)

30

20

10

-10 0 10 20 30 40 50
Temperature ( oC)

Figure 5.30: Influence of temperature on strength of concrete cast and cured at the
temperature indicated (Neville and Brooks 2010)

One more noticeable difference between normally cured specimens and those subjected
to f-t cycles is the retained moisture content immediately before testing. Therefore, to
investigate the effect of moisture, at the end of the UCS test, the moisture contents of all
the tested specimens were measured. The moisture variations of the specimens
subjected to f-t procedure 1 (Series 1 and 2) are shown in Figure 5.32 which
demonstrates that the moisture content gradually decreased with the number of f-t
cycles and ended up having only 2% after 24 f-t cycles. Figure 5.33 shows the moisture
contents measured for the Series 3 and 4 specimens for which f-t procedure 2 was
adopted. Compared with the f-t procedure 1 specimens, the f-t procedure 2 specimens
retained a 1.5-2.5% higher moisture content because of the moisture pads kept
underneath them during the thaw cycle. Moreover, after 12 f-t cycles, the moisture
contents of all the specimens were recorded as being above 5%. However, the normally
cured specimens of QM retained moisture contents of between 7 and 7.5% irrespective

195
Chapter 5 F-t, w-d durability study

of the number of curing days. Therefore, as well as from deleterious f-t effects, it’s
likely that differences between the moisture contents of specimens subjected to f-t
cycles and normally cured specimens could also affect their strengths to varying degrees
depending on the moisture level.

Moreover, the relationship between the UCS and moisture contents of normally cured
specimens was established in a separate study of QM in which the specimens were
compacted at OMC (9%) and cured for 14 days. After curing, they were dried in the
oven at 600C for different time durations to obtain different post-compaction moisture
contents before being tested for their UCS. As shown in Figure 5.34, decreases in the
moisture content up to 5% helped the specimens gain strength which was similar to the
outcome reported by Nazarian and Yuan (2008) for the seismic modulus of a fine-
grained material compacted at different compaction moisture contents. This moisture
content and UCS relationship can be explained using the concept of unsaturated soil
mechanics.

The effective stress of unsaturated soil can be determined using equation 5.8 which was
proposed by Bishop (1959). In this equation, X w is 1 for saturated soils and zero for dry
soils while matric suction is high at low moisture content and decreases with an increase
in the moisture content. Although X w and matric suction have opposite trends, there
exists a moisture content at which their combined effect is maximum, giving maximum
strength and stiffness, which is on the dry side of the OMC.

 '  (  u a )  X w (u a  u w ) Eq. (5.8)

where:  ' = effective stress

 = total stress

X w =Bishop’s effective stress parameter and Ua and Uw are the pore air and pore water
pressures respectively.

Therefore, decreases in the moisture content until the optimum point is reached are
likely to increase the UCS. As the normally cured specimens did not show great
moisture variations with different numbers of curing days, the moisture effects were
minimal. However, for f-t specimens, containing more moisture is detrimental, as could

196
Chapter 5 F-t, w-d durability study

easily be seen from specimens subjected to the f-t 1 and f-t 2 testing procedures. When a
specimen is subjected to a f-t cycle with a high moisture content, the destructive nature
of f-t cycle is likely to be higher than specimens with lower moisture content. The water
in the voids expands while freezing which cause microstructural changes and
disintegrates the soil-water structure. With higher water contents in the voids, the
heaving pressure is high which creates a more destructive force and this was the reason
for f-t procedure 2 being more destructive than f-t procedure 1. Khoury and Zaman
(2006) also adopted two different procedures and presented similar observations. In
their study, one f-t procedure was similar to the f-t procedure 1 adopted in this study
while, in the other, free water was made available through a chamber placed beneath the
samples and filled with gravel. The results showed that, due to the supply of water, the
moisture content increased by 2.3% after 8 cycles compared with that of the normal
procedure and excessive degradation in flexural properties was observed.

Regarding f-t procedure 1, decreases in the moisture content with the number of f-t
cycles was likely to limit adverse f-t effects. Going back to Figure 5.31 where
specimens subjected to f-t procedure 1 showed higher strengths than normally cured
specimens up to 12 f-t cycles. As previously mentioned, this could be because of the
decline in the moisture content with the number of f-t cycles that could have limited the
destructive f-t effects. Moreover, decreasing moisture content also increases the
effective stress which eventually might have increased the UCS. The deleterious effect
of f-t cycles, further decreased the moisture content and its adverse effect on the UCS,
and the decreasing rate of cementitious reactions caused a decline in strength after 12 f-t
cycles.

Therefore, based on this study, for a lightly stabilised granular material, f-t procedure 2
is a better choice than f-t procedure 1 because: firstly, it better simulates field
conditions; and, secondly, the deleterious effect is noticeable even after 4 f-t cycles
which is important in a laboratory study where the timeframe has to be reasonably short.
Also, its experimental procedure is easy to implement.

197
Chapter 5 F-t, w-d durability study

Normally cured specimens


60 Specimens subjected to f-t procedure 1
Specimens subjected to f-t procedure 2

40

Percentage in crease in UCS

20

-20

-40

0 5 10 15 20 25
Number of cyles

Figure 5.31: Percentage changes in UCS relative to 0 cycles of normally cured QM


specimens and those subjected to f-t procedures 1 and 2

MC Vs # f-t cycles
1.5%_QM
6 1.5%_QM_Avg
3%_QM
3%_QM_Avg
Moisture content (%)

0 5 10 15 20 25
Number of f-t cycles

Figure 5.32: Variations in moisture content with number of f-t cycles for QM with
1.5% and 3% CF binders

198
Chapter 5 F-t, w-d durability study

7.5
MC Vs # f-t cycles
QM
CM
7 QM_ft2
CM_ft2

Moisture content (%) 6.5

5.5

0 4 8 12
Number of f-t cycles

Figure 5.33: Variations in moisture content with number of f-t cycles for QM and
CM subjected to f-t procedures 1 and 2

2.4

14 Days UCS
14 Days UCS Trend Line

2
UCS (MPa)

1.6

1.2

0.8

0 2 4 6 8 10
Moisture Content (%)

Figure 5.34: Variations in UCS with different moisture contents

199
Chapter 5 F-t, w-d durability study

Another observation from this study was the differences between CM and QM in terms
of their vulnerability to f-t cycles. All the UCS values presented earlier in this section
demonstrated that CM possessed higher strength than QM. Figure 5.35 shows the
strength losses and gains in all four series tested for f-t cycles, where losses are
considered positive. In each series, strength losses were calculated relative to the UCS
of specimens not subjected to any f-t cycles (0 f-t cycles). In Series 2 and 4, significant
differences in percentages of strength loss could be observed, with the UCS of QM
lower than that of CM. Also, in Series 3, the effects of f-t were greater for QM.
However, as shown in the first set of bars in Series 1, CM did not show as great a
strength gain as QM and demonstrated less susceptibility to f-t cycles. One possible
reason for this could be the differences in the materials’ particle size distributions.

Considering both soil matrices, where QM’s soil matrix contains more fine soil particles
than that of CM, as shown in Figure 5.36, the coefficient of uniformity for QM is 80
and, for CM, 24 which shows that QM has wider and more well-graded particle size
distribution (Table 5.15). A well-graded material tends to have smaller void spaces and
require more moisture to achieve the MDD. The reason for QM having a 9% OMC
compared with CM’s 8% is the differences in their fines content. Larger fines content
helped to fill the voids, there were smaller void spaces, and a higher MDD was
achieved. With smaller void spaces, the f-t effect was expected to be high because of the
volume changes in the voids that occurred during the freezing phase when the liquid
water changed to ice which damaged the soil matrix and resulted in less strength and
stiffness. As QM possessed smaller voids, its soil matrix was considerably more
affected than that of CM, as can be seen from the strength measurements after different
numbers of f-t cycles.

200
Chapter 5 F-t, w-d durability study

QM
40 CM

20
UCS loss %

Series 1
0
Series 2 Series 4
Series 3

-20

-40

Figure 5.35: Comparisons of QM and CM

Table 5.15: Physical properties determined from sieve analysis and MDD

Properties QM CM
Dry density 2.3 kN/m3 2.1 kN/m3
OMC 9% 8%
D10 0.075 mm 0.25 mm
D60 6 mm 6 mm
Cu (D60/ D10) 80 24

201
Chapter 5 F-t, w-d durability study

100

CM
QM
80

Percentage Passing(%)

60

40

20

0
0.01 0.1 1 10 100
Particle Size (mm)

Figure 5.36: Particle size distributions of CM and QM

5.3.2.2 Freeze-Thaw Durability Study using Repeated Load Triaxial Test

The absence of any widely available durability procedure of a lightly stabilised material
demanded different tests and other variables. In the previous section, series of UCS tests
revealed certain aspects of each testing procedure and the effects of different variables.
However, as the RLT test is preferred than UCS because of its cyclic loading and ability
to simulate different stress conditions, it was used for the set of combinations shown in
Table 5.16. In the RLT tests, specimens subjected to f-t cycles were tested in terms of
their Mr, PDs and monotonic shear stresses.

202
Chapter 5 F-t, w-d durability study

5.3.2.2.1 Freeze-Thaw Effects on Resilient Modulus

Table 5.16: Program for RLT testing of materials stabilised with 1.5% CF binder

Granular F-t Number of f-


Series Curing Binder
Material Procedure t Cycles

CM 7 days - normal 1.5% 1 0,12,24


1
QM 7 days - normal 1.5% 1 0,12,24

CM 28 days - normal 1.5% 1 0,12,24


2
QM 28 days - normal 1.5% 1 0,12,24

3 CM 7 days at 400C 1.5% 2 0,4,12

4 CM 28 days - normal 1.5% 2 0,4,12

As described in section 5.2.2.6, the Mr at 16 selected stress sequences were determined


using RLT Mr testing, with details of the procedure presented in section 5.2.2. This
section discusses the Mr results obtained from the testing program shown in Table 5.16
all of which are presented in both tabular and graphical formats. As a graphical
representation is a convenient way of comparing the Mr from different tests, the k-θ
model, which relates the Mr to the bulk stress, was chosen to evaluate the results. The
regression equations obtained above were then used to determine the Mr for different
stress sequences, with typical calculated variations in the Mr with the number of f-t
cycles presented for each series of tests for stress sequences 4, 8, 12 and 16. Moreover,
the regression equations suggested that the coefficient of determination (R2) remained
high (<0.75) for all the tests except those with specimens cured for 400C and subjected
to 4 f-t cycles. Therefore, as the selected model was proven to be statistically
significant, it could be used to predict the Mr values.

Table 5.17 presents the Mr of QM cured for 7 days, subjected to f-t procedure 1 (Series
1) and tested after 0, 12 and 24 f-t cycles. As shown in Figure 5.37, the selected k-θ
model demonstrated good correlations between the Mr and bulk stress for all three tests.
The dependency of the Mr on the bulk stress can be noted in both Table 5.17 and Figure
5.37, as the Mr increased from a lower to higher value with an increase in the bulk

203
Chapter 5 F-t, w-d durability study

stress. As shown in Figure 5.38, although after 12 f-t cycles, a 30% increase in the Mr
(stress sequence 16, θ=1407.6 kPa) over that of 0 f-t cycles was observed, the Mr
reduced significantly after 24 f-t cycles and achieved values almost equal to those of 0
f-t cycles. This is consistent with the UCS test results which showed that an increase in
the UCS for the first 12 f-t cycles was followed by a decrease in UCS afterwards.

CM specimens were tested similarly using the same set of variables. Table 5.18 and
Figure 5.39 show the Mr measured after 0, 12 and 24 f-t cycles. The Mr values for 0 f-t
cycle were higher for CM than QM which demonstrates that CM possesses higher
stiffness characteristics. Moreover, as shown in Figure 5.40,similar to QM, there was an
increase in the Mr after 12 f-t cycles followed by a decrease after 24 f-t cycles,.
However, unlike QM, after 24 f-t cycles, CM retained considerably lower values of the
Mr than after 0 f-t cycles, showing a 23% decrease in the Mr (stress sequence 16,
θ=1407.6 kPa).

Series 1 test results indicates that the f-t cycles did not have much influence in the first
12 f-t cycles. Next 12 f-t cycles caused decline in Mr which can be clearly seen from the
test results both QM and CM. First impression from Figures 5.38 and 5.40 is that the
changes in absolute values in Mr between different the tests are approximately equal
irrespective of the bulk stress (it can be seen from the constant grading line). In other
words, constant reduction in Mr can be noticed between specimens subjected to
different f-t cycles, irrespective of stress level.

204
Chapter 5 F-t, w-d durability study

Table 5.17: Mr of QM stabilised with 1.5% CF binder, cured for 7 days and
subjected to f-t procedure 1

Stress 0 Cycle 12 Cycles 24 Cycles


Sequence (MPa) (MPa) (MPa)
1 260 435 329
2 291 466 364
3 361 554 418
4 391 558 401
5 392 591 391
6 440 617 433
7 470 654 476
8 476 679 439
9 469 676 390
10 509 706 446
11 570 787 545
12 576 792 553
13 572 770 516
14 626 826 596
15 657 882 670
16 687 929 738

1000
0 f-t cycle
12 f-t cycles
24 f-t cycles
0 f-t cycle
Resilient Modulus (MPa)

800 12 f-t cycles


24 f-t cycles

600

400

200
200 300 400 500 600 700 800 900
1000
Bulk Stress (kPa)

Figure 5.37: Variations in Mr with bulk stress of QM stabilised with 1.5% CF


binder, cured for 7 days and subjected to f-t procedure 1

205
Chapter 5 F-t, w-d durability study

Mr = k 1(?)k 2
Sequence 4_ ? = 422.3
Sequence 8_ ? = 703.8
1000 Sequence 12_ ? = 1055.7
Sequence 16_ ? = 1407.6

800
Resilient modulus (MPa)

600

400

200

0 5 10 15 20 25
Number of f-t cycles

Figure 5.38: Variations in Mr with number of f-t cycles for typical stress sequences
4, 8, 12 and 16 (Series 1, QM)

Table 5.18: Mr of CM stabilised with 1.5% CF binder, cured for 7 days and

subjected to f-t procedure 1

Stress 0 Cycle 12 Cycles 24 Cycles


Sequence (MPa) (MPa) (MPa)
1 439 679 481
2 532 789 538
3 689 852 639
4 725 911 703
5 635 890 707
6 759 928 806
7 814 1038 834
8 733 1000 830
9 680 944 786
10 775 1006 842
11 888 1111 985
12 885 1152 993
13 853 1123 959
14 959 1220 1021
15 1008 1257 1119
16 941 1203 1075

206
Chapter 5 F-t, w-d durability study

0.4232
Mr=56.093( )
R2=0.93
0 f-t cycle
0.2793
12 f-t cycles Mr=175.06( )
24 f-t cycles R2=0.92
1200
0 f-t cycle
12 f-t cycles 0.3807
Mr=67.463( )
24 f-t cycles
R2=0.85

1000
Resilient Modulus (MPa)

800

600

400

100 1000
Bulk Stress (kPa)

Figure 5.39: Variations in Mr with bulk stress of CM stabilised with 1.5% CF


binder, cured for 7 days and subjected to f-t procedure 1

Mr = k1(?)k2
Sequence 4_ ? = 422.3
Sequence 8_ ? = 703.8
Sequence 12_ ? = 1055.7
1400
Sequence 16_ ? = 1407.6

1200
Resilient modulus (MPa)

1000

800

600

0 5 10 15 20 25
Number of f-t cycles

Figure 5.40: Variations in Mr with number of w-d cycles for typical stress
sequences 4, 8, 12 and 16 (Series 1, QM)

207
Chapter 5 F-t, w-d durability study

The next series of specimens were cured for 28 rather than 7 days before being
subjected to f-t cycles. Table 5.19 presents the Mr obtained for QM after 0, 12 and 24 f-t
cycles. Similar to the specimens in Series 1, this series also showed good correlations
between the Mr and bulk stress, as shown in Figure 5.41.

The Mr of the QM specimens subjected to 12 f-t cycles retained lower values than 0 f-t
cycle, except for the last stress sequence. However, after 24 f-t cycles, significant
decline in the Mr, 25-30% for the stress sequences were observed, as shown in Figure
5.42. The Mr results obtained for CM is shown in Figure 5.43 and Table 5.20. After 12
f-t cycles greatly deteriorated the material which was exacerbated after 24 f-t cycles.
Figure 5.44 illustrates that a greater than 40% decline in the Mr was recorded for the
first 12 f-t cycles and the next 12 f-t cycles caused further damage, except for stress
sequence 16. The effect of curing can be noticed in the Series 1 and 2 Mr test results
which show that the former, which were cured for 7 days, did not show deterioration for
the first 12 f-t cycles because of their strength gain whereas the latter, which were cured
for 28 days, deteriorated much more quickly and lost considerable stiffness after 24 f-t
cycles.

Table 5.19: Mr of QM stabilised with 1.5% CF binder, cured for 28 days and
subjected to f-t procedure 1

Stress 0 Cycle 12 Cycles 24 Cycles


Sequence (MPa) (MPa) (MPa)
1 587 464 412
2 607 481 459
3 668 535 541
4 640 542 506
5 657 569 510
6 711 599 546
7 717 630 545
8 685 650 502
9 697 690 479
10 735 700 517
11 824 760 600
12 825 766 587
13 815 790 582
14 907 848 650
15 923 865 682
16 902 920 675

208
Chapter 5 F-t, w-d durability study

0 f-t cycle
1000 12 f-t cycles
24 f-t cycles
0 f-t cycle
12 f-t cycles
24 f-t cycles

0.3938
0.2597
Mr=53.663( )
800 Mr=140.87( ) R2=0.93
Resilient Modulus (MPa)

R2=0.85
0.2329
Mr=123.48( )
R2=0.76

600

400

100 1000
Bulk Stress (kPa)

Figure 5.41: Mr varying with bulk stress of QM stabilised with 1.5% CF binder,
cured for 28 days and subjected to f-t procedure 1

Mr = k 1(?)k 2
1000
Sequence 4_ ? = 422.3
Sequence 8_ ? = 703.8
Sequence 12_ ? = 1055.7
Sequence 16_ ? = 1407.6
900
Resilient modulus (MPa)

800

700

600

500

0 5 10 15 20 25
Number of f-t cycles

Figure 5.42: Mr varying with number of w-d cycles for typical stress sequences 4, 8,
12 and 16 (Series 2, QM)

209
Chapter 5 F-t, w-d durability study

Table 5.20: Mr of CM stabilised with 1.5% CF binder, cured for 28 days and
subjected to f-t procedure 1

Stress 0 Cycle 12 Cycles 24 Cycles


Sequence (MPa) (MPa) (MPa)
1 934 483 311
2 1014 494 343
3 1225 597 428
4 1415 632 505
5 1271 598 532
6 1456 702 595
7 1562 772 688
8 1458 792 732
9 1331 742 689
10 1496 805 777
11 1664 914 898
12 1655 912 890
13 1536 845 866
14 1758 953 957
15 1722 1021 1036
16 1552 1004 1023

0.2916
0 f-t cycle Mr=228( )
2000 12 f-t cycles R2=0.86
24 f-t cycles
0 f-t cycle
12 f-t cycles
24 f-t cycles
1600
Resilient Modulus (MPa)

1200

0.3917
Mr=64.898( )
800
R2=0.96

400 0.7036
Mr=7.4477( )
R2=0.94

100 1000
Bulk Stress (kPa)

Figure 5.43: Mr varying with bulk stress of CM stabilised with 1.5% CF binder,
cured for 28 days and subjected to f-t procedure 1

210
Chapter 5 F-t, w-d durability study

Mr = k 1(?)k 2
2000
Sequence 4_ ? = 422.3
Sequence 8_ ? = 703.8
Sequence 12_ ? = 1055.7
Sequence 16_ ? = 1407.6
1600
Resilient modulus (MPa)

1200

800

400

0 5 10 15 20 25
Number of f-t cycles

Figure 5.44: Mr varying with number of w-d cycles for typical stress sequences 4, 8,
12 and 16 (Series 2, CM)

Although the previous two series of tests are potential laboratory durability tests, the
times they took to complete the tests are not acceptable. It is important to note that a f-t
test could be considered a potential test when it actually begins to capture the f-t effects
(i.e., a decline in strength or stiffness). In Series 1 and 2, this required minimum of 24
and of 12 f-t cycles respectively which is more than 50 days, that is not acceptable for a
typical durability test. . Therefore, as previously discussed, two more series with
different f-t procedures were carried out.

The Series 3 and 4, Mr tests were conducted adopting f-t procedure 2 on CM. Unlike the
two previous series, Mr tests after 0, 4 and 12 f-t cycles were considered. Table 5.21 and
Figure 5.45 show the results for Series 3 specimens which were cured for 7 days and
subjected to f-t procedure 2. Comparing the results for 0 and 12 f-t cycles, although
their differences were marginal at lower stresses, reductions in their Mr at 156 MPa
were measured for the last load sequence, as shown in Figure 5.46, while 4 f-t cycles
did not cause significant changes in the Mr.

211
Chapter 5 F-t, w-d durability study

The Series 4 Mr results showed significant amounts of destruction after 4 and 12 f-t
cycles, as shown in Table 5.22 and Figure 5.47, which can quite clearly be seen at both
low and high stresses. As shown in Figure 5.48, after 4 and 12 f-t cycles, Mr losses of
14% and 43% respectively were observed (θ=1407.6 kPa). The introduction of moisture
pads as well as accelerated curing escalated the destructibility of the f-t cycles.
Accelerated curing is noticeable in the 0 f-t test results where the stiffness is higher than
series one 0 f-t specimen which was cured for 7 days under normal curing environment
and almost equal to 28 days cured specimens which was cured for 28 days.

Table 5.21: Mr of CM stabilised with 1.5% CF binder, cured for 7 days and
subjected to f-t procedure 2

Stress 0 Cycle 4 Cycles 12 Cycles


Sequence (MPa) (MPa) (MPa)
1 439 587 505
2 532 607 542
3 689 668 572
4 725 640 539
5 635 657 582
6 759 711 619
7 814 717 651
8 733 685 674
9 680 697 699
10 775 735 739
11 888 824 774
12 885 825 772
13 853 815 778
14 959 907 851
15 1008 923 865
16 941 902 879

212
Chapter 5 F-t, w-d durability study

0 f-t cycle
4 f-t cycles 0.3807
Mr=67.463()
12 f-t cycles
0 f-t cycle R2=0.85
4 f-t cycles 0.2597
1000 Mr=140.87()
12 f-t cycles
R2=0.85
0.3143
Resilient Modulus (MPa) Mr=93.17()
R2=0.89
800

600

400

100 1000
Bulk Stress (kPa)

Figure 5.45: Mr of Canberra material stabilised with 1.5% CF binder, cured for 7
days and subjected to f-t procedure 2

1100 Mr = k 1(?)k 2
Sequence 4_ ? = 422.3
Sequence 8_ ? = 703.8
Sequence 12_ ? = 1055.7
1000 Sequence 16_ ? = 1407.6
Resilient modulus (MPa)

900

800

700

600

0 4 8 12
Number of f-t cycles

Figure 5.46: Mr varying with number of w-d cycles for typical stress sequences 4, 8,
12 and 16 (Series 3, CM)

213
Chapter 5 F-t, w-d durability study

Table 5.22: Mr of CM stabilised with 1.5% CF binder, cured for 7 days at 40oC
and subjected to f-t procedure 2

Stress 0 Cycle 4 Cycles 12 Cycles


Sequence (MPa) (MPa) (MPa)
1 777 543 217
2 834 615 235
3 1094 761 283
4 1268 836 326
5 1020 742 324
6 1178 868 367
7 1370 945 417
8 1287 730 487
9 1078 544 432
10 1199 667 494
11 1393 825 609
12 1382 822 615
13 1220 684 544
14 1402 831 645
15 1496 975 762
16 1393 1045 805

0.2987
0 f-t cycle Mr=181.16()
1600 4 f-t cycles R2=0.81
12 f-t cycles
0 f-t cycle
4 f-t cycles
12 f-t cycles

0.2921
1200 Mr=163.06()
R2=0.42
Resilient Modulus (MPa)

800
0.7579
Mr=3.7061()
R2=0.97

400

100 1000
Bulk Stress (kPa)

Figure 5.47: Mr of CM stabilised with 1.5% CF binder, cured for 7 days at 40oC
and subjected to f-t procedure 2

214
Chapter 5 F-t, w-d durability study

Mr = k 1(?)k 2
Sequence 4_ ? = 422.3
1600
Sequence 8_ ? = 703.8
Sequence 12_ ? = 1055.7
Sequence 16_ ? = 1407.6

1200

Resilient modulus (MPa)

800

400

0 4 8 12
Number of f-t cycles

Figure 5.48: Mr varying with number of w-d cycles for typical stress sequences 4, 8,
12 and 16 (Series 3, CM)

There are many similarities in the outcome of this Mr test results compared with UCS.
The influence of strength gain, number of curing days, curing method and f-t procedure
can also be noticed in the Mr test. Series 1 test results indicates that the f-t cycles did not
have much influence in the first 12 f-t cycles. Next 12 f-t cycles caused decline in Mr
which can be clearly seen from the test results of both QM and CM. Series 2 also
requires at least 12 f-t cycles to notice the f-t effects which rules out the possibility of
being considered as a potential laboratory test because of the duration. Compared with
series 3, series 4 showed more vulnerability to f-t cycles, hence, could be a better
choice.

As mentioned earlier, one of the main advantages of Mr is the ability to simulate


different stress combinations. In the previous section, RLT test results were mainly
discussed for selected four stress sequences 4,8,12 and 16. To study the effect of stress,
for each series of test, the absolute increase/decrease in Mr for these selected stresses
was calculated relative to 0 f-t cycle. In Figure 5.49, for the selected stress sequences,
the differences were stacked up on top of each other to get the cumulative differences.

215
Chapter 5 F-t, w-d durability study

Here, Mr reduction was considered as positive whereas the Mr increase was considered
as negative. The accumulated values did not show much difference between the stress
sequences. Only 12% difference was noticed between sequence 4 and 16, however, the
bulk stress of sequence 16 is more than three times of bulk stress of sequence 4
(θ4=422.3 kPa, θ16=1407.6 kPa). Therefore, it can be concluded that irrespective of the
level of stress, the absolute differences in Mr can be expected to be approximately same.
It must be recalled that the RLT test was considered to be superior to the UCS test
because of the ability to simulate different stress conditions. However the outcome of
this test may not act as a ‘sweetener’ to select rigours RLT Mr test opposed to simple
UCS test.

3200

2700
Accumulation of differences in

2200
resilient modulus (MPa)

1700

1200

700

200

-300
Sequence 4 Sequence 8 Sequence 12 Sequence 16

Figure 5.49: Accumulation of differences in Mr (MPa) of selected stress sequences

216
Chapter 5 F-t, w-d durability study

5.3.2.2.2 Effects of Freeze-Thaw on Repeated Load Triaxial Permanent


Deformation Test

The series of Mr tests in the previous section evaluated the effects of f-t on two different
granular materials in terms of the stiffness of the materials after different numbers of f-t
cycles. However, it did not assess the effects of f-t cycles on rutting behaviour which is
important for a stabilised base material. Therefore, on the completion of the Mr tests,
PD testing was carried out on the same specimens. Although using a new specimen for
each stress combination of repeated load PD testing is recommended in order to avoid
the effect of the stress history, the Mr test is non-destructive and only 1700 load cycles
were applied and majority of them are of very low magnitude. Moreover, the permanent
strain developed in the Mr test was found to be less than 1% and multiple non-
destructive tests on a specimen are not uncommon. For example, N and Zaman (2007)
carried out multiple Mr tests on the same specimen after different w-d cycles. Therefore,
to save time and make the test more practical, the same specimens were used for PD
testing.

The aim of this testing was to determine the PD responses of lightly stabilised materials
subjected to different numbers of f-t cycles varying with number of loading cycles and
stress levels to evaluate the impact of f-t cycles on rutting behaviour. The outcomes of
this experiment will be compared with the other mechanical properties presented earlier
in this chapter.

As explained in section 5.2.2.7, four different stress sequences, each with 10,000
loading cycles, were applied. The stress ratio of the maximum cyclic stress to bulk
stress was kept at approximately two for all tests. All the specimens tested for the Mr
were also tested for PD, with the permanent and resilient deformations of each cycle
calculated based on the initial and final readings from the LVDTs. Initial analyses of the
PD test results were carried out by comparing their absolute PD values. To begin the
analysis, the cumulative PDs varying with the number of load cycles were plotted, with
a typical one for QM in Series 1 shown in Figure 5.50. Although the first three stress
sequences did not produce significant amounts of PD, as the fourth produced
considerable amounts, they were extracted for comparisons of different specimens.

217
Chapter 5 F-t, w-d durability study

Figure 5.51 shows the PDs of QM specimens cured for 7 days and subjected to 0, 12
and 24 f-t cycles (Series 1) for the fourth stress sequence. All three tests shows the at
the PD increases rapidly for the first 1000 to 2000 load cycles before showing
significant reduction in the rate of accumulation. During the last 10,000 cycles, the
deformation of the specimen subjected to 24 f-t cycles was greater than those of the
other two. It is also important to note that the specimen subjected to 12 f-t cycles
deformed at only 0.19 mm compared with 0.47 mm for the 0 f-t specimen. It must be
recalled that similar results were found in both the UCS and Mr tests, in which the first
12 f-t cycles did not produce any sign of f-t effects.

However, the differences more revealing in the PD test. For example, specimen
subjected to 24 f-t cycles deformed as much as four times as the 0 f-t specimen.
Similarly, PDs from the test conducted on specimens moulded with CM are shown in
Figure 5.52. Compared with QM, CM did not deform much, as expected based on the
previous test outcomes (CM showed higher strengths and Mr and UCS than QM) and,
similar to QM, the f-t effect was noticeable after 24 f-t cycles while specimens
subjected to 12 f-t cycles showed less PDs.

1.6

0 f-t cycle
1.2 12 f-t cycles
Permanent deformation (mm)

24 f-t cycles

0.8

0.4

0 10000 20000 30000 40000


Number of load cycles

Figure 5.50: PDs of QM stabilised with 1.5% CF binder, cured for 7 days and
subjected to f-t procedure 1

218
Chapter 5 F-t, w-d durability study

The Series 2 specimens cured for 28 days were tested after 0, 12 and 24 f-t cycles, with
Figures 5.53 and 5.54 showing the PD responses for QM and CM respectively. Similar
to Series 1, QM deformed significantly higher than CM; for example, the 0 f-t
specimens of CM and QM deformed 0.20 mm and 0.71 mm respectively. Also, in this
series, 24 f-t cycles caused significant degradation, with the f-t effects noticeable even
after 12 f-t cycles, particularly for QM. The explanation for these behaviours was
discussed in details in the UCS test results.

As previously mentioned, the Series 3 and 4 tests were based on f-t procedure 2, which
were carried out on only CM. The Series 3 specimens were cured in a normal curing
environment whereas Series 4 at 400C. Moreover, in contrast to Series 1 and 2, these
two series were tested for up to 12 f-t cycles. The cumulative PDs for Series 3 and 4 are
shown in Figures 5.55 and 5.56 respectively. The f-t effects were noticeable even after 4
f-t cycles, in which deformations were more than 50% greater than those of 0 f-t cycles,
and the additional 8 f-t cycles generated even more adverse effects, particularly in series
3 specimens.

In summary, the accumulated PD captures the effects of f-t cycles similar to the
previous tests. However, the selection of deviator stress for a given confining pressure
has to be high enough to cause considerable PD. In this study, first three stress
combinations did not produce significant PD and was not used for the evaluation. It can
be recalled that, the deviator stress to confining pressure ratio of 6 was used for all four
stress sequence. Increasing this ratio, by either reducing the confining pressure or
increasing the deviator stress would have helped to produce significant permanent strain
even for other stress sequences. Moreover, PD tests also suggest the series 3 or 4 is a
better choice for the laboratory durability study.

219
Chapter 5 F-t, w-d durability study
Series 1_7days_ft1_QM
1.6 0 f-t cycle
12 f-t cycles
24 f-t cycles

1.2

Permanent deformation (mm)


0.8

0.4

0 2000 4000 6000 8000 10000


Number of load cycles

Figure 5.51: PD of QM stabilised with 1.5% CF binder, cured for 7 days and
subjected to f-t procedure 1

Series 1_7days_ft1_CM
1 0 f-t cycle
12 f-t cycles
24 f-t cycles

0.8
Permanent deformation (mm)

0.6

0.4

0.2

0 2000 4000 6000 8000 10000


Number of load cycles

Figure 5.52: PDs of CM stabilised with 1.5% CF binder, cured for 7 days and
subjected to f-t procedure 1

220
Chapter 5 F-t, w-d durability study

1.6
Series 2_28days_ft1_QM
0 f-t cycle
12 f-t cycles
24 f-t cycles
1.2

Permanent deformation (mm) 0.8

0.4

0 2000 4000 6000 8000 10000


Number of load cycles

Figure 5.53: PDs of QM stabilised with 1.5% CF binder, cured for 28 days and
subjected to f-t procedure 1

Series 2_28days_ft1_CM
0.4 0 f-t cycle
12 f-t cycles
24 f-t cycles

0.3
Permanent deformation (mm)

0.2

0.1

0 2000 4000 6000 8000 10000


Number of load cycles
Figure 5.54: PDs of CM stabilised with 1.5% CF binder, cured for 28 days and
subjected to f-t procedure 1

221
Chapter 5 F-t, w-d durability study
0
Series 4_40 C_ft2_CM
0.4 0 f-t cycle
4 f-t cycles
12 f-t cycles

0.3

Permanent deformation (mm)


0.2

0.1

0 2000 4000 6000 8000 10000


Number of load cycles

Figure 5.55: PD of CM stabilised with 1.5% CF binder, cured for 7 days at 40oC
and subjected to f-t procedure 2

Series 3_7days_ft2_CM
1 0 f-t cycle
4 f-t cycles
12 f-t cycles

0.8
Permanent deformation (mm)

0.6

0.4

0.2

0 2000 4000 6000 8000 10000


Number of load cycles

Figure 5.56: PD of CM stabilised with 1.5% CF binder, cured for 7 days and
subjected to f-t procedure 2

222
Chapter 5 F-t, w-d durability study

PD tests are time consuming and require a sophisticated testing facility as they are high
frequency and require thousands of cyclic loadings. Establishing models predicting PDs
are helpful to limit the number of tests and number of cycles. Modelling of PDs can be
performed with respect to the number of load cycles or related different stress
conditions. In addition shakedown theory can be used to analyse PD responses as well.

All the above PD data were correlated with two empirical equations, equations 5.7 and
5.8. On completion of a comprehensive study of a base course material, Barksdale
(1972) proposed equation 5.7. According to him, the accumulated PD can be linearly
related with the logarithm of the number of load cycles. Equation 5.8 was proposed by
Sweere (1990a) after analysing PD test results of 106 loading cycles applied on granular
materials. These two equations were fitted with the PD data obtained in this series of
experiments. Table 5.23 shows the regression coefficients and coefficient of
determination obtained for each curve fitting. The coefficient of determination shows
that logarithmic equation (eq 5.9) fits the data verify well and can be used to predict the
PD.

ep = a1ln(N)+b1 Eq. (5.9)

ep = a2Nb2 Eq. (5.10)

ep = Accumulated PD (mm)

N = Number of load cycles

a1, a2, b1,b2 = Regression parameters

223
Chapter 5 F-t, w-d durability study

Table 5.23: Regression parameters of two empirical equations correlating PD with


number of load cycles

ep = a1ln(N)+b1 ep = a2Nb2
a1 b1 R2 a2 b2 R2
0.1165 -0.2419 0.9829 0.1287 0.2066 0.9512
Series 1
0.0900 -0.1802 0.9961 0.0891 0.2212 0.8846
QM
0.1922 -0.4027 0.9931 0.1935 0.2173 0.9147
0.0824 -0.3067 0.9733 0.0291 0.3050 0.9799
Series 1
0.0294 -0.0857 0.9725 0.0214 0.2389 0.9899
CM
0.0971 -0.0383 0.9966 0.2105 0.1549 0.9615
0.115 -0.1909 0.9912 0.1321 0.2098 0.8677
Series 2
0.1380 -0.3155 0.9838 0.1237 0.2274 0.8860
QM
0.1972 -0.4919 0.9372 0.1790 0.2215 0.9316
0.0300 -0.0740 0.9945 0.0245 0.2349 0.9333
Series 2
0.0282 -0.0389 0.9974 0.0424 0.1826 0.9615
CM
0.0599 -0.0247 0.9703 0.0130 0.3518 0.9767
0.0824 -0.3067 0.9733 0.0291 0.305 0.9799
Series 3
0.0884 -0.0817 0.9363 0.0662 0.2733 0.4453
CM
0.115 -0.1909 0.9912 0.1321 0.2098 0.8677
0.0193 -0.0216 0.9917 0.0301 0.1826 0.9061
Series 4
0.0443 -0.1560 0.9433 0.0255 0.2671 0.9966
CM
0.0638 -0.2466 0.9678 0.0180 0.3281 0.9554

In the following section, PDs responses were further analysed using the shakedown
theory which better explains a material’s behaviour under cyclic loading by classifying
its responses in the following four categories as shown in Figure 5.57.

1. Purely elastic

This category implies that the material does not undergo PD, i.e., its total deformation is
recoverable. However, generally, for a pavement material, purely elastic conditions may
not be noticeable, even for a small load. Hence, purely elastic condition does not exist
for pavement materials.

2. Elastic shakedown

This is a condition whereby the material undergoes plastic deformations for a finite
number of cycles and ‘shakes down’ to a pure elastic response. Although the applied
load is sufficiently high to cause plastic deformation, it is less than the amount required

224
Chapter 5 F-t, w-d durability study

to produce plastic shakedown. The maximum load that can produce a shakedown elastic
response is defined as the ‘elastic shakedown limit’.

3. Plastic shakedown

At this stage, the material no longer operates under elastic deformation but shows a
hysteretic plastic response and accumulates plastic deformation until it reaches the state
in which cyclic plasticity occurs. The maximum load at which plastic shakedown is
achievable is termed the ‘plastic shakedown limit’.

4. Incremental collapse or ratcheting

This occurs when the applied load is considerably higher (higher than the plastic
shakedown limit) than the strength of the material, whereby permanent strain
accumulates much faster and failure occurs within a short period.

However, after analysing numerous RLT test data of unbound granular materials,
Werkmeister et al. (2005) stated that three different deformation responses (Ranges A,
B and C) could be observed under cyclic loading. According to him, range A behaviour
exhibits a small amount of permanent strain for the first few thousand load cycles and
then ‘shakes down’ to an elastic response. This behaviour is preferred in a pavement
because the rutting phenomenon can be limited to only the initial permanent strain.
Materials showing range B behaviour accumulate PDs with the number of load cycles
and fail due to rutting at a higher number of load cycles. This behaviour could be
allowable in pavements provided that the number of cycles required to cause rutting
failure is greater than the number of design cycles. In range C behaviour, resilient
deformation increases with the number of load cycles and fails at a lower number.

A better way of representing and classifying these three ranges of shakedown responses
is to plot graphs showing the permanent vertical strain rate versus vertical permanent
strain. Also, as these three ranges show different resilient responses which can be used
to classify materials’ responses, all the triaxial data acquired in this study were analysed
to obtain the vertical permanent strain rates and resilient responses, and then classified
in one of the three categories. In the following sections, typical responses obtained in
this study are discussed using graphical plots.

225
Chapter 5 F-t, w-d durability study

Figure 5.57: Elasto-plastic responses of material under repeated loading (Johnson


1986)

Range A_ Plastic Shakedown Range

Figure 5.58 shows a material’s typical responses in range A from a few selected tests.
The PD strain rate decreased drastically from a high to very small value (1×10-5) with
increases in the PD strain. The plastic response ceased after a certain number of cycles
and the deformation was entirely resilient. It can be seen in Figure 5.59 that the vertical
resilient response reduced to a stable value after a certain number of load cycles.
Therefore, it was not necessary to continue the test for greater numbers of load cycles
(105) for the selected stress value because the material responses would remain the
same. In the next sequence, applying a higher stress value would help to identify the
material response for that selected value which is the main advantage of multi-stage
over single-stage triaxial testing. Obtaining knowledge about a material’s behaviour by
looking at the data after 10,000 cycles helps to identify or predict the longer-term
responses of the material and avoids the need for long-term testing. Moreover, although
it is always desirable to have pavement materials operating in range A, this is not viable
because either the load has to be very small compared with the ultimate capacity of the

226
Chapter 5 F-t, w-d durability study

material or the thicknesses of the layers have to be increased to match the load which
could lead towards a conservative design.

0.01

0.001

Vertical Permanent Strain Rate (10-3/load cycle)


0.0001

1E-005

1E-006

1E-007

1E-008

1E-009

1E-010

0 0.2 0.4 0.6 0.8


Vertical Permanent Strain (10 -3)

Figure 5.58: Vertical permanent strain rates versus vertical permanent strains -
range A

0.0012

0.001
Verticle resilient strain

0.0008

0.0006

0.0004

0.0002

0 2000 4000 6000 8000 10000


Number of load cycles

Figure 5.59: Vertical resilient strain varying with number of load cycles - range A

227
Chapter 5 F-t, w-d durability study

Range C_ Incremental collapse range

Figure 5.60 shows typical responses under Range C obtained from few selected test.
Range C responses obtained in this series of tests are very less compared with other two
responses. The vertical permanent strain rate remains considerably higher in range C
after showing decline in rate in the first few thousands of load cycles. As shown in
Figure 5.60, the strain rate going towards a steady state which means for each load
cycle, constant amount of PD is getting developed within the material. It is important to
remember that none of the specimens tested were failed during the PD test because the
load cycles were limited to only 10,000 for each stress sequences. However, the
response clearly showed that the accumulation of PD at higher rate will make the
specimen fail. This is more evident from the resilient responses shown in Figure 5.61.
Resilient deformation increases with number of load cycles which implies that the Mr
reduces with number of load cycles. This is a sign of a range C behaviour where the
material is likely to fail prematurely.

0.1
Vertical Permanent Strain Rate (10-3/load cycle)

0.01

0.001

0.0001

1E-005

0 4 8 12
Vertical Permanent Strain (10 -3)

Figure 5.60: Vertical permanent strain rates versus vertical permanent strain_
range C

228
Chapter 5 F-t, w-d durability study

0.002

0.0016

Verticle resilient strain

0.0012

0.0008

30000 32000 34000 36000 38000 40000


Number of load cycles

Figure 5.61: Vertical resilient strain varying with number of load cycles_range C

Range B_ Intermediate Response

Figure 5.62 shows typical permanent deformation responses obtained for Range B
behaviour from few selected test. It showed considerable amount of permanent
deformation in the initial load cycles before declining to a lower values. Compared with
Range A and C behaviour, Range B behaviour showed intermediate response in which
the number of cycles to cause a failure is generally high.

229
Chapter 5 F-t, w-d durability study

0.1

Vertical Permanent Strain Rate (10-3/load cycle)


0.01

0.001

0.0001

0 4 8 12
Vertical Permanent Strain (10 -3)

Figure 5.62: Vertical permanent strain rates versus vertical permanent strain_
range B

The above classification of range A, B and C was clearly based on the shape/trend of
the permanent vertical strain rate versus vertical permanent strain graphs and resilient
responses. It is possible to assign a single-state stress to a particular category based on
the criteria proposed by Werkmeister et al. (2005) and EN_13286-7 (2004). The
shakedown limits published for unbound granular materials are based on their
accumulated permanent strains from 3000 to 5000 load cycles. The stress level is
considered to reach the plastic shakedown limits and plastic creep limits for strains of
0.045*10-3 and 0.4*10-3 respectively. It is important to define these limits to ensure that
a pavement operates at a stress level within the shakedown limits to avoid rutting failure
during its function.

All the test results were analysed in order to classify them in one of the ranges
mentioned above. The permanent vertical strain rate versus vertical permanent strain
behaviour, resilient strain as well as the numerical classification were used. Table 5.24
shows outcome of the classification. It can be seen that first three stress sequences were
either classified as range A or B and use them to classify or compare different test

230
Chapter 5 F-t, w-d durability study

results was not possible because of the lower values of PD. Classifying the PD
responses to one of the ranges is helpful to predict the material’s performance instead of
continuing the tests for longer period of time. Classification also helps to compare the
outcome between different tests.

Table 5.24: Classification of PD response

Number of
Series Material Sequence 1 Sequence 2 Sequence 3 Sequence 4
f-t cycles
0 A A/B B B/C
QM 12 A A B B
24 A A B C
Series 1
0 B B B B
CM 12 B B B B
24 B B B B
0 A A A B
QM 12 A A A C
24 A A B C
Series 2
0 A A A B
CM 12 B B B B
24 B B B B
0 A A A B
Series 3 CM 4 A A A B
12 A A A B
0 B B B B
Series 4 CM 4 A A A B
12 B B B C

5.3.2.2.3 Freeze-Thaw Effects on Residual Unconfined Compressive Strength

It is not uncommon to determine the RUCS at the end of RLT tests because of the non-
destructive natures of resilient and PD tests; for example, Sireesh Saride et al. (2014)
determined the ‘retained UCS’ of specimens tested using the RLT resilient modulus test
and compared the results with those from the UCS. As previously discussed the
shakedown analysis showed that, at the end of stage 4 load sequence, all the specimens
apart from a few, remained in range A/B. Also permanent strain of less than 1% was
observed at the end of the RLT and not a single specimen failed after repetitive

231
Chapter 5 F-t, w-d durability study

loadings. Moreover, the stress histories of all the specimens were similar because the
stress values and number of loading cycles were kept constant for all tests. Therefore, it
was decided to perform a monotonic test on the same specimens to determine the failure
load, with the objective of conducting a relative comparison of different series of tests.
The RUCS was determined at the end of the PD test using the same triaxial assembly.
The confining pressure during the monotonic test was kept at 0 kPa and the strain rate
of 0.5 mm per minute chosen was similar to that of the UCS for comparison purposes.

The failure loads and failure stresses obtained are shown in Table 5.25. Specimens of
CM and QM cured for 7 days and subjected to f-t procedure 1 (Series 1) achieved
higher RUCS after 12 f-t cycles than those subjected to 0 f-t cycles. However, a
significant strength decline was observed after 24 f-t cycles, similar to observations of
specimens subjected to similar curing and f-t cycles in the UCS test. However, as shown
in Figure 5.63, the adverse effect of f-t was greater in the triaxial than UCS specimens.
The percentage differences between the UCS and RUCS shown in Figure 5.63 were
determined relative to 0 f-t cycles. After the first 12 f-t cycles, the strength gains in the
RUCS were not as high as those in the UCS and the ultimate reductions after 24 f-t
cycles higher. In Series 1, the RUCS of triaxial specimens showed much greater
vulnerability to f-t than the UCS.

Figure 5.64 shows that Series 2 of QM cured for 28 days and subjected to f-t procedure
1 showed approximately 20% and 23% reductions in the RUCS after 12 and 24 f-t
cycles respectively and CM, 20% and 27% respectively. Figure 5.64 further illustrates
that the percentage decline in the UCS for QM was as high as 30% but only 8% for CM.

As shown in Figure 5.65, specimens cured at 400C (Series 3) lost 13% after 4 f-t cycles
and 23% after 12 f-t cycles. The series 3 UCS followed similar trend of RUCS,
however, UCS of Series 4 much lesser vulnerability to f-t cycles. Series 4 specimens
cured for 7 days and subjected to f-t procedure 2 did not show high strength reductions
after 4 f-t cycles but, after 12, the RUCS reduced by 22%.

All the UCS and RUCS values discussed in this section were empirically related, as
shown in Figure 5.66. The data points of specimens subjected and not subjected to f-t
cycles were separated from each other’s data sets and defined two different equations.

232
Chapter 5 F-t, w-d durability study

Regression coefficient suggests that the relationship obtained for specimens subjected to
f-t cycles is not convincing, however, the specimens not subjected to f-t cycles showed
strong correlation though sample size is small. Moreover, Table 5.25 shows that the
UCS was higher than the RUCS. One of the factors affecting the results from these two
compression tests could be differences in their aspect ratios (the ratio between the
height (h) and diameter (d)). It should be noted that, with an increasing h/d ratio, the
failure stress tended to decline. The h/d ratio was approximately 2 for the triaxial and 1
for the UCS specimens which explains why the RUCS ha to be less than the UCS. Apart
from the h/d ratio, the RUCS was obtained after Mr and PD tests which could have
reduced the ultimate capacities of the specimens.

Table 5.25: Residual UCS

Parent No. of Curing Number Failure Failure % Diff Failure


% Diff -
Series Material Days & F-t of F-t Load Stress - Stress -
UCS
& Binder Procedure Cycles (kN) (MPa) RUCS UCS (MPa)
0 13.2 1.7 0.0 1.73 0.0
CM 7-day
12 14.4 1.8 11.3 2.14 23.7
1.5% CF f-t procedure 1
Series 24 12.0 1.5 -14.5 1.91 10.4
1 0 12.4 1.6 0.0 1.71 0.0
QM 7-day
12 14.4 1.8 9.0 2.56 49.0
1.5% CF f-t procedure 1
24 10.6 1.3 -9.0 1.68 -1.8
0 17.1 2.2 0.0 2.29 0.0
CM 28-day
12 13.8 1.8 -19.8 2.22 -3.1
1.5% CF f-t procedure 1
Series 24 12.5 1.6 -22.8 2.10 -8.3
2 0 13.6 1.7 0.0 2.20 0.0
QM 28-day
12 10.9 1.4 -19.3 1.92 -12.7
1.5% CF f-t procedure 1
24 10.5 1.3 -26.9 1.49 -32.3
0 17.1 2.2 0.0 2.38 0.0
Series CM 7-day
4 14.8 1.9 -13.5 2.21 -7.1
3 1.5% CF f-t procedure 2
12 13.1 1.7 -23.4 1.80 -24.4
7-day 0 13.2 1.7 0.0 1.73 0.0
Series CM
f-t procedure 2 4 12.8 1.6 -3.0 1.77 2.3
4 1.5% CF 0
Curing at 40 C 12 10.3 1.3 -22.0 1.62 -6.4

233
Chapter 5 F-t, w-d durability study
60
7-day curing, f-t procedure 1
QM_RUCS
CM_RUCS
CM_UCS
QM_UCS
40

% Change in (R)UCS (MPa)


20

-20

0 5 10 15 20 25
Number of f-t cycles

Figure 5.63: Percentage declines in RUCS/UCS with number of f-t cycles of


specimens cured for 7 days and subjected to f-t procedure 1

28-day curing, f-t procedure 1


QM_UCS
CM_UCS
CM_RUCS
0
QM_RUCS

-10
% Change in RUCS/UCS

-20

-30

-40

0 5 10 15 20 25
Number of f-t cycles

Figure 5.64: Percentage declines in RUCS/UCS with number of f-t cycles of


specimens cured for 28 days and subjected to f-t procedure 1

234
Chapter 5 F-t, w-d durability study

10 Curing at 400C, f-t procedure 2_RUCS


Normal curing, f-t procedure 2_RUCS
Curing at 400C, f-t procedure 2_UCS
Normal curing, f-t procedure 2_UCS

0
UCS (MPa)

-10

-20

-30

0 4 8 12
Number of f-t cycles

Figure 5.65: Percentage decline in RUCS/UCS with number of f-t cycles


(procedure 2) of specimens cured at 400C and normally cured specimens

UCS vs RUCS
Not considererd for the
regression analysis Subjected to f-t cycles
2.8 Not subjected to f-t cycles

UCS=1.2106*RUCS
R2=0.6714
2.4 UCS=1.0601*RUCS
R2 =0.9732
AUCS (MPa)

1.6

1.2

1.2 1.6 2 2.4


UCS (MPa)

Figure 5.66: Relationship between RUCS and UCS

235
Chapter 5 F-t, w-d durability study

5.3.2.3 Freeze-Thaw Effects on Modulus of Rupture and Static Stiffness Modulus

The previous two sections evaluated the materials’ responses under compressive
loading. Although UCS and RLT tests on unbound materials are also common, the
flexural properties are unique only for stabilized materials. As studies focussing on the
environmental effects of the flexural properties of stabilised materials are limited, a
durability study of the flexural properties of QM was carried out.

As shown in table 5.26, triplicate beam specimens stabilized with two different binder
content of 1.5% and 3% CF were subjected to different numbers of f-t cycles based on
f-t procedure 1. All the tested specimens were cured for 28 days. Then the specimens
were tested using 3-point bending test following the experimental procedures described
in section 5.2.3. Both the loads and deformations were measured and then used to
calculate the flexural stress, flexural strain and flexural modulus.

Table 5.27 shows the outcomes of the flexural testing of specimens stabilised with a
1.5% CF binder. As previously mentioned the strength and stiffness values at 0 f-t
cycles represent those of specimens tested after 28 days of curing without being
subjected to f-t cycles and it measured MoR of 413 kPa. However, there was a sharp
decline in flexural strength from 413 kPa to 250 kPa (approximately 40%) after only a
single f-t cycle with further increases in the number of f-t cycles resulting in mild
decreases in the MoR and, after 12 f-t cycles, nearly 51% of strength loss was observed.
The measured MoR varying with the number of f-t cycles are shown in Figure 5.67.
Similar behaviour was noticed in the SSM as only 25% of the stiffness was retained
after a single f-t cycle, as shown in Figure 5.68 and Table 5.28, and after 12 f-t cycles,
there was 83% more stiffness reduction than after 0 f-t cycles. Sharp declines after the
first f-t cycle followed by mild reductions were also observed for samples stabilised
with 3% of CF binders, as shown in Figures 5.69 and 5.70, with one f-t cycle causing
65% and 87% reductions in the MoR and SSM respectively. However, a notable
observation from this study is that the specimens remained integrated even after 12 f-t
cycles.

236
Chapter 5 F-t, w-d durability study

Moreover, in situations where deflection data is not available, the relationship between
the flexural modulus and MoR is useful for calculating the flexural modulus which
requires sophisticated testing and a data acquisition system. The deflection
measurement arrangement used for this experiment produced consistent results, as can
be seen from the flexural stress-strain graph obtained for each individual specimen for
each f-t cycle (three tests in each) shown in Appendix A. The average stress-strain
graphs obtained for each f-t cycles are shown in Figures 5.71 and 5.72 for 1.5% and 3%
of CF binders respectively. It is evident that there were significant differences between
the specimens subjected and not subjected to f-t cycles. Moreover, Figure 5.73 and
equations 5.11 and 5.12 illustrate the relationships between the MoR and flexural
modulus of these specimens, with the former established in this study and the latter in
Paul (2012). As previously mentioned, these empirical relationships are useful for
determining one parameter if the other exists.

The results from the UCS and RLT Series 2 tests, in which the specimens were cured
for 28 days and subjected to f-t procedure 1, were compared with those from flexural
testing which showed that flexural testing exhibited greater vulnerability to f-t cycles.
Although, in the Series 2 tests, f-t cycles impaired properties such as the UCS, resilient
modulus, RUCS and PD, this never reached the extent observed in flexural testing. As
far as the series stabilised with 1.5% CF is concerned, the tensile capacity is very less
compared with compressive strength, hence the f-t cycles could have affected a much
heavier. However, specimens stabilised with 3% CF has tensile capacity nearly four
times that of 1.5% specimens also showed similar response. Therefore, it can be
concluded that f-t cycles affect tensile properties to a greater extent than other
compressive properties.

Table 5.26: Testing Program for flexural test

Granular Number of
Binder Curing Binder %
material cycles
QM CF 28 Days 1.5% 0,1,2,4,8,12

QM CF 28 Days 3% 0,1,2,4,8,12

237
Chapter 5 F-t, w-d durability study

Table 5.27: Variations in MoR and flexural modulus with number of f-t cycles for QM stabilised with 1.5% CF binder

Average Average
Binder Number of MoR Strain Average SSM
Material MoR CoV CoV SSM CoV
% f-t cycles (kPa) (10-3) strain (10-3) (MPa)
(kPa) (MPa)
433 157 2760
0 406 413 4 149 135 23 2725 2955 12
401 100 3379
286 311 921
1 243 243 18 361 331 8 672 739 22
201 321 625
225 278 810
2 228 230 3 217 283 24 1051 844 23
1.5% 238 354 672
QM
CF 222 363 612
4 213 227 8 289 337 12 736 677 9
247 360 684
204 418 487
8 207 202 3 383 392 6 540 514 5
194 375 516
164 226 726
12 192 187 11 262 252 9 734 742 3
205 268 766

238
Chapter 5 F-t, w-d durability study

Table 5.28: Variations in MoR and flexural modulus with number of f-t cycles for QM stabilised with 3% CF binder

Binder Number of Average Average Average


Material MoR CoV Strain CoV SSM CoV
% f-t Cycles MoR strain SSM
1842 110 16813
0 1654 1854 11 94 106 11 17641 17435 3
2066 116 17850
704 311 2262
1 679 656 10 275 299 10 2470 2203 14
585 312 1876
640 323 1981
2 674 687 8 392 361 8 1718 1910 9
CF 3% 746 367 2032
QM
CF 633 327 1937
4 629 627 1 322 329 1 1954 1910 3
620 338 1838
627 286 2189
8 520 590 10 304 304 10 1710 1945 12
622 321 1935
709 323 2193
12 700 697 2 398 371 2 1757 1900 13
682 390 1749

239
Chapter 5 F-t, w-d durability study

500

400
MoR of 1.5% Vs # of f-t cycles
MoR of Individual specimen

Modulus of Rupture (kPa)


Mean values of triplicate specimens

300

200

100

0
0 1 2 4 8 12
Number of f-t cycles

Figure 5.67: Variations in MoR with number of cycles for 1.5% CF binder

4000

SSM of 1.5% Vs # of f-t cycles


3000
MoR of Individual specimen
Static stiffness modulus (MPa)

Mean values of triplicate specimens

2000

1000

0
0 1 2 4 8 12
Number of f-t cycles

Figure 5.68: Variations in SSM with number of cycles for 1.5% CF binder

240
Chapter 5 F-t, w-d durability study

2500

2000
MoR of 3% Vs # of f-t cycles
MoR of Individual specimen
NModulus of rupture (kPa) Mean values of triplicate specimens

1500

1000

500

0
0 1 2 4 8 12
Number of f-t cycles

Figure 5.69: Variations in flexural modulus with number of cycles for 3% CF


binder

20000

16000
SSM of 3% Vs # of f-t cycles
MoR of Individual specimen
Static stiffness modulus (MPa)

Mean values of triplicate specimens

12000

8000

4000

0
0 1 2 4 8 12
Number of f-t cycles

Figure 5.70: Variations in MoR with number of cycles for 3% CF binder

241
Chapter 5 F-t, w-d durability study

0.4
0 f-t cycle
1 f-t cycle
2 f-t Cycles
4 f-t Cycles
8 f-t Cycles
0.3
12 f-t Cycles

Flexural Stress (MPa)

0.2

0.1

0 200 400 600 800 1000


Strain (10 -3)

Figure 5.71: Flexural stress-strain relationship of specimens stabilised with 1.5%


of CF binder and subjected to various f-t cycles

2
0 f-t cycle
1 f-t cycle
4 f-t cycles
1.6
4 f-t cycle
8 f-t cycles
12 f-t cycles
1.2
Flexural Stress (MPa)

0.8

0.4

-0.4

0 200 400 600 800 1000


Strain (10 -3)

Figure 5.72: Flexural stress-strain relationship of specimens stabilised with 3% of


CF binder and subjected to various f-t cycles

242
Chapter 5 F-t, w-d durability study

3000

2500

Flexural Modulus (MPa)


2000

1500

1000

500

0
0 0.2 0.4 0.6 0.8
Modulus of Rupture(MPa)

Figure 5.73: Relationship between flexural modulus and MoR of specimens


subjected to f-t cycles and specimens not subjected to f-t cycles

Relationship between flexural modulus (E) and MoR of samples subjected to f-t cycles,

E = 3101* MoR + 31 (R2 = 0.95) Eq. (5.11)

Relationship between flexural modulus (E) and MoR of samples not subjected to f-t
cycles,

E = 4840 * MoR (R2 = 0.96) Eq. (5.12)

243
Chapter 5 F-t, w-d durability study

5.3.3 Wet-Dry Durability Test using Different Element Testings

It’s widely understood that the mechanical properties of granular base materials are
sensitive to moisture fluctuations. Construction specifications recommend that these
materials be compacted at or near the OMC with attainment of maximum dry densities
(Qian 2010). Although, during construction, granular bases are compacted closer to the
OMC, the moisture content beneath the pavement changes and achieves an equilibrium
condition referred to as the equilibrium moisture content (EMC) (Perera Y.Y et al.
2004). Models for predicting the EMC of a granular base layer are based on material
properties as well as other environmental variables has helped the current M-E design
standard to incorporate moisture variations in the design (NCHRP 2004). However, the
EMC is still likely to fluctuate due to seasonal climatic changes and other factors, such
as f-t in cold regions, shallow water tables, improper drainage and porous asphalt layers,
and may not remain stable for the entire life of a pavement (Perera Y.Y et al. 2004).
Such moisture variations lead to w-d cycles which, in turn, have deleterious effects and
raise concerns regarding durability, particularly for stabilised materials. Also, stabilised
materials might experience leaching of their stabilisers as a result of moisture
movement. Therefore, durability studies and strength characterisations of stabilised
materials are carried out in laboratory mix designs.

The only relevant standard available for w-d durability testing is ASTM D 559 which
sets out the procedure for soil-cement mixes but has been criticised for its long testing
duration and ambiguity associated with its specified brushing portions. Moreover,
Khoury (2007) reported that Kalankamary and Donald (1963) considered that the
ASTM D 559 test method is too severe and not representative of field conditions.
Considering changes in the UCS as a criterion for assessing durability rather than
conducting durability testing is common among highway agencies (Zhang and Tao
2008). Although the UCS test is preferred because of its ease of implementation and
relatively short duration, its unconfining nature is criticised as not being representative
of field stress conditions. Therefore, using triaxial testing to evaluate the resilient
modulus (Mr) with different stress sequences is considered more relevant. Zaman et al.
(1999), Khoury and Zaman (2002) and Khoury (2007) presented series of studies using
RLT testing to evaluate the effects of w-d cycles on stabilised aggregates.

244
Chapter 5 F-t, w-d durability study

Moreover, literature studies revealed that there is little knowledge available regarding
the effects of w-d cycles on stabilised granular materials, with even fewer findings
published concerning lightly stabilised granular materials. Therefore, in this
investigation, UCS, RLT and flexural tests were conducted to evaluate the resistance of
a lightly stabilised material to w-d cycles using the parameters shown in Table 5.29.
Triplicate specimens were prepared for each cycle for UCS and flexural testing and a
single specimen was used for RLT testing. At the end of specific curing periods, each
specimen was subjected to different numbers of w-d cycles before being tested for its
UCS, triaxial and flexural properties. In the following section, the outcomes of the
laboratory tests are discussed in detail.

Table 5.29: Testing program for QM stabilised with CF binder subjected to w-d
cycles

Granular Numbers of w-d


Test Binder % Curing
Material cycles
QM UCS 1.5% CF 7 days 0,1,4,8,12

QM UCS 3% CF 7 days 0,1,4,8,12

QM RLT 1.5% CF 28 days 0,1,4,12

QM Flexural 1.5% CF 28 days 0,1,4,8,12

5.3.3.1 W-d Durability Study using UCS Test

W-d durability studies of QM were carried out on 105 mm diameter, 115 mm high
cylindrical specimens stabilised with CF binder according to the preparation and testing
procedures explained in section 5.2.1. The specimens were sheared using monotonic
axial loading at the rate of 0.5mm/min to obtain the UCS, stabilised with two binder
contents, 1.5% and 3%, and cured for 7 days in a moist curing room kept at a
temperature of 210C and relative humidity of 95%.

Table 5.30 shows the measured UCS of specimens stabilised with 1.5% and 3% CF
binders subjected to various numbers of w-d cycles. Those specimens stabilised with a
1.5% CF binder became badly damaged and could not be tested beyond 4 w-d cycles.

245
Chapter 5 F-t, w-d durability study

Figures 5.74 and 5.75 show the specimens stabilised with a 1.5% CF binder after 1 w-d
and 8 w-d cycles respectively, with the extent of damage clearly visible after 8 w-d
cycles. However, increasing the CF binder content to 3% helped the stabilised material
retain its strength and shape even after 12 w-d cycles. The images of 3% stabilised
specimens subjected to 0 and 12 w-d cycles are shown in Figures 5.76 and 5.77
respectively. After 12 w-d cycles, no significant damage was suffered by the specimens
and, more importantly, their UCS remained considerably high.

As shown in Table 5.30 and Figure 5.78 and 5.79, after 1 w-d cycle, the UCS increased
significantly over that of 0 w-d cycle (0 w-d cycle represent the specimens which did
not undergo w-d cycles but were tested at the end of curing). The percentage increase
was 66% and 68% for the 1.5% and 3% CF binder respectively. However, specimens
stabilised with 1.5% CF deteriorated markedly after the first w-d cycle and, after 4 w-d
cycles, showed only 0.65 MPa. The specimens stabilised with 3% CF also displayed
some evidence of the destructive nature of w-d cycles but retained their shapes even
after 12 w-d cycles when their UCS remained significantly higher than after 0 w-d
cycles.

Figure 5.74: UCS specimens after 1 w-d cycle curing (1.5% CF binder - QM)

246
Chapter 5 F-t, w-d durability study

Figure 5.75: UCS specimens after 8 w-d cycles (1.5% CF binder)

Figure 5.76: UCS specimens at end of curing (3% CF binder)

247
Chapter 5 F-t, w-d durability study

Figure 5.77: UCS specimens after 12 w-d cycles (3% CF binder)

248
Chapter 5 F-t, w-d durability study

Table 5.30: UCS obtained after different numbers of w-d cycles

Average
Number of UCS
Material Binder % UCS CoV
w-d cycles (MPa)
(MPa)
0 1.66
0 1.73 1.71 2.3
0 1.73
1 2.80
1 2.66 2.84 7.5
1 3.08
1.5% CF
4 0.37
4 0.91 0.65 41.9
4 0.66
12 0
12 0 _ _
12 0
QM
0 4.18
0 4.09 4.38 9.6
0 4.86
1 6.29
1 7.59 6.94 9.4
1 6.93
3% CF
4 6.62
4 6.57 6.30 2.7
4 6.30
12 6.75
12 5.88 6.08 9.7
12 5.62

249
Chapter 5 F-t, w-d durability study

2
UCS (MPa)

0 4 8 12
Number of w-d cycles

Figure 5.78: UCS of specimens subjected to w-d cycles and normally cured
specimens with 1.5% CF binders

6
UCS (MPa)

0 4 8 12
Number of w-d cycles

Figure 5.79: UCS of specimens subjected to w-d cycles and normally cured
specimens with 3% CF binders

250
Chapter 5 F-t, w-d durability study

The occurrence of w-d reversals in a pavement is expected to weaken the materials’


mechanical properties. However, specimens stabilised with 1.5% and 3% CF binders
showed sudden increases in their UCS after 1 w-d cycle. Moreover, the UCS of
specimens stabilised with 3% CF binder remained considerably higher after 12 w-d
cycles than with 0 w-d cycles. Different factors contributed to this strength gain which
overcame the adverse effects of w-d in the early cycles, with one possible cause the rate
of change in the hydration/pozzolanic reaction rate.

The rate of strength development of a stabilised material is directly proportional to the


rate of the hydration or pozzolanic reaction. In the early days of curing, due to the high
rate of cementitous reaction, the stabilised material gains strength quickly but the rate of
its increase in UCS reduces with the age (number of cycles) of the specimen, as can be
seen in Table 5.31. However, specimens subjected to 1 w-d cycle showed significant
increase in strength compared with those cured in the curing room for the same number
of days, for example, a 7-day cured specimen subjected to 1 w-d cycle showed an
increase of 66% whereas a sample cured in the curing room for 9 days (2 additional
days to match the 1 w-d cycle) showed only a 14% increase which is possibly due to the
rise in temperature in the w-d cycle. As previously mentioned, the specimens subjected
to w-d cycles spent 7 days in the curing room at a temperature and relative humidity of
21°C and 95% respectively before exposure to 5 hours of soaking in water and drying at
710C for 42 hours. The increase in temperature to 710C could have favoured the early
cementitious reaction and helped the un-reacted binder increase the rate of reaction. As
shown in Table 5.31, specimens cured for 7 days attained only 74% of the 28-day
strength of other specimens which means that a significant amount of un-reacted
cementitious material was available at the end of 7 days. It appears that changing the
environmental conditions in a w-d cycle benefited the binder reaction and enabled the
gain of early strength, particularly in specimens cured for fewer days. To further
illustrate this concept, specimens cured for different numbers of days were tested after 1
w-d cycle and results compared with those after 0 w-d cycles, as shown in Table 5.31.

This illustrates the impact on the strength gain of different numbers of days of curing on
w-d cycle. Samples cured for 7 days stabilised with a 1.5% CF binder gained 66% after
1 w-d cycle while the 28-day and 56-day cured samples showed less than an 11%

251
Chapter 5 F-t, w-d durability study

increase. Also, samples stabilised with a 3% CF binder and cured for 7 and 196 days
showed approximately 68% and 16% strength increases. These observations indicate
that samples cured for fewer days have higher percentage of un-reacted cementitious
products which is considerably accelerated by a single w-d cycle. However, samples
cured for a higher number of days have a lower percentage of un-reacted cementitious
products which is accelerated marginally by a single w-d cycle. Moreover, a sample
stabilised with a 1.5% CF binder and cured for 7 days showed a higher UCS than those
cured for 28 and 56 days. Therefore, it can be concluded that, depending on the
availability of un-reacted binders, increases in the UCS vary.

Khoury and Zaman (2006) also reported that samples cured for 28 days showed increase
in the resilient modulus up to 12 cycles before showing reduction after 30 cycles while a
specimen cured for 3 days did not show any reduction for up to 30 cycles. Therefore,
when an increase in strength due to cementitous reactions becomes insignificant or less
sufficient, the deleterious effects of f-t or/and w-d cycles become noticeable.

The UCS durability test results indicate that the stabilisation provided by 1.5% CF
binder was not adequate to withstand w-d cycles. The specimens stabilised with 1.5%
CF binder could not be tested beyond 4 w-d cycles because of heavy deterioration.
However, increasing the binder content to 3% showed enough resistance to w-d even
after 12 w-d cycles. Furthermore, it had been found that the selected environmental
parameter of w-d cycles accelerates the hydration and pozzalanic reaction which
overcome the effects of w-d cycles in the short term.

252
Chapter 5 F-t, w-d durability study

Table 5.31: Specimens subjected to 1 w-d cycle and cured for different numbers of
days

1 W-d %
Binder Number of 0 Cycles -
Cycle - UCS Difference
Content Curing Days UCS (MPa)
(MPa)

7 1.71 2.84 66.1

1.5% 28 2.2 2.34 6.4

56 2.30 2.54 10.4

7 4.14 6.94 67.6


3%
196 6.37 7.42 16.5

5.3.3.2 W-d Durability Study using RLT Test

As well as the UCS test, the RLT test was performed on specimens of QM subjected to
w-d cycles. Unlike the UCS test, these specimens were cured for 28 days and had only
one binder content (1.5% CF) because of the heavy deterioration noticed in the UCS test
for 7-day cured specimens and their inability to complete 12 w-d cycles. It was expected
that the increased strength could provide more resistance to w-d cycles. Figures 5.80
and 5.81 show images of specimens after different numbers of w-d cycles in which
heavy deterioration was observed after both 2 and 4 w-d cycles. The specimens
subjected to 4 or more w-d cycles were not tested, because they lost their shapes, their
installation did not appear to be practical. Similar to the f-t study, a resilient modulus
test was first carried out and then a PD test conducted on the same specimen. The
sampling and testing procedures were similar to those for the f-t tests discussed in
section 5.2.2. The results from these tests are discussed in the next section.

Resilient Modulus (Mr)

The Mr obtained for specimens subjected to 0, 1 and 2 w-d cycles as well as the
unbound granular material are shown in Table 5.32, with that of each stress sequence
the average of the last five loading cycles. Similar to f-t study, to make this comparison,

253
Chapter 5 F-t, w-d durability study

the Mr was plotted against the bulk stress, as shown in Figure 5.82. Irrespective of the
bulk stress, the specimen subjected to 1 w-d cycle showed higher Mr than that subjected
to 2 w-d cycles while 0 w-d cycle specimen showed less sensitivity to the bulk stress
than those subjected to w-d cycles. At lower stresses, the Mr of 0 w-d cycle specimen
was higher than those of the specimens subjected to 1 w-d and 2 w-d cycles but, at
higher stress levels, were lower.

As shown in Figure 5.82, the bulk stresses of the tests differed by a small margin largely
because of variations in the deviatoric stress (the confining stresses were controlled
within +1 kPa). To make the comparison of different specimens more meaningful, the
basic k-θ model was used to obtain an equation. As shown in Figure 5.82, a good
correlation was observed between the Mr and bulk stress. The regression equation was
then used to determine the Mr for different stress sequences, with typical calculated
variations in the Mr with the number of w-d cycles shown in Table 5.33 and Figure 5.83
for stress sequences 4, 8, 12 and 16. Except for the fourth stress sequence, increase in
Mr can be seen at the end of 1 w-d cycle than 0 w-d cycle. However, after 2 w-d cycles,
the specimens showed significant loss in Mr. Although the Mr were not tested beyond 2
w-d cycles, the nature of the specimens after 4 w-d cycles indicated that they were
likely to decrease.

254
Chapter 5 F-t, w-d durability study

Figure 5.80: Triaxial specimens after 0 and 1 w-d cycles

Figure 5.81: Triaxial specimens after 2 and 4 w-d cycles

255
Chapter 5 F-t, w-d durability study

Table 5.32: Resilient modulus of stabilised specimens with different numbers of w-


d cycles

W-D Unbound
Stress
0 Cycle 1 Cycle 2 Cycles Material
Sequence
(MPa) (MPa) (MPa) (MPa)
1 499 395 350 365
2 545 417 374 375
3 649 487 442 446
4 628 496 438 444
5 565 513 424 418
6 632 553 481 443
7 657 611 525 481
8 596 636 563 495
9 537 615 510 409
10 583 666 570 469
11 668 769 679 553
12 673 768 688 572
13 636 759 638 484
14 706 838 729 564
15 756 895 804 599
16 784 912 894 624

256
Chapter 5 F-t, w-d durability study

1000

0 w-d Cycle
1 w-d Cycle
2 w-d Cycles
800
Resilient Modulus (MPa)

600

400

200

250 500 750 1000 1250 1500


Bulk Stress (kPa)

Figure 5.82: Resilient modulus varying with bulk stress of specimens subjected to
w-d cycles

Table 5.33: Calculated Mr based on regression equation

Mr (MPa) Mr (MPa) Mr (MPa) Mr (MPa)


Number of Regression Load Load Load Load
W-d Cycles Parameters Sequence 4 Sequence 8 Sequence 12 Sequence 16
θ =422.3 kPa θ =703.8 kPa θ =1055.7 kPa θ =1407.6 kPa
k1=177.88
0 w-d k2=0.1977 588 650 705 746
R2=0.67
k1=25.278
1 w-d k2=0.5026 528 682 836 966
R2=0.96
k1=17.937
2 w-d k2=0.0.5333 451 592 735 857
R2=0.97

257
Chapter 5 F-t, w-d durability study

Stress sequence 4
1000
Stress sequence 8
Stress sequence 12
Stress sequence 16

800
Resilient modulus (MPa)

600

400

200
0 1 2 3 4
Number of w-d cycles

Figure 5.83: Resilient modulus varying with number of w-d cycles for typical stress
sequences 4, 8, 12 and 16

Permanent Deformation

The multi-stage PD test carried out after Mr test showed clear evidence of deterioration
as a result of w-d cycles. Similar to the f-t test, first three stress sequences did not
produce significant amount of PD and last 10,000 loading cycles showed a rapid rate of
PD accumulation as shown in Figure 5.84. Two w-d cycles weakened the specimen
substantially and caused significant permanent strain with the increase in number of
load cycles.

As discussed in the f-t permanent deformation tests, to analyse the PD experimental


results further, categorisation presented by Werkmeister et al. (2005) was used since
lightly stabilised granular materials are not ‘fully-bound’ and their behaviour are close
to those of unbound granular materials (Austroads 2012). A better representation to
classify the ranges of shakedown responses is to plot permanent vertical strain per load
cycle (hereafter will be referred as ‘rate of strain’) against vertical permanent strain.
Figure 5.85 shows permanent vertical strain rate versus vertical permanent strain of

258
Chapter 5 F-t, w-d durability study

specimen subjected to 2 w-d cycles from the multi stage cyclic load testing (Note:
accumulated vertical permanent strain for the four stress sequences/stages are depicted
as I, II, III and IV in Figure 5.85). First three stages (I, II, III) of stress sequences
produced only a small amount of permanent strain. During these stages, rate of strain
increase declined to very lower values showing that the PD has almost ceased. Decline
in rate of strain also emphasises the fact that the PD only occurs in the initial stages of
the loading before achieving steady state. Similar observation was made in the 0 w-d
and 1 w-d specimens as well. According to Werkmeister et al. (2005), this type of
behaviour is classified as range A behaviour.

However, the last loading stage (stage IV) resulted in the sample experiencing range C
behaviour and it clearly demonstrates the w-d effect on the material. As shown in Figure
5.86, comparing the results of specimens subjected to different number of w-d cycles, it
is evident that samples subjected to 0 w-d cycle and 1 w-d cycle reached a steady state
whereas 2 w-d cycles sample showed Range C behaviour.

Compared with Mr test, the PD test showed clear w-d effects at the last load sequence.
In the Mr test, the reduction in resilient deformation was not clearly noticeable for 1 w-d
and 2 w-d specimens. However, PD test showed that at higher stresses the effects of w-d
cycles can be observed clearly.

Figure 5.84: Permanent deformation with different numbers of w-d cycles

259
Chapter 5 F-t, w-d durability study

0.1

Vertical Permanent Strain Rate (10-3/load cycle)


0.01

IV
0.001

III

0.0001

II

1E-005

0 10 20 30
Vertical Permanent Strain (10 -3)

Figure 5.85: Permanent strain rates versus vertical permanent of specimen


subjected to 2 w-d cycles

0.1

0 Cycle
Vertical permanent strain rate (10-3/cycle)

1 w-d cycle
2 w-d cycles

0.01

0.001

0.0001

0 10 20 30
Vertical permanent strain (10 -3)

Figure 5.86: Permanent strain rates versus vertical permanent of loading sequence
4

260
Chapter 5 F-t, w-d durability study

5.3.3.3 Wet-Dry Durability Test using Flexural Test

Finally, a set of beam specimens of QM stabilised with 1.5% CF binders were prepared
for a w-d durability study. Similar to f-t durability testing, 3-point monotonic loading at
a rate of 0.5 mm/min was carried out with load and deflection measurements. Figure
5.87 and Table 5.34 show the MoR, flexural strain and SSM determined from the tests
which, unlike those from the UCS and RLT tests, showed significant decline after 1 w-d
cycle and continued to reduce for the next 8 w-d cycles but, similarly, did not last for 12
w-d cycles. As shown in Figure 5.88, the specimens subjected to 4 w-d cycles retained
their shapes, similar to those subjected to 0 and 1 w-d cycles, which helped
measurements of their deflections. However, for specimens subjected to 8 w-d cycles,
as the deflection measurement setup was not able to be installed because of the rough
surface, the strain and SSM were not able to be calculated and the specimens could not
be tested beyond 8 w-d cycles due to their heavy deterioration.

Table 5.34: MoR and SSM for different numbers of w-d cycles

Number Average
Strain Average
of w-d MoR Average CoV strain CoV SSM CoV
(10-3) -3 SSM
cycles (10 )
0 433 157 2760
0 406 413 4.1 149 141 14.3 2725 2955 12.5
0 401 119 3379
1 192 220 872
195 2.2 223 1.3 877 0.8
1 198 225 882
4 176 459 384
173 2.5 482 6.7 360 9.2
4 170 505 337
8 148
155 5.9 - - - - - -
8 161

261
Chapter 5 F-t, w-d durability study

500

400
MoR Vs # of w-d cycles
MoR of Individual specimen
Mean values

300

200

100

0
0 1 4 8 12
Number of w-d cycles

Figure 5.87: MoR for different numbers of w-d cycles

Figure 5.88: Specimen subjected to 4 w-d cycles and the specimen subjected to 8 w-
d cycles is being tested for failure load only

It is important to understand the nature of the three different tests used, the UCS, RLT
PD and Mr. The UCS is based on the ultimate failure of a specimen whereas the PD test
investigates the rutting phenomenon. Basically, both these tests characterise the long-

262
Chapter 5 F-t, w-d durability study

term performance of a material whereas, the Mr test, which is limited by the number of
loading cycles and deviatoric stresses, focusses on a material’s short-term behaviour. A
comparison of results showed that, in contrast to the Mr test, the PD and UCS tests
exhibited similar trends and revealed clear distinctions in behaviour for different
numbers of w-d cycles. At the end of the w-d cycles, the measured moisture contents of
all the samples, irrespective of the number of loading cycles to which they were
subjected, ranged between 2% and 2.5%, and they became more brittle in nature. As
such brittleness could have concealed the detrimental effects of w-d cycles in the Mr
test, in a durability study, it is important to select suitable mechanical properties and
tests to assess the destructive nature of w-d cycles.

As previously noted, this experimental study could not be continued beyond 4 w-d
cycles, except for the UCS test conducted with a 3% CF binder, because the specimens
stabilised with a 1.5% CF binder were badly damaged after 4 w-d cycles which
indicated that stabilisation using a 1.5% CF binder is inadequate in terms of durability
against w-d cycles. On the other hand, it’s likely that the environmental parameters
(wetting and drying phases) selected were much more severe for lightly stabilised
granular materials; for example, Khoury and Zaman (2002) adopted 24-hour wetting
and drying phases which was different from the recommendations in ASTM D559
(2007). Therefore, further study using different environmental parameters is essential
before arriving at any conclusion regarding the ‘weakness’ of a granular material
stabilised with a 1.5% CF binder and subjected to w-d cycles.

Although it is true that w-d cycles have detrimental effects, this may not be noticeable
in the early stages. For example, in this study, 1 w-d cycle specimen performed better
than 0 w-d cycle specimen in terms of the UCS and PD tests and specimens stabilised
with 3% CF binders showed 39% higher UCS after 12 w-d cycles than 0 w-d cycle. A
possible explanation for such behaviour could be the accelerated cementitious and
pozzolanic reactions during the w-d cycles as the hydration of cement and pozzolanic
reaction of flyash are sensitive to temperature. Accelerated hydration due to a high
drying temperature (71oC) could be a reason for the high strength gain. The effect of
hydration becomes less prominent if a sample subjected to w-d cycles is cured for a
longer time or cured in an accelerated curing environment. Therefore, it is important to

263
Chapter 5 F-t, w-d durability study

select an appropriate curing time and environmental parameters for w-d durability
testing so that the reductions in a material’s mechanical properties can be observed
within a certain number of w-d cycles.

5.4 Summary

This chapter presented an investigation on durability of lightly stabilised material


against f-t and w-d cycles using UCS, RLT and flexural testings which tested different
mechanical properties. Monotonic loading was selected for the UCS and flexural testing
whilst cyclic loading was chosen for the RLT test. The RLT testing studied the resilient,
PD behaviours and RUCS and flexural test determined MoR and SSM. Variables such
as type of granular material, curing days, binder percentage, curing methods and type of
f-t procedures were added to the investigation for evaluating their impacts on durability.

The f-t durability test using UCS test was categorised into four series based on the
variables and f-t procedures. All four series of UCS test showed that the f-t cycles had
an effect on UCS, however depending on the combinations of the variables and f-t
procedure, the effects of f-t cycles varied. The specimens subjected f-t procedure 2,
showed greater vulnerability to f-t cycles than f-t procedure 1 and specimens cured for
less number of days needed higher number of f-t cycles to show declining in UCS.
Therefore, based on this study, for a lightly stabilised granular material, f-t procedure 2
is a better choice than f-t procedure 1 because: firstly, it better simulates field
conditions; and, secondly, the deleterious effect is noticeable even after 4 f-t cycles
which is important in a laboratory study where the timeframe has to be reasonably short.
Also, its experimental procedure is easy to implement.

The preliminary investigation of RLT resilient modulus test revealed that the small
stress values suggested in the Mr standard needed to be eliminated. It was found that the
small stress values were not only difficult to control but also caused inconsistent
measurements, particularly for a lightly stabilized granular material. The stress
combinations with 16 stress values were chosen for the RTL resilient modulus test.

The RLT resilient modulus test results were expressed using k- model and found that
the data showed good correlation. There were many similarities in the outcome of this

264
Chapter 5 F-t, w-d durability study

Mr test results compared with UCS test results. The effect of f-t procedure was very
much visible in the Mr results as well. The influence of strength gain, number of curing
days, curing method and f-t procedure were also noticed in the Mr test. Series 1 and 2
showed that 12 f-t cycles were not enough to cause any major deterioration in the
specimens, but f-t procedure 2 showed it is more destructive than f-t procedure 1. One
of the main advantages of Mr is the ability to simulate different stress combinations, but
it was found that irrespective of the level of stress, the absolute differences in Mr were
approximately same. Although, RLT PD test at lower stress did not reveal the effects of
f-t cycles, at higher deviatoric stress, the effects were clearly visible and it is a better test
method than RLT resilient modulus test.

F-t durability study performed on flexural properties, MoR and SSM, showed that f-t
effects was much higher and significant reduction was observed even after 1 f-t cycle
due to very limited tensile capacity.

One of the objectives of the study was to find out a better durability test method to
evaluate the resistance of lightly stabilised granular material against f-t cycles.
Compared with clay or silt content, the selected granular materials had high proportion
of larger particles and pore sizes which meant that effects of f-t were minimal.
Therefore when selecting testing procedures, there were factors and variables needed
greater concern which may go un-noticed in finer materials. As the f-t effects are
minimal, the testing procedures starting from sample preparation till the testing needed
greater precision. It can be recalled that the RLT resilient modulus test eliminated lower
deviatoric stress values to improve the consistency of the results.

It was found that there were two factors needed attention in the f-t durability test; one is
number of days of curing and the other one is the moisture content. The UCS and RLT
tests revealed that the curing age has an influence on the properties obtained after f-t
cycles. Stabilised material continuously gained strength during the f-t cycles and in
early days of its curing, due to the high cementitious reaction rate, it overcame
deleterious f-t effects and did not show or decline in the earlier f-t cycles. When rate of
strength development became insignificant or less sufficient, the deleterious effects of f-
t were noticeable. Therefore, a better approach is to cure in an accelerated environment
and minimize the effects of strength gain.

265
Chapter 5 F-t, w-d durability study

Supplying moisture during the thawing phase was shown to be important and it revealed
the role moisture content plays in the f-t cycles. F-t procedure 2 supplied moisture
during the thaw phase and generated an adverse effect on the material and it was
apparent that the f-t procedure 2 had more destructive effects than f-t procedure 1.
Moreover, f-t procedure 2 more closely simulated field conditions than f-t procedure 1.
For a laboratory test, it is important that the selected f-t procedure should be able to
create an adverse effect on the material within a reasonable time period. Therefore f-t
procedure 2 with accelerated curing is a better choice for a laboratory test which is
Series 3 in this study.

Although a laboratory f-t durability test of a stabilised material may not necessarily
require test on its mechanical properties, it ultimately affects the properties and a better
approach is to test its mechanical properties than using an indirect test procedure; for
example tube suction test. In this study different mechanical properties were tested, but
it was clear that the tensile tests can’t be used because of the limited tensile capacity of
lightly stabilised material.

UCS is a simple and easy test which produced reliable results and the UCS of triplicate
specimens showed greater consistency as measured by coefficient of variance (CoV).
Easy to test the influence of different variables and UCS has wider acceptability in the
industry hence should not be avoided.

Compared with RTL resilient modulus test, PD test clearly identified the f-t effects
when the specimens subjected to higher deviatoric stresses. Since PD test was
performed for 40,000 cycles, the rutting behaviour was better revealed which is more
critical for a lightly stabilised material. It is also important to note that it would have
been very difficult to design this RLT test in the absence of UCS test results. Therefore,
UCS is a good beginning for this lightly stabilised material which can be succeeded by
RLT PD test to assess the durability against f-t cycles.

Similar to f-t durability tests, series of tests were performed to assess a lightly stabilised
granular material’s resistance against w-d cycles focussed on same mechanical
properties. One common observation from all three different testings was that
specimens stabilised with 1.5% CF binder showed heavy deterioration after few w-d

266
Chapter 5 F-t, w-d durability study

cycles. In the UCS test, after 1 w-d cycle, the UCS increased significantly over that of 0
w-d cycle and after 4 w-d cycles deteriorated markedly to 0.65 MPa, but increasing CF
binder content to 3% showed enough resistance to w-d even after 12 w-d cycles.
Therefore, the UCS w-d durability test results indicated that the stabilisation provided
by 1.5% CF binder was not adequate to withstand w-d cycles.

Also, the RLT tests could not be continued beyond 2 w-d cycles as the specimen
damaged significantly. Similar to f-t study, to make the comparison, the Mr was plotted
against the bulk stress and all the test results showed good correlations. Although
shakedown analysis of RLT PD tests revealed that w-d cycles have less influence on
lower deviatoric stress, higher deviatoric led to plastic creep of specimens subjected to
w-d cycles. Because of the limited tensile capacity, the MoR and SSM were found to be
more vulnerable to w-d cycles than other properties matching the f-t test results.

The dominance of cementitious reaction over the deleterious w-d was very much visible
in samples cured for less number of days (7 days). The accelerated reaction of unreacted
binders caused by favourable w-d environmental conditions (i.e., accelerated
temperature) masked the adverse w-d effect in the first cycle. Therefore, it is also
possible that the accelerated strength gain in the earlier stages of w-d cycles could
outweigh the detrimental effects of w-d cycles.

UCS and RLT PD tests at higher stress distinctively showed the effects of w-d cycles
than RLT resilient modulus test. As mentioned earlier, UCS is a simple test and the use
of RLT PD along with it is a better approach for evaluating the w-d durability of lightly
stabilised granular materials than Mr test. Because of heavy deterioration obtained in
this series of test, environmental parameters of w-d cycles to assess the lightly stabilised
material needs further study and the appropriate values has to be selected.

In summary, durability studies against f-t and w-d cycles of a lightly stabilised material
are essential and the selected two different granular materials showed vulnerability and
all the mechanical properties tested had impaired and degraded. Different variables,
durability procedures and element testings suggested that particular few combinations
performed better and can be used to assess the durability of lightly stabilised material. It
is important to remember that the all the specimens tested were prepared and tested

267
Chapter 5 F-t, w-d durability study

under laboratory environment which could differ considerably in the actual field
scenarios. Its applicability to filed conditions needs to be verified with field data for
better interpretation.

268
Chapter 6 Permanent Deformation Behaviour using RLT

6 Chapter 6

Characterising Permanent Deformation Behaviour `of Lightly


Stabilised Granular Material using Repeated Load Triaxle
(RLT) Test

6.1 Introduction

As discussed earlier, the lightly stabilised granular base material is an important


structural element in a pavement structure. The rutting of base and subgrade layers
cause progressive fatigue cracking of bituminous layers which ultimately leads towards
failure.

The permanent deformation (PD) of unbound granular material is modelled using


analytical models, plasticity theory based models and shakedown theory. Analytical
model is based on a constitutive relationship which can be used to predict the PD for a
particular stress and loading cycle. The approach based on the plasticity theory is
modelled using elasto-plastic models that use incremental deformation. The shakedown
theory is based on the concept that the deformation response stabilises with the number
of loading cycles given the applied stress is low compared with the strength of the
material.

In this study, PD of lightly stabilised material under repeated loading is presented using
series of repeated load triaxial (RLT) test. Initially, the material responses are classified
based on the shakedown theory and the shakedown limits are determined. Analysis of
resilient deformation and influence of moisture on resilient modulus are also presented.
Last part of the chapter is dedicated to the development of the permanent deformation
behaviour model.
Chapter 6 Permanent Deformation Behaviour using RLT

6.2 Experimental Methodology

6.2.1 Granular Materials and Binders

The multi-stage RLT tests were performed on Canberra material (CM) and the flexural
and UCS tests on Queensland material (QM). General blended (GB) cement and fly ash
in a ratio of 75:25 were used as the stabiliser. The physical properties of these two
granular materials and general descriptions of the binders were presented in Chapter 3.
As shown in Table 6.1, multi-stage RLT testing was performed on specimens stabilised
with 0.5% and 1.5% binder contents, and UCS and flexural tests on specimens
stabilised with 1.5% binder contents.

To evaluate the moisture effects using RLT tests, the moulding moisture contents were
changed during compaction. However, for the UCS and flexural tests, specimens were
compacted at OMC and dried in an oven at 600C for different durations to obtain
different moisture contents. Both these procedures for studying moisture effects have
been reported in the literature (Ba. et al. 2012; Khoury and Zaman 2004; Li et al. 2010;
Nazzal and Mohammad 2010; Yuan and Nazarian 2003)

Another approach for changing the moisture content in a triaxial specimen is to


supply/extract moisture to/from the bottom of a specimen by applying water pressure, as
reported by Drumm et al. (1997). To facilitate moisture movement the specimen has to
be compacted on top of a saturated porous disk and the water in the soil pores connected
to the water in the porous disk. However as, because a lightly stabilised material
specimen has to be cured before testing, compacting it on top of the porous disk to
maintain water continuity is not possible, it was decided to change the moulding
moisture content.

270
Chapter 6 Permanent Deformation Behaviour using RLT

Table 6.1: RLT testing program

Granular Moisture Confining


Test Binder %
material content % pressure (kPa)
50
100
8 0.5%
140
180
50
CM RLT test
6.5 0.5% 100
140
8 50
6.5 1.5% 50
5 50

6.2.2 Multi-stage Permanent Deformation Testing with Varying Moisture


Contents

Table 6.1 shows the program used for multi-stage permanent deformation testing in
which specimens were compacted with CM. The procedures involved in sampling,
compacting and curing of the 195 mm height and 100 mm in diameter cylindrical
specimens were similar to those described in Chapter 5. The reconstituted particle size
distribution of CM were mixed with water and binder and compacted in 7 layers.
Specimens stabilised with 0.5% of CF were tested for moisture contents of 8% and
6.5% for four and three different confining pressures respectively in which 8% was the
OMC. Three moisture contents of 8%, 6.5% and 5% were selected for specimens
stabilised with 1.5% of CF binder at only one confining pressure of 50 kPa. For each
moisture content, a certain compaction effort was applied to obtain a dry density close
to the MDD. The compaction efforts applied, which increased for decreasing moisture
contents to improve compaction, and MDDs obtained are shown in Table 6.2 for the
selected moisture contents. The compacted specimens were cured for 7 days and
assembled in the MTS system with the internal deformation measurement setups
explained in Chapter 5.

271
Chapter 6 Permanent Deformation Behaviour using RLT

Table 6.2: Compaction efforts

Moisture Number of Proctor Achieved relative


content (%) standard compactions MDD (%)
8.0 25 100

6.5 33 98.5

5.0 40 97

Cyclic loading was applied using a harversine-shaped load at a 10 Hz frequency with a


0.9s resting period. The deviatoric stresses were selected based on the confining
pressures and, for each, 10,000 load cycles were applied. For each specimen, four to six
deviatoric stresses were selected from Table 63 for a particular confining pressure with
new specimens used for each confining pressure.

Table 6.3: Stress combinations for multi-stage permanent deformation testing

Confining
Stress Maximum axial
pressure (3) 1/3
sequence stress (1) (kPa)
(kPa)
50 150 3
1 50 300 6
2 50 450 9
3 50 650 13
4 50 1000 20
5 50 1300 26
6 50 1600 32
1 100 300 3
2 100 550 5.5
3 100 800 8
4 100 1000 10
5 100 1300 13
6 100 1500 15
1 140 280 2
2 140 560 4
3 140 840 6
4 140 1120 8

272
Chapter 6 Permanent Deformation Behaviour using RLT

5 140 1400 10
6 140 1680 12
Confining
Stress Maximum axial
pressure (3) 1/3
sequence stress (1) (kPa)
(kPa)
1 180 360 2
2 180 540 3
3 180 720 4
4 180 900 5
5 180 1080 6
6 180 1260 7
7 180 1440 8
8 180 1620 9

6.3 Results and Discussion

The RLT test results were analysed to determine the specimens’ resilient and permanent
deformation behaviours and dependence on the moisture content. As discussed in
chapter 5, these results can be interpreted well using the shakedown theory.

6.3.1 Classification of Material Responses using Shakedown Limits

According to the shakedown theory, a material subjected to cyclic loading can be


classified as:

1. Purely elastic
2. Elastic shakedown
3. Plastic shakedown
4. Incremental collapse or ratcheting

However, Werkmeister et al.(2005) showed that unbound granular base materials under
cyclic loading exhibit three different deformation responses called ranges A, B and C.
Range A behaviour shows a small amount of permanent strain for the first few
thousands of load cycles and ‘shakes down’ to an elastic response. A material with

273
Chapter 6 Permanent Deformation Behaviour using RLT

range B behaviour accumulates permanent deformation with each load cycle and fails
due to rutting at a high number of them. During range C behaviour, resilient
deformation increases with each load cycle and fails at a lower number of them.
Classifying a material’s response in a particular loading to any of these categories can
be carried out by analysing its permanent and resilient deformation responses.

However, to make categorisation easy, EN_13286-7 (2004) and Werkmeister et


al.(2005) published shakedown limits for unbound granular materials based on the
accumulated permanent strain from 3000 to 5000 load cycles. A stress level is
considered as reaching the plastic shakedown limit (PSL) for a strain of 0.045*10-3 and
plastic creep limit (PCL) for a strain of 0.4*10-3. One significant reason for defining
these limits is that a RLT test can be terminated at fewer load cycles (typically 10,000)
instead of continuing for higher numbers for determining the type of response to a
particular loading. Also, they facilitate multi-stage RLT testing, similar to that in this
study in which different loading combinations are performed on a single specimen. The
importance of categorising particular loading combinations in one of these ranges is to
ensure that a pavement operates at a stress level within the shakedown limits to avoid
rutting failure.

Based on these limits, all the RLT test results were categorised as being in range A, B or
C and tabulated in Tables 6.4 to 6.6. As few stress sequences produced strain
differences almost equal to the limits of either 0.045*10-3 or 0.4*10-3, two ranges, A/B
or B/C were defined. Table 6.4 shows the classification of specimens compacted at an
8% moisture content and stabilised with 0.5% of CF. For a particular confining
pressure, with an increase in the stress ratio, the shakedown range changed from A to B
and then C. Specimens tested with confining pressures of 100 kPa and 180 kPa did not
have stress sequences with range A because the initial load was higher than required to
cause range A behaviour. There were a few final stress sequences sufficiently high to
cause a specimen to fail after only a few thousand load cycles which were automatically
allotted range C behaviour. Another two series of test results similarly categorised are
presented in Tables 6.5 and 6.6.

As the stress values selected for a particular confining pressure in individual tests were
different, it is difficult to compare them based on either the shakedown categorisation or

274
Chapter 6 Permanent Deformation Behaviour using RLT

permanent deformation values. Also, it must be remembered that, although these limits
were suggested for unbound granular materials, they were used for lightly stabilised
granular materials.

0.045*10-3 < (ep,5000 -ep,3000) Range A

0.045*10-3 < (ep,5000 -ep,3000) < 0.4*10-3 Range B

0.4*10-3 < (ep,5000 -ep,3000 ) Range C

Table 6.4: Classification of shakedown ranges of specimens compacted with 8%


moisture content and stabilised with 0.5% CF

Confining Maximum Stress


Stress pressure axial stress ratio (ep,5000 - Shakedown
sequence ep,3000)×10-3 range
(3) kPa (1) kPa (1/3)
1 50 345 6.9 0.0040 A
2 50 525 10.5 0.1166 B
3 50 693 13.9 0.3981 B
4 50 1020 20.4 _ C
1 100 549 5.5 0.1045 B
2 100 811 8.1 0.2550 B
3 100 1054 10.5 0.6318 C
4 100 1375 13.8 2.9335 C
5 100 1607 16.1 _ C
1 140 555 4.0 0.0488 A/B
2 140 851 6.1 0.1528 B
3 140 1159 8.3 0.4348 C
4 140 1468 10.5 0.8185 C
5 140 1742 12.4 3.0386 C
1 180 737 4.1 0.1344 B
2 180 1120 6.2 0.2424 B
3 180 1510 8.4 0.9949 C
4 180 1848 10.3 _ C

275
Chapter 6 Permanent Deformation Behaviour using RLT

Table 6.5: Classification of shakedown ranges of specimens compacted with 6.5%


moisture content and stabilised with 0.5% CF

Confining Maximum Stress


Stress (ep,5000 - Shakedown
pressure axial stress ratio
sequence ep,3000)×10-3 range
(3) kPa (1) kPa (1/3)
1 50 353 7.1 0.0942 B
2 50 501 10.0 0.1275 B
3 50 660 13.2 0.1976 B
4 50 975 19.5 1.7370 C
5 50 1352 27.0 _ C
1 100 545 5.4 0.0951 B
2 100 782 7.8 0.1759 B
3 100 985 9.9 0.5098 C
4 100 1301 13.0 5.9834 C
5 100 1507 15.1 _ C
1 140 284 2.0 0.0171 A
2 140 574 4.1 0.0522 B
3 140 841 6.0 0.0718 B
4 140 1102 7.9 0.2222 C
5 140 1369 9.8 0.8728 C
6 140 1624 11.6 _ C

Table 6.6: Classification of shakedown ranges of specimens compacted with 8.0%,


6.5% and 5.0% moisture contents and stabilised with 1.5% CF

Confining Stress
Moisture Stress (1) (ep,5000 - Shakedown
pressure ratio
content % sequence kPa ep,3000)×10-3 range
(3) kPa (1/3)
1 50 156 3.1 0.006 A
2 50 272 5.4 0.008 A
8%
3 50 412 8.2 0.045 A/B
(OMC)
4 50 639 12.8 0.097 B
5 50 986 19.7 0.275 B
1 50 496 9.9 0.013 A
2 50 620 12.4 0.030 A
6.5% 3 50 984 19.7 0.156 B
4 50 1324 26.5 0.418 B/C
5 50 1659 33.2 - C
1 50 450 9.0 0.046 A
2 50 515 10.3 0.060 B
5% 3 50 913 18.3 0.064 B
4 50 1267 25.3 0.061 B
5 50 1608 32.2 0.198 B

276
Chapter 6 Permanent Deformation Behaviour using RLT

6.3.2 Calculation of Shakedown Limits

Figure 6.1 shows the stress path selected for each confining pressure which relates the
stress ratio to the peak axial stress in which each stress sequence marked with a symbol
is classified as one of the three ranges as explained in the previous section. In this
section, based on the categorisation used in the previous section, calculations for
determining the PSL and PCL are presented.

The PCL, which divides ranges B and C and the region, can be obtained by drawing a
line between the upper-limit range B and lower-limit range C values. As shown in
Figure 6.2, to obtain the PCL for each confining pressure when the transition took place
from range B to range C (i.e., the stress ratio and axial stress corresponding to the PCL),
interpolation was used. The stress ratio and axial values calculated in the interpolation
were based on the difference between the cumulative permanent strains of points X and
Y which was also used to define different ranges of the shakedown limits. As
Werkmeister et al. (2005) explained, if the test is performed over a range of confining
pressures, where points X and Y are relatively close, , the average values can be used to
obtain the stress ratio and peak axial stress values of the shakedown limit lines. A
similar approach can also be used to obtain the PSL for which shake down limit change
from range A to B has to be considered.

Figures 6.1 and 6.3 show the shakedown limits calculated for specimens stabilised with
0.5% of CF binder and compacted with 8% and 6.5% moisture contents respectively. In
Figure 6.2, as only two transitions took place from range A to B at confining pressures
of 50 kPa and 140 kPa, these two points were used to form the PSL equation. For
specimens compacted with 6.5% moisture content, only the PCL was determined as
there was insufficient data to obtain the PSL equation. Although shakedown limit
calculations could not be performed on specimens stabilised with 1.5% of CF as they
were tested for only one confining pressure, a similar approach could be used to obtain
the limit equations.

277
Chapter 6 Permanent Deformation Behaviour using RLT

Figure 6.1: Shakedown limit lines of specimens moulded with 8% moisture content
and 1.5% CF binder

Y -3
(ep,5000 - ep,3000= 0.435 )×10
PCL Stress ratio = 8.3
Axial stress = 1159
-3
(ep,5000 - ep,3000) = 0.4×10
Stress ratio = 8.0
Axial stress = 1121 kPa

X
-3
(ep,5000 - ep,3000) = 0.153×10
Stress ratio = 6.1
Axial stress = 851 kPa

Figure 6.2: Interpolation for determining exact point of shakedown limit line

278
Chapter 6 Permanent Deformation Behaviour using RLT

Range C

Range B

Figure 6.3: Shakedown limit lines of specimens moulded with 6.5% moisture
content and 1.5% CF binder

The shakedown limits established in Figures 6.1 and 6.3 can be represented by
exponential as well as power equations, as shown in Table 6.7 in which both types
indicate good correlations. The accuracy of these equations can be further improved by
increasing the number of confining pressures and decreasing the interval between the
axial stresses.

Moreover, the PCL equations for 8% and 6.5% moisture contents were plotted for
comparison in Figure 6.4. It can be seen that, for a particular stress ratio, the required
axial stress to bring the material to the PCL was higher for a 6.5% than 8% moisture
content, particularly at lower axial stresses. This shows that reducing the moisture
content from 8% to 6.5% gives more resistance against rutting.

This section demonstrated that the PSL and PCL can be established given the criteria
available for defining different shakedown limits. Using the limits suggested for an
unbound granular material, these equations were developed for specimens stabilised

279
Chapter 6 Permanent Deformation Behaviour using RLT

with 0.5% of CF and could be used to predict the material’s performance for a given
stress combination while considering the impact of the moisture content.

Table 6.7: Shakedown limit equations

Moisture
Plastic creep limit (PCL) Plastic shakedown limit (PSL)
content

 1  1965 e  0 .076 ( 1 /  3 )  1  739 e  0 .072 ( 1 /  3 )


8% R2 = 0.94 R2 = 1

 1  1965 ( 1 /  3 )  0 .781  1  991 ( 1 /  3 )  0 .481


R2 = 1
R2 = 0.96

 1  2108 e  0 .08 ( 1 /  3 ) -
R2 = 0.87
6.5%
 1  7355 ( 1 /  3 )  0 .897 -
R2 = 0.89

2000

PCL_MC = 8% (Exponential eqn)


PCL_MC = 6.5% (Exponential eqn)
1600 PCL_MC = 8% _ (power eqn)
PCL_MC = 6.5% (power eqn)
Peak axial stress (s 1) (kPa)

1200

800

400

4 8 12 16 20
Stress ratio (s 1/s 3)

Figure 6.4: Comparison of equations for PSL and PCL of specimens compacted at
8% and 6.5% moisture contents

280
Chapter 6 Permanent Deformation Behaviour using RLT

6.3.3 Analysis of Resilient Deformation

According to the shakedown theory, the resilient response can be a good indicator for
categorising the range of responses. Its variation with the load cycles obtained for all
stress sequences of both materials stabilised with 0.5% and 1.5% of CF binder and,
along with the categorisation previously made based on the accumulated permanent
strain, is provided in Annex C.

Resilient behaviour in range A

Almost all the responses in shakedown range A began with a high resilient deformation
value and decreased to a constant one after a few thousand load cycles. A typical
response obtained for a specimen moulded with 8% moisture content, stabilised with
0.5% CF and tested for 50 kPa confining pressure is shown in Figure 6.5, with
responses for the other stress sequences presented in Annex C. This behaviour remained
similar irrespective of the confining pressure, stress ratio, moisture content and binder
content. These decreases in resilient deformation meant that the resilient modulus
increased for the first few thousand load cycles and then remained constant. According
to the shakedown theory, continuing the test for more load cycles won’t fail the
material. Although it’s ideal to have a pavement operating in range A, in terms of
design, achieving an appropriate layer thickness for such conditions may be expensive.

Resilient behaviour in range B

The majority of stress sequences were classified as range B behaviour, with resilient
deformation exhibiting two different responses depending on the binder content. For
specimens stabilised with 0.5% CF, their deformation responses were similar to those in
range A, decreasing from higher values to stable ones after a few thousand load cycles.
According to the theory, range B eventually reaches failure at a higher number of load
cycles which would not have been possible in this test as the loading was limited to
10,000 cycles. Therefore, differences between the resilient deformation responses in
ranges A and B may not be noticeable if the numbers of load cycles are limited. Typical

281
Chapter 6 Permanent Deformation Behaviour using RLT

responses obtained for specimens compacted with 6.5% moisture content and tested for
stress sequence 1 when subjected to 100 kPa are shown in Figure 6.6. However, the
range B responses in specimens stabilised with 1.5% of CF did not show any reduction
but, instead, increased in the first 1,000 load cycles and remained at stable values.

Resilient behaviour in range C

Typically, in range C behaviour, a material accumulates permanent deformation at a


rapid rate with increasing resilient deformation and fails at a relatively low number of
load cycles. In fact, Austroads (2010a) defines a reduction in the resilient modulus to
half its initial value as a failure criteria for unbound materials and, for a cemented
material, the flexural modulus is measured instead. Figure 6.7 shows range C response
immediately before failure in which deformation increased with the number of load
cycles. However, in this study, not all the stress sequences falling in range C showed
increasing resilient responses, as most resilient deformations were turbulent responses,
particularly in stress sequences in which failure occurred. Although increase in resilient
deformation were observed in many stress sequences which completed all the 10,000
load cycles, some stress sequences showed similar trends to range A or B in which
resilient deformation declined and reached a stable value after a certain number of load
cycles.

In summary, although resilient deformation responses showed some characteristics of


the three shakedown ranges, there were stress sequences which could be classified as
being in any one of them. This could have been partly due to the number of load cycles
being limited to 10,000.

282
Chapter 6 Permanent Deformation Behaviour using RLT

0.44
Range A typical response
MC = 6.5%, Conf.pre = 50 kPa, 1.5% CF
Stress sequence 1
0.4
Resilient strain (10-3)

0.36

0.32

0.28

0.24
0 2000 4000 6000 8000 10000
Number of load cycles

Figure 6.5: Typical resilient deformation response in shakedown range A

283
Chapter 6 Permanent Deformation Behaviour using RLT

Range B typical responses


MC = 8.0%, Conf.pre = 50 kPa, 1.5% CF
1.4 Stress sequence 5
MC = 6.5%, Conf.pre = 100 kPa, 0.5% CF
Stress sequence 1

1.2
Resilient strain (10-3)

0.8

0.6
0 2000 4000 6000 8000 10000
Number of load cycles

Figure 6.6: Typical resilient deformation responses in shakedown range B

284
Chapter 6 Permanent Deformation Behaviour using RLT

2.8

2.4
Resilient strain (10-3)

Range C
2

1.6

1.2
0 1000 2000 3000 4000
Number of load cycles

Figure 6.7: Typical resilient deformation response in shakedown range C just before failure of specimen

285
Chapter 6 Permanent Deformation Behaviour using RLT

The resilient strains measured after the 1,000th load cycle and plotted against the stress
ratios for specimens stabilised with 1.5% of CF binder are shown in Figure 6.8 in which
they are in all three shakedown ranges and increase almost linearly with an increasing
stress ratio. For a particular stress ratio, the highest resilient strain was recorded for a
specimen stabilised with a high moisture content; in other words, a decreasing moisture
content increased the resilient modulus. Similarly, Figure 6.9 shows that specimens
compacted with 8% moisture content with 0.5% of CF deformed higher than those with
6.5% moisture content, particularly at with a 50 kPa confining pressure.

1.6 Moisture content = 8%


Moisture content = 6.5%
Moisture content = 5%

1.2
Resilient strain(?r) (  10-3)

0.8

0.4

0
0 10 20 30 40
Stress ratio (s 1/s 3)

Figure 6.8: Resilient strain varying with stress ratio for specimens stabilised with
1.5% CF binder and compacted at different moisture contents

286
Chapter 6 Permanent Deformation Behaviour using RLT

MC = 6.5%, Confining pressuure = 50 kPa


MC = 6.5%, Confining pressuure = 100 kPa
2
MC = 8.0%, Confining pressuure = 50 kPa
MC = 8.0%, Confining pressuure = 100 kPa

1.6
Resilient strain (  10-3)

1.2

0.8

0.4
4 8 12 16 20
Stress ratio (s 1/s 3)

Figure 6.9: Resilient strain varying with stress ratio for specimens stabilised with
0.5% CF binder and compacted at different moisture contents

6.3.4 Influence of Moisture on Resilient Modulus

In the previous section, it was shown that the Mr varied over the course of testing and
depended on the shakedown range. As it is common to measure the Mr after 1,000 load
cycles, the resilient deformations and deviatoric stresses obtained after 1,000 load
cycles for each stress sequence were used to determine the Mr and study its dependence
on the moisture content. As the data were limited for specimens with 1.5% of CF, only
those stabilised with 0.5% of CF were used for the comparison.

The Mr of specimens stabilised with 8% and 6.5% compaction moisture contents, along
with their confining pressures and axial stresses, are shown in Table 6.8. Comparing the
absolute values may be difficult as the bulk stresses are not the same in both moisture

287
Chapter 6 Permanent Deformation Behaviour using RLT

contents. Since the k-θ model was successfully used to relate the Mr of a lightly
stabilised material to the bulk stress in the previous chapter, a similar approach was
selected to compare the Mr (Figure 6.10). The Mr measured from the last stress sequence
in each confining pressure was omitted as failures of the materials in these sequences
and their dramatic range C behaviours did not produce Mr consistent with the rest of the
data.

Although the coefficients of determination of 0.85 and 0.79 showed good fits with the
data, in the previous chapter, resilient modulus testing consistently produced R2 greater
than 0.9. This could be because the test results presented here were combinations of
those from three or four different tests whereas, in the previous chapter, a resilient
modulus test was performed on a single specimen.

The Mr of specimens compacted with 6.5% moisture content showed a marginally


higher value than that one compacted at 8%; in particular, the differences were high at a
low bulk stress and reduced with higher bulk stresses. Therefore, the Mr values were
consistent with the findings presented in the previous section, that is, a reduction in the
moisture content helped the material to gain more resistance against rutting

288
Chapter 6 Permanent Deformation Behaviour using RLT

Table 6.8: Resilient modulus obtained for CM

Moisture content = 8% Moisture content = 6.5%


Confining Bulk Resilient Resilient
Stress Deviatoric Stress Confining Deviatoric Bulk stress
pressure stress msodulus modulus
sequence stress (kPa) sequence pressure (kPa) stress (kPa) (kPa)
(kPa) (kPa) (MPa) (MPa)
1 50 268 445 369 1 50 276  453 402
2 50 432 625 405 2 50 410  601 510
3 50 584 793 429 3 50 555  760 546
4 50 882 1120 - 4 50 841  1075 579
5 50 1184  1452 -
1 100 408 749 510
2 100 646 1011 597 1 100 404  745 511
3 100 867 1254 611 2 100 620  982 606
4 100 1159 1575 639 3 100 805  1185 607
5 100 1370 1807 - 4 100 1092  1501 644
5 100 1279  1707 -
1 140 378 835 499
2 140 647 1131 632 1 140 101  564 400
3 140 926 1439 825 2 140 387  854 525
4 140 1208 1748 899 3 140 644  1121 584
5 140 1456 2022 1056 4 140 893  1382 739
5 140 1147  1649 858
1 180 506 1097 516 6 140* 1388  1904 -
2 180 854 1480 595
3 180 1209 1870 690
4 180 1516 2208 -

289
Chapter 6 Permanent Deformation Behaviour using RLT

1200

1000

Resilient modulus (MPa)


Mr=20.643( )0.4928
800
R2=0.85 Mr=6.6384( )0.6391
R2=0.79

600

400

200
0 500 1000 1500 2000 2500
Bulk stress (kPa)

Figure 6.10: Resilient modulus obtained for 8% and 6.5% moisture contents

6.3.5 Characterisation of Permanent Deformation Behaviour

The permanent deformation obtained for each stress sequence was analysed using two
existing models. There are numerous models available for predicting permanent
deformations, with some based on the number of loading cycles and some on the
stresses. Two models were used to analyse the data, with the regression coefficients
obtained for both curve fittings shown in Table 6.9.

The log-based model shown in equation (6.1), in which the accumulated permanent
strain is related to the logarithmic number of load cycles, was proposed by Barksdale
(1972). The regression constant value ‘a1’ is related to the permanent deformation
obtained after the first loading cycle and has to be a positive number. After testing for
106 load cycles, Sweere (1990a) found that this log-based model does not fit the data
well and proposed the power-based equation (6.2) (Lekarp et al. 2000). Similar to the

290
Chapter 6 Permanent Deformation Behaviour using RLT

log-based one, parameter ‘a1’ is related to the permanent deformation in the first loading
cycle.

In multi-stage RLT tests, as multiple axial stresses are applied in succession, the effect
of the stress history is unavoidable. Virgin specimens deform higher than those
preloaded with different axial stresses because of the stiffening effect resulting from
particle re-arrangement and increased density. Because of the effects of the stress
history, the permanent deformations measured in the first loading sequence showed
higher values than those in subsequent sequences irrespective of their having higher
deviatoric stresses. Therefore, deformation of the first 100 load cycles from the initial
load sequence in each test was omitted from the results analysis. In their proposed
model, Paute et al. (1996) suggested eliminating permanent deformation from the first
100 load cycles.

εp = a1 + b1log(N) Eq. (6.1)

εp = a2N b2 Eq. (6.2)

where:
εp = permanent strain,
N = number of load cycles
a1 , a2 , b1, b2 = regression coefficients

Annex C shows the permanent deformation varying with the number of load cycles
obtained from RLT tests. The above two equations were fitted with the experimental
data to obtain the regression coefficients based on which, as can be seen in Tables 6.9
and 6.10, the log model showed a better fit with the experimental data than the power
model. As previously mentioned, the regression constant ‘a1’ related to the permanent
strain in the first loading should be a positive value. In fact, the experimental data used
for the regression analysis had positive values. However, when the regression line fitted
the data in the best possible way to obtain the highest R2, the ‘a1’ values became
negative. Furthermore, it can be noted that the ‘a1’ value continued to decrease with
subsequent loadings because it depends on the gradient of the initial load cycles. With
an increase in axial stress, the gradient becomes less and reaches further down the y axis
and, for lower loads, the grading becomes high and reaches the y axis at a lower level.

291
Chapter 6 Permanent Deformation Behaviour using RLT

Therefore, the parametric study did not consider the regression constant ‘a1’. The log
model was chosen for further analysis over the power model as it fitted well with the
data.

Table 6.9: Regression coefficients obtained for 8% moisture content

Confining Equation (6.1) Equation (6.2)


pressure
a1 b1 R2 a2 b2 R2
(kPa)
-0.4 0.15 0.90 0.08 0.28 0.77
50 -1.0 0.46 0.89 0.23 0.30 0.66
-4.1 0.99 0.95 0.04 0.54 0.74
-0.6 0.17 0.93 0.07 0.29 0.91
-1.7 0.61 0.98 0.27 0.30 0.75
100
-5.9 1.31 0.97 0.04 0.58 0.80
-25.2 5.16 0.96 0.09 0.62 0.82
-0.3 0.13 0.98 0.07 0.27 0.81
-1.0 0.39 0.97 0.23 0.27 0.80
140 -3.3 0.97 0.96 0.16 0.41 0.74
-8.6 2.01 0.97 0.08 0.54 0.78
-28.6 5.32 0.94 0.02 0.81 0.89
-0.6 0.16 0.88 0.01 0.43 0.89
180 -0.9 0.80 0.95 0.77 0.23 0.75
-8.4 2.34 0.99 0.32 0.42 0.72

After eliminating ‘a1’ from the log equation, the permanent deformation equation
becomes

εp = b1log(N) Eq. (6.3)

The regression constant ‘b1’ of each confining pressure varying with the bulk stress is
shown in Figures 6.11 and 6.12 for 8% and 6.5% moisture contents respectively, and it
was found that the exponential equation (6.4) was a good fit with the values.

ɷp
b1 = ye Eq. (6.4)

where:
p = bulk stress (kPa) and
y, ɷ - regression constants

292
Chapter 6 Permanent Deformation Behaviour using RLT

By substituting b1 in equation (6.3), the original permanent deformation can be


modified to equation (6.5). As shown in Table 6.11, y and ɷ decreased with an
increasing confining pressure which can be expressed by a power or exponential law as
in equations (6.6) to (6.9).

εp = yeɷp log(N) Eq. (6.5)

βσ3
y=αe Eq. (6.6)
β
y=ασ3 Eq. (6.7)
δσ3
ɷ= λe Eq. (6.8)
δσ3
ɷ= λe Eq. (6.9)

where α, β, λ and δ are regression constants.

Table 6.12 shows that all regression constants showed good correlation for the selected
equation. Hence, substituting any combinations of equations 6.6 to 6.9 in equation 6.5,
the PD can be expressed in terms of bulk stress and confining pressure.

Table 6.10: Regression coefficients obtained for 6.5% moisture content

Confining Equation (6.1) Equation (6.2)


pressure 2
a1 b1 R a2 b2 R2
(kPa)
0.18 -0.6 0.99 0.07 0.30 0.88
0.28 -0.9 0.98 0.08 0.34 0.79
50
0.49 -2.1 0.96 0.03 0.52 0.8
3.14 -13.6 0.98 0.26 0.46 0.822
0.30 -0.5 0.99 0.06 0.30 0.9
0.41 -1.1 0.99 0.22 0.28 0.82
100
1.09 -4.9 0.97 0.04 0.55 0.82
8.07 -43.1 0.89 0.11 0.64 0.92
0.02 0.1 0.34 0.05 0.15 0.42
0.14 1.2 0.92 1.34 0.07 0.83
140 0.23 -0.2 0.94 0.33 0.20 0.70
0.58 -1.8 0.96 0.14 0.37 0.72
1.79 -8.3 0.97 0.04 0.60 0.81

293
Chapter 6 Permanent Deformation Behaviour using RLT

Conf. pressure 50 kPa Regression line 50 kPa


Conf. pressure 100 kPa Regression line 100 kPa
6 Conf. pressure 140 kPa Regression line 140 kPa
Conf. pressure 180 kPa Regression line 180 kPa

4
Regression constant 'b'

0 400 800 1200 1600 2000 2400


Bulk stress (kPa)

Figure 6.11: Regression constant ‘b1’ varying with bulk stress (MC=8%)

Conf. pressure 50 kPa Regression line 50 kPa


Conf. pressure 100 kPa Regression line 100 kPa
10 Conf. pressure 140 kPa Regression line 140 kPa

8
Regression constant 'b'

0 400 800 1200 1600 2000


Bulk stress (kPa)

Figure 6.12: Regression constant ‘b1’ varying with bulk stress (MC=6.5%)

294
Chapter 6 Permanent Deformation Behaviour using RLT

Table 6.11: Regression constant obtained for b1

Moisture Confining
y ɷ R2
content pressure (kPa)
50 0.0132 0.0055 0.99
100 0.0087 0.0041 0.99
8%
140 0.0109 0.0031 0.99
180 0.0039 0.0035 0.99
50 0.0183 0.0047 0.97
6.5% 100 0.0074 0.0045 0.93
140 0.0024 0.0041 0.96

Table 6.12: Regression constant obtained for exponential and power equations

    8% moisture content 6.5% moisture content


      R2 R2
Exponential eq. α = 0.021 α = 0.060
0.90 0.98
(6.6) β = -0.009 β =-0.022

α = 0.359 α =30.71
Power eq. (6.7) 0.81 0.94
β = -0.823 β =-1.873

Exponential eq. λ =0.006 λ =0.005


0.79 0.93
(6.8) δ =-0.004 δ =-0.001
ɷ 
λ =0.026 λ =0.008
Power eq. (6.9) 0.89 0.85
δ =-0.403 δ =-0.122

6.3.6 Relationship between Permanent and Elastic Deformations

Resilient deformation may not be a good indicator for classifying PD characteristics


(Puppala et al. 1999). However, the permanent strain rate is a good indicator for
classifying materials’ responses and is determined by dividing the permanent strain by
the number of load cycles. To establish a relationship between permanent and elastic
deformations, the permanent strain rate and resilient deformation obtained after 1,000
and 10,000 load cycles were related, as shown in Figures 6.13 and 6.14 for 8% and
6.5% moisture contents respectively. Better correlations can be seen for the measured

295
Chapter 6 Permanent Deformation Behaviour using RLT

values obtained after 10,000 than 1,000 cycles, particularly for an 8% moisture content
in which the coefficient of determination increased from 0.37 to 0.68. Although
correlations were not overly promising for specimens compacted at an 8% moisture
content, those at 6.5% showed better relationships, with R2=0.88 and 0.91 for the
measured values obtained after 1,000 and 10,000 load cycles respectively.

0.012
Measured after 1000 load cycles
Exponential fit_1000 load cycles
0.01 Measured after 10000 load cycles
Exponential fit_10000 load cycles
Permanent strain rate ( 10-3)

0.008

0.006

0.004
y = 0.0001e1.509x
R2 =0.68
0.002

0.8 1.2 1.6 2


Resilient strain (  10-3)

Figure 6.13: Permanent strain rate varying with resilient strain of specimens
compacted at 8.0% moisture content

296
Chapter 6 Permanent Deformation Behaviour using RLT

Measured after 1000 load cycles


Exponential fit_1000 load cycles
Measured after 10000 load cycles
Exponential fit_10000 load cycles
0.012
y = 0.0001e2.6181x
Permanent strain rate ( 10-3) R2 =0.88

0.008

y = 0.00001e3.4758x
R 2 =0.91

0.004

0 0.4 0.8 1.2 1.6 2


Resilient strain (  10-3)

Figure 6.14: Permanent strain rate varying with resilient strain of specimens
compacted at 6.5% moisture content

6.4 Summary

PD behaviour of lightly stabilised granular material was studied using RLT test was
performed on CM stabilised with 0.5% and 1.5% of CF and cured for 7 days.
Specimens stabilised with 0.5% of CF were moulded with two moisture contents of 8%
and 6.5 and three moisture contents of 8%, 6.5% and 5% were selected for specimens
stabilised with 1.5% of CF binder %. More than one deviatoric pressure was applied for
a particular confining pressure, typically, each specimen subjected to four to five axial
pressure.

The obtained PD behaviour from each stress sequences were classified based on the
accumulated permanent strain from 3000 to 5000 load cycles to range A, B or C
shakedown behaviour. These shakedown limits were used to obtain plastic creep line
(PCL) and plastic shakedown line (PSL). The established shakedown limit equations

297
Chapter 6 Permanent Deformation Behaviour using RLT

can be used to predict the material’s performance for a given stress combinations while
considering the impact of moisture content. However, the addition test data with more
stress combinations can be used to improve the accuracy of the equation. The
shakedown range categorisation was based on the limits suggested for the unbound
granular material. The validity of the equation to lightly stabilised material can be
performed by conducting series of single stage RLT test.

Resilient deformation responses analysed showed some characteristics of the three


shakedown ranges, but there were stress sequences which could be classified as any one
of these ranges. This could be partly because the number of load cycles being limited to
10,000. Mr of specimens stabilised with 8% and 6.5% compaction moisture contents
were studied using k-θ model which shows a good fit the data. Mr of specimen
compacted with 6.5% showed marginally higher value than specimen compacted with
8% moisture content particularly the differences were high at the lower bulk stress and
reduced with increasing bulk stress.

Two permanent deformation models, log and power based models, which relate
permanent strain with number of loadings were used to analyse the permanent
deformation obtained for each stress sequence. The regression fit of the date with the
equation suggested that the log model showed better fit to the experimental data than
the power model. Hence the power model was used to further analyse the data and to
obtain its dependency on stress values and to establish an equation to predict PD in
terms of bulk stress and confining stress.

298
7 Chapter 7

Characterising Deformation Behaviour of Lightly Stabilised


Granular Material using Accelerated Pavement Model (APM)
Test

7.1 Introduction

Laboratory studies of pavement materials in terms of characterisations, performance


evaluations and durability tests often involve element testing. The aim of all laboratory
element tests is to somehow relate their outcomes to real-world problems. Investigating
different variables and studies involving multiple variables require series of tests which
could only be feasible with element testings. In earlier chapters of this study also,
element tests, such as the unconfined compressive strength (UCS), flexural and triaxial,
were presented as part of a durability study. Outcomes from laboratory element tests
cannot be directly used to in-service pavements because their uncertainties and those of
the models developed require validation which can be addressed through accelerated
pavement model (APM) testings.

APM tests provide a link between the outcomes from laboratory element tests and
pavement structures constructed using the same pavement materials as in the field and
models developed in laboratories are better translated to in-service pavements through
APM tests. The first empirical pavement design standard, developed by the Association
of State Highway Officials (AASHO), was based on series of APM tests performed in
the late 1950s. Nowadays, extensive APM tests performed around the world provide
more insights into the mechanical behaviours of pavement materials. Current
mechanical empirical (M-E) designs have many empirical models that require
mathematical models to fully transform them based on analytical solutions which could
possibly be obtained through APM testing.
Chapter 7 Characterising Deformation Behaviour using APM Test

APM tests are performed on either field test roads or model pavements constructed in a
laboratory by applying linear or circular loads, usually for large numbers of load cycles,
which is difficult to achieve in a triaxial or any other element test. In a typical APM test,
specially designed wheel loadings with controlled pressures closely simulate in-situ tyre
pressures are generally applied on top of pavement layers to monitor rutting and fatigue
behaviours. The term ‘accelerated’ means that wheel loads are applied at higher rates or
frequencies than it could possibly be experienced in an actual pavement. Therefore,
APM requires less time than field testing and interruptions to traffic are avoided.

Nowadays, heavy vehicle simulators (HVS), which are mobile devices with linear
loadings, are used for APM testing in a section of a field or model pavements. This
arrangement enables engineers to conduct tests under closely simulated field conditions.
The mobility of HVS allow any pavement section to be assessed in terms of the
feasibility of introducing new materials or new construction techniques, and for
ascertaining the requirements for rehabilitation. During testing, wheel pressures are
controlled and allowed to run back and forth over the selected pavement section.

The first HVS testing program was commissioned in South Africa in the early 1970s
and currently provides major inputs for its pavement design guide. Several programs
using HVSs have covered the vast majority of issues, including rehabilitation methods
for pavements, validations of laboratory findings, evaluations of the performance of
porous asphalt, tests of ultra-thin continuously reinforced concrete pavements and
assessments of the long-term performances of pavements (Kannemeyer and Plessis
2006; Rust et al. 1992; Steyn et al. 1997).

As previously mentioned APM testing of a model pavement in a laboratory is also


common and may be preferred over field studies in the absence of HVSs and with
limited resources. In Australia, the Australian Road Research Board (ARRB) has a
fixed APM testing facility which is used to simulate linear heavy vehicle traffic loads
using a load assembly trolley. APM tests are preferred over in-situ testing because of
the easiness in performing the tests.

In this investigation, an APM test was performed on a model constructed in a concrete


cylindrical tank of diameter 1.175 m for characterising the deformation behaviours of

300
Chapter 7 Characterising Deformation Behaviour using APM Test

lightly stabilised granular material. Apart from measuring deformations, sensors to


measure moisture and pressure were installed to understand the insight of the material’s
behaviour. A granular base material stabilised with 1.5% CF of thickness of 150 mm
constructed on top of 600 mm subgrade was used for the testing. Four cyclic pressures
were applied at a rate of 2 Hz and the applied pressure, vertical deformations, moisture
and pressures were measured at different locations. The obtained deformations data
were used to back calculate the moduli of both subgrade and base using a finite element
(FE) model developed in ABAQUS. The measured pressures were compared against
pressures obtained in the FE model. Finally, using repeated load triaxial (RLT) test data
and FE analysis, relationship between regression coefficients obtained from
accumulated permanent deformation was developed in terms of bulk-stress.

7.2 Accelerated Pavement Model Testing Facility

In this section, the APM testing facility available in the civil engineering laboratory at
the University of New South Wales, Canberra, is discussed in detail. The entire testing
setup, which includes a cylindrical tank, steel loading frame and loading actuator, is
shown in Figure 7.1.

The concrete cylindrical tank shown in Figure 7.2, which has an internal diameter of
1.175 m and height of 1 m, was used to construct a model pavement. Its wall is made of
reinforced concrete, its bottom surface a thick steel plate fastened to the cylinder and it
rests on three steel ‘I’ sections fixed to its bottom surface.

The loading actuator is fitted with a steel beam as shown in Figure 7.3, has the capacity
to apply an axial load of 100 kN. Its steel frame has two vertical columns and a
horizontal beam, all three of which are ‘I’ sections with the dimensions shown in Figure
7.4. Initially, it was decided to apply a loading at the rate of 10 Hz to match the
frequency used for similar cyclic experiments but, as the loading frame could not handle
a high frequency, 2 Hz was selected.

The Instron servo hydraulic actuator was used to apply axial pressures of 750 kPa, 1000
kPa, 1250 kPa and 1500 kPa. Table 7.1 shows the number of loading cycles applied for

301
Chapter 7 Characterising Deformation Behaviour using APM Test

each stress sequences. Servo controllers have state-of-the-art technology for applying
continuously higher numbers of load cycles and can be programmed to automate the
loading system. However, as the control program did not have a harversine load pattern,
to apply the load, sinusoidal loading was selected.

302
Chapter 7 Characterising Deformation Behaviour using APM Test

Figure 7.1: Schematic diagram of accelerated pavement model testing facility

303
Chapter 7 Characterising Deformation Behaviour using APM Test

Figure 7.2: Cylindrical tank used for APM testing

Figure 7.3: Loading actuator assembly

304
Chapter 7 Characterising Deformation Behaviour using APM Test

190 mm

460 mm 10 mm

12 mm

Figure 7.4: ‘I’ section used for beam and columns in the steel frame

Table 7.1: Applied loading pressures and numbers of cycles

Stress Maximum axial Minimum axial Number of load


sequence pressure (kPa) pressure (kPa) cycles
1 750 50 8×104
2 1000 50 8×104

3 1250 50 8×104

4 1500 50 7×105

7.2.1 Measurement of Vertical Deformation

The vertical deformations of the model pavement were measured at three different
locations using linear variable differential transformers (LVDTs). The total vertical
deformation of both the base and subgrade was measured by determining the
deformation of the loading plate assuming that it was incompressible compared with the
soils. The 184-mm diameter loading plate is shown in Figure 7.5 with four small steel

305
Chapter 7 Characterising Deformation Behaviour using APM Test

plates attached to it in diametrically opposite directions on which to mount the four


LVDTs. The average of these four readings gives the total deformation of the pavement.

The deformation of the subgrade was measured on its top surface (at the interface) using
two simple in-depth vertical deformation measurement devices made in-house (Figure
7.6). The copper house acted as a shield to protect the steel rod from the soil and allow
it to move freely following movements of the bottom plastic circular disk, which had to
be placed upright on a firm base, with the LVDT mounted on the top circular disk. The
height of the total device had to be adjusted based on the depth of the soil it needed to
cover. Similarly, two more in-depth vertical deformation measurement devices were
placed in the middle of the stabilised base to measure vertical deformations at the mid-
depth of the base layer.

A plan view of all eight vertical deformation measurement points is presented in Figure
7.7. The in-depth vertical deformation measurement devices were arranged in a circle
with a diameter of 220 mm while measurements on the loading plate were arranged in a
circle approximately 200 mm in diameter.

Figure 7.5: 184-mm diameter loading plate with steel plates attached on which to
mount LVDTs for measuring top surface deformation

306
Chapter 7 Characterising Deformation Behaviour using APM Test

LVDT

Circular plastic disk

Copper house

Steel rod

Circular plastic disk buried into the soil

Figure 7.6: In-depth measurement device for measuring vertical deformations

Surface deformation
Mid-depth deformation

Subgrade deformation

Figure 7.7: Locations at which vertical deformation measurements were measured

307
Chapter 7 Characterising Deformation Behaviour using APM Test

7.2.2 Pressure Measurements

Four earth pressure cells (EPCs), each with a capacity of 1 MPa, were used to measure
the horizontal and vertical pressures at the locations shown in Figure 7.8. EPC 1 and
EPC 2 measured the vertical pressures at the bottom and middle of the stabilised layer
respectively, EPC 3 the horizontal pressure 184 mm from the centre and EPC4 the
horizontal pressure at the boundary wall. Before using the EPCs, they were calibrated
using a cylindrical split device. A simpler calibration using universal testing machine
could not be performed because the top surface of the EPC is not flat.

7.2.3 In-situ Moisture Measurements

In-situ measurements of the moisture content are needed in many field and large-scale
experiments and often involve continuous monitoring over a period of time. The
moisture content, which is one of the environmental factors that has a great influence on
a pavements’ performance changes during its life, due to many factors. As excessive
moisture can cause adverse effects, this is of serious concern for in a pavement design.
Therefore, it is essential to measure moisture variations, particularly in long-term in-situ
testing, in order to obtain better interpretations of load deformation characteristics
which requires continuous measurements of moisture while tracking the load and
deformation characteristics of a pavement

308
Chapter 7 Characterising Deformation Behaviour using APM Test

Loading plate (184 mm)


EPC 4 184 mm
EPC 3
EPC 2

Lightly stabilised granular 150 mm


base

Subgrade
EPC 1

1175 mm

Figure 7.8: Locations of installed EPCs

There are different methods available for measuring the in-situ moisture contents of
soils. The gravimetric moisture content (GMC) is a common traditional but destructive
method which measures the weights of soil samples taken directly from the field before
and after they are dried in an oven. Although proven to be accurate, it requires soil
samples, is often slow and is also difficult to use for obtaining ongoing moisture
measurements and those at greater depths. Therefore, non-destructive continuous
indirect test methods are preferred for measuring moisture contents. However, it must
be noted that the calibrations of indirect moisture measurement techniques are generally
performed by measuring the GMC. Soil-specific calibration is essential for improving
the accuracy of measurements as relying on a manufacturer’s calibration function is
likely to produce erroneous results (Mittelbach et al. 2012). Other factors that could
contribute to the selection of a moisture measurement technique include price, accuracy
and response time. Also, the size of the device matters if the probe has to be inserted
into a small soil specimen.

309
Chapter 7 Characterising Deformation Behaviour using APM Test

An indirect measurement technique, as the name implies, does not directly measure the
moisture content but, instead, measures different variables in a soil matrix which it
converts to a volumetric moisture content (VMC%) after calibration. These
measurements are non-destructive, and have been proven to be accurate after soil-
specific calibration, fast and capable of continuously monitoring moisture movements.
Time domain reflectometry (TDR), the standing wave, capacitance and neutron probes
are a few examples that use indirect measurement techniques to measure the moisture
content.

7.2.3.1 Time Domain Reflectometer

TDR is becoming an increasingly popular technique for measuring the VMCs of soils in
laboratories and fields. A TDR instrument generates short-duration electromagnetic
waves at a speed of light and sends them through a coaxial cable which eventually
transmits them to a soil via two or three waveguides embedded in the soil. Each
reflected wave is then sent back to the source and, based on its travel time, its velocity is
calculated which is then calibrated to read its VMC%.

The velocity of a reflected wave decreases with an increasing dielectric constant (DC)
of the soil. The DC of a material is the ratio between its permittivity to free space which
is 80 for water, 1 for air and 4 for solid particles. A change in the DC of a given
unsaturated soil is directly related to a change in the soil’s moisture content because of
the significantly different DCs of water, air and solid particles.

One of the main disadvantages of TDR sensors is its higher cost compared with
alternatives.

7.2.3.2 Neutron Probe

Unlike the previous moisture sensor, a neutron probe does not measure the DC but uses
the technique of neutron scattering to measure the hydrogen atoms in a soil. It emits fast
neutrons that collide with the hydrogen atoms in the soil, which are predominantly from
water molecules, with the number of neutrons that collide measured and then calibrated

310
Chapter 7 Characterising Deformation Behaviour using APM Test

to obtain the VMC%. Calibration is performed by relating the neutron probe readings to
the VMC% measured at different moisture contents. However, a neutron probe is not a
perfect device for use near the surface of a soil.

7.2.3.3 Standing-Wave Ratio (STR)

Moisture probes use the technology of the standing wave ratio (STR), a relatively new
method first proposed by Gaskin and Miller (1996), and, as they are less expensive than
TDR moisture devices, enables large numbers of them to be installed. The STR is
defined as the ratio of the transmission line impedance to electric impedance of a stem
probe (Wang and Zhao 2010). The electric impedance of probe depends on the
dielectric value of the soil and the transmission line impedance is a constant for a
particular device. Therefore, a change in the STR is caused by the dielectric value which
predominantly depends on the moisture content. TDR and the STR are based on the
principle of the dielectric value of a soil although the former measures velocity changes
in the electromagnetic waves whereas the latter is based on impedance of the
electromagnetic pulse.

7.2.3.4 Capacitance Probe

A capacitance probe uses radio frequencies to determine the dielectric permittivity of a


soil which is then converted to its VMC% after calibration. Although capacitance
probes are relatively cheap, their measurements are likely to be effected by the salinity
and temperature of a soil.

Comprehensive studies of different types of moisture sensors, which compare them in


terms of accuracy, repeatability and factors influencing measurements, are available
(Leib et al. 2003; Plauborg et al. 2005; Walker et al. 2004). Table 7.2 compares
different aspects of moisture sensors discussed above which can be used as a basic
guide for selecting a suitable moisture probe for a particular application.

311
Chapter 7 Characterising Deformation Behaviour using APM Test

7.2.3.5 Moisture Probe (MP306)

In this study, for measuring the moisture content, MP306 moisture probes manufactured
by ICT International were used. They used STR techniques to measure the VMC%
which was then converted to the GMC% after calibration. Figure 7.09 shows a MP306
moisture probe which has three 60 mm stainless steel needles arranged in a single plane
for full insertion into the soil to measure the moisture content. As its exterior is made of
ABS plastic which is sufficiently strong to withstand high soil pressures, it is possible to
use it repeatedly. A MP306 is directly connected to a soil moisture meter (SMM) from
which continuous measurements can be taken, with the data logger used to store the
data using the storage device. The whole moisture measurement system, MP306 probes,
data logger and SMM, is shown in Figure 7.10.

Table 7.2: Comparisons of different types of soil moisture sensors

TDR Neutron
STR sensor Capacitance probe
sensor probe
Excellent with soil Good with soil
Accuracy Excellent Excellent
specific calibration specific calibration
Excellent
Resolution Excellent Excellent Good (with
calibration)
Cost High Moderate Low High
Temperature
No No Yes No
Sensitivity
source: www.ictinternational.com

Table 7.3 shows some of the specifications of the MP306 obtained from the
manufacturer. Although response time, price and accuracy are the main advantages of a
MP306, as it is larger than other potential alternatives and may not be suitable for some
applications in which size is an issue, using TDR or capacitance based moisture probes
which are relatively smaller is more appropriate. ]

312
Chapter 7 Characterising Deformation Behaviour using APM Test

Table 7.3: MP306 specification

Measurement range 0-100 VSW%


± 5 VSW% using supplier’s
Accuracy calibration, 1% VSW% after
soil specific calibration
Response Time Less than 0.5 seconds
approximately 3 seconds from
Stabilisation Time
power-up
Cable length 4.5 m

source: www.ictinternational.com

7.2.3.6 Laboratory Calibration of MP306 Probes

Each moisture probe was individually calibrated to obtain a soil-specific VMC% in the
laboratory using a cylindrical sample 115 mm high and 105 mm in diameter compacted
in a standard UCS mould. An oven-dried base material was selected based on its
reconstituted particle size distribution, mixed with water and compacted using a
standard Proctor hammer. As discussed in Chapter 4, different moulding moisture
contents had different compaction efforts applied to achieve dry densities closer to their
MDDs. The weight of a specimen was measured with and without its mould to obtain
its bulk and dry densities. Then, a MP306 sensor was inserted at the middle, and the
VMC% measured for the selected moisture content, following which the specimen was
dismantled and two soils samples taken to measure its GMC% using the oven-dried
method. Similar steps were performed to calibrate the MP306s for the subgrade material
as well. Correlations between the moisture distributions determined from the oven-dried
method and VMC% readings for the base and subgrade materials are shown in Figures
7.11 and 7.12 respectively. A better correlation was obtained for the subgrade material
while, for the base one, calibration was certainly needed to improve the accuracy of the
readings.

Two MP306s were installed at the top of the base to measure the surface moisture, one
buried at the base-subgrade interface to measure the moisture movements between them
and the fourth buried 100 mm from the top of the subgrade, as shown in Figure 7.13.

313
Chapter 7 Characterising Deformation Behaviour using APM Test

119 mm

20 mm 60 mm

Figure 7.9: Moisture probe (MP306) selected for experiments

Soil moisture meter

Data logger

MP306 moisture probe

Figure 7.10: MP306, data logger and SMM

314
Chapter 7 Characterising Deformation Behaviour using APM Test

MP calibration_base material
24 MP 1
MP 2
MP 3
MP 4
Regression line (MP 1)
Regression line (MP 2)
20 Regression line (MP 3)
Regression line (MP 4)
MP306_VSW (%)

16 MP 1 (Y = 0.4613*X +7.7164, R2=0.91)


MP 2 (Y = 0.4933*X +7.1296, R2=0.89)
MP 3 (Y = 0.4821*X +7.3400, R2=0.91)
MP 4 (Y = 0.4849*X +7.4346, R2=0.90)

12

8 12 16 20 24 28 32
Soil sample_VSW (%)

Figure 7.11: MP306 moisture sensors’ calibration curves for base material

MP calibration_subgrade material
60 MP 1
MP 2
MP 3
MP 4
Regression line (MP 1)
50 Regression line (MP 2)
Regression line (MP 3)
Regression line (MP 4)
MP306_VSW (%)

40

30 MP 1 (Y = 1.4015*X - 4.5167, R2=0.97)


MP 2 (Y = 1.3999*X - 4.1749, R2=0.97)
MP 3 (Y = 1.4639*X - 5.6363, R2=0.97)
MP 4 (Y = 1.3920*X - 4.4816, R2=0.98)

20

10

10 20 30 40 50
Soil sample_VSW (%)

Figure 7.12: MP306 moisture sensors’ calibration curves for subgrade material

315
Chapter 7 Characterising Deformation Behaviour using APM Test

Figure 7.13: Locations of installed MP306s

7.2.4 Preparation of Subgrade and Base Layers

The preparation of a model pavement is very labour intensive and requires careful
attention during instrumentation to obtain a quality outcome. Canberra granular base
and subgrade materials were chosen for construction of the base and subgrade layers
respectively and their material properties were presented in Chapter 3. As previously
mentioned, a 600-mm thick subgrade was selected for the model testing. Its required
height was achieved by compacting the materials in several multiple sub-layers and the
selection of their thicknesses important for two reasons: (1) to achieve the target
density; and (2) to obtain a uniform mix. As the subgrade had to be mixed with water in
a concrete mixer to obtain a uniform mix within a reasonable amount of time, it was
essential to mix it in small quantities. Moreover, achieving the target density of MDD
that was established in the Proctor test was only possible to achieve by compacting the

316
Chapter 7 Characterising Deformation Behaviour using APM Test

material in thinner layers. Based on these factors, a sub-layer thickness of 25 mm was


selected, with twenty-four sub-layers forming the 600 mm thick subgrade.

The total amount of material required for base and subgrade is shown in Table 7.4. The
required subgrade material for 600 mm thickness was oven-dried in small quantities and
kept in sealed airtight containers to prevent moisture absorption until the required
amount was collected(due to the high volume, materials had to be oven dried for a week
and kept in seal containers to prevent moisture absorption). Then, they were divided
into twenty four equal layers and mixed in the concrete mixer with water for 15 to 20
minutes to produce a uniform soil mix. Then the mix was poured into the tank and
compacted using a vibratory compactor until the required thickness was achieved.
Inside the tank, marks were placed at every 50 mm, one for each quadrant, to ensure
that each layer achieved the required dry density. The subgrade was constructed in two
phases; on the first day, 400 mm was prepared and, on the next, the remaining 200 mm
plus 150 mm of the base layer were compacted. Figure 7.14 shows the subgrade
construction.

Table 7.4: Requirements for subgrade and base materials

Material required Material required


per layer Total

Subgrade 36.05 (25 mm) 865.3 kg (600 mm)

Base 55.18 (25 mm) 331.1 kg (150 mm)

317
Chapter 7 Characterising Deformation Behaviour using APM Test

Figure 7.14: Subgrade construction in the cylindrical tank

Similarly, the 150-mm thick base layer base was compacted in six equal layers.
However, unlike the subgrade, 1.5% of CF binder was added prior to the addition of 8
% (=OMC) of water. The granular material and binders were dry-mixed for roughly 10
minutes before the water was added and then mixed for another 10 minutes to obtain a
uniform mix.

Most of the sensors were installed while constructing the subgrade and base, with their
placements carefully managed to be, and remain, in their correct locations. Also, during
compaction, additional precautions were taken to prevent any damage to them. An
important concern when installing a number of sensors are the possible disturbances
they could cause to the soil, hence, overcrowding the sensors should be avoided.

While installing the in-depth deformation system, clogging of the soil particles between
the steel tube and rod was avoided by sealing the top opening using duct tape which
otherwise would have resisted free movement of the steel rod. The bottom disk was

318
Chapter 7 Characterising Deformation Behaviour using APM Test

placed on a flat, firm surface during compaction and its position and verticality
constantly checked and adjusted. Similarly, as previously explained, EPCs and MPs
were installed in their respective locations.

After compaction, 28 days were allowed for curing before testing as this is more widely
accepted than a short curing period. Curing is important for any stabilised material as,
during curing, the stabilisers reacts with water and bind the granular particles together
which provides greater strength. In Chapter 5, the importance of curing was well
explained in the context of durability as it was found that, even for a low percentage of
binder (1.5%), the effects of curing could be significant. Curing of most of those
specimens was performed by keeping them in the curing room at 95% relative humidity
and 230C. As this method of curing was not possible for APM testing, in this test, air
curing was performed during which no-excess supply of water added to the stabilised
materials. However, rubber mats were placed on top of the base layer to cover the whole
area and prevent an excessive amount of moisture being lost to the environment.

The hydration of cement requires a w/c ratio of 0.25 which is 0.4% of water to 1.5% of
binder. As previously mentioned, the material was compacted with 8% water, which is
plentiful for 1.5% of CF binders for the reaction. The curing of normal concrete or
stabilised pavements should not be confused with that of a lightly stabilised material
because, in the former, an excess supply of water during curing is essential to absorb
heat of hydration to avoid thermal cracking. Therefore, the curing method used in this
experiment was acceptable. The moisture probe readings also suggested that the
material lost very little moisture during the curing period, as is discussed in the results
section.

Also, the measured UCS of batch mix indicated that the selected method of curing was
acceptable. It’s good practice to test the UCS of a batch mix as it could possibly go
wrong because of the enormous quantity involved. Therefore, after completing mixing,
the material required to prepare a UCS specimen was separated from the mix. Three
specimens were prepared for three different batch mixes and cured at room temperature
in a sealed container to prevent moisture loss which was identical to the curing method
used for the AMP test. The UCS of specimens of the same material obtained from this

319
Chapter 7 Characterising Deformation Behaviour using APM Test

and a standard UCS test in which they were cured in a curing chamber at 95% RH and
230C are shown in Table 7.5. The UCS obtained from the APM test specimens showed
slightly higher values than the specimens under standard testing procedures which could
be because of their slightly different moisture contents. Therefore, these evidence
suggested that using air curing for APM testing is acceptable.

Table 7.5: Comparisons of UCS obtained from two different curing procedures

Standard UCS test APM test batch mixes


Sample UCS (MPa) Layer Sample UCS (MPa)
1 2.21 1 1 2.57
2 2.34 3 2 2.91
3 2.32 5 3 2.08
Average 2.29 2.52

7.2.5 Data acquisition

The installed deformation measurement system, pressure sensors and load output from
the Instron load cells in the pavement model were connected to a national instrument
box with different types of modules for different types of input; for example, the
pressure sensors and load sensors had to be connected to different modules. A signal
conditioner in the national instrument box converted the voltage pulses and input them
into a data acquisition program called LabVIEW which generated .dat raw data files
which were then converted to .xlsx format before being processed using MATLAB to
determine the peak values. The frequency of acquisition of the sensor readings had to be
sufficiently high to capture the peak values but not too high as to create a huge data file
which would be very difficult to process. The moisture measurements from the MP306s
were acquired using its own data logger which had a micro-storage device for storing
files.

320
Chapter 7 Characterising Deformation Behaviour using APM Test

7.2.6 Testing

After 28 days of curing, the testing arrangement for loading the pavement model began
with aligning the loading actuator with the centre of the base on the same axis. As the
model was expected to be applied with symmetrical loadings, it was essential to align
the actuator right above the centre of the pavement. To avoid any ‘rocking’ effects that
could potentially arise during loading, a master flow mix was placed between the steel
beams, where the whole testing tank rested, and the floor. As the master flow would
settle while it gained strength, sufficient time (4 to 6 hours) was needed to avoid
differential settlement.

The next step was to set up the loading plate and LVDTs. The loading plate had to be
pasted to the base layer on a perfectly flat surface to avoid any lateral movement while
subjected to cyclic loading. As LVDTs are mounted on the loading plate measuring
vertical deformation, lateral movements of it may cause erroneous measurements. The
master flow mix was used to stick the base and loading plate together and build the
plate in a horizontal plane.

The grid system made of aluminium shown in Figure 7.15 was used to support all the
eight LVDT holders, into which the LVDTs were clamped, that were mounted on top of
the disks of the in-depth deformation measurement system, with the steel plates attached
to/with the loading plate. The LVDTs’ initial values were adjusted to ensure that their
measurement ranges were sufficiently high to read the maximum permanent
deformations.

After ensuring that the data acquisition program received signals from all the sensors,
stress-controlled cyclic loading was applied using a sinusoidal pulse with a frequency of
2 Hz and rest period of 0.5 second. As the control load control system did not have a
harversine load-shaped pulse, a sinusoidal pulse with a rest period was selected, as
shown in Figure 7.16, and typical outputs from the loadings and LVDTs are shown in
Figure 7.17. The initial intention was to apply a loading frequency of 10 Hz to match
the current Mr laboratory loading rate but the highest rate the selected loading frame
assembly could handle was 2 Hz.

321
Chapter 7 Characterising Deformation Behaviour using APM Test

As shown in Table 7.6, four cyclic loading pressures, 750 kPa, 1000 kPa, 1250 kPa and
1500 kPa were chosen. A minimum pressure of 50 kPa was maintained in all four stress
sequences to ensure that there was always contact between the loading assembly and
loading plate which was essential to keep them in the same axis. Approximately 80,000
load cycles were applied for the first three load sequences and 700,000 for the last
before terminating the test. Then, the raw data collected was processed to study the
material’s behaviour which is discussed in the results section.

Figure 7.15: Aluminium grid for supporting LVDT holders

322
Chapter 7 Characterising Deformation Behaviour using APM Test

Maximum
cyclic stress

Rest period of 0.5 s

Minimum cyclic
stress (50 kPa)

Figure 7.16:Typical loading cycle with rest period

Figure 7.17: Typical outputs showing LVDT readings and cyclic loadings

323
Chapter 7 Characterising Deformation Behaviour using APM Test

Table 7.6: Cyclic loading maximum and minimum pressures

Maximum Minimum
Cyclic stress Maximum Number of
cyclic load pressure
sequence pressure (kPa) cycles
(kN) (kPa)
1 19.94 750 50 80,000

2 26.59 1000 50 80,000

3 33.24 1250 50 80,000

4 39.89 1500 50 700,000

7.3 Repeated Load Triaxial Test

As previously mentioned, APM tests are primarily used to investigate the rutting
behaviours of base and subgrade materials. As they require a huge investment and
months to perform, relating their outcomes to those from laboratory RLT testing is very
useful. RLT testing is commonly used to characterise the rutting behaviour of a
pavement material but has limited value unless verified against field or large-scale data.
Therefore, in this study, the outcomes from APM and RLT laboratory testing were
compared adding some data gathered from a finite element study.

For the results comparison, data from two RLT tests of a similar material to that of
APM (CM), compacted with the same moisture content of 8% and cured for 28 days
were used. In both tests, attaining a MDD close to 100% was ensured by compacting the
required volumes. As detailed descriptions of the laboratory procedures for RLT testing
are described in Chapter 4, they are not repeated here. The deviatoric stresses and
confining pressures used for the RLT tests are tabulated in Table 7.7. RLT test 1 was
presented in Chapter 4 for the same CM with a similar curing age (28 days). It can be
recalled that, in the chapter 4, a repeated cyclic loading frequency of 10 Hz was used.
However, RLT test 2 was performed at a 2 Hz frequency which matched the rate used in
the APM test. Although the frequency was different in RLT test 1, the time duration of
a cycle was kept at 1 second in both the RLT and APM tests. The other differences
between the RLT and APM tests were (1) the method of curing and (2) number of load

324
Chapter 7 Characterising Deformation Behaviour using APM Test

cycles. Also, although the RLT specimens were cured in the curing chamber in contrast
air curing of APM, differences could be ignored, as explained in section 7.2.4. The
numbers of load cycles were limited to 5000 and 10,000 in RLT tests 1 and 2
respectively. In the last part of the results section, the results obtained from all these
tests are analysed.

Table 7.7: RLT testing program

Confining
Deviatoric Number of
Sequence pressure
stress (kPa) cycles
(kPa)
1 250 25 5,000
2 350 40 5,000
Test 1
3 450 50 5,000
4 500 70 5,000
1 450 75 10,000
2 600 100 10,000
Test 2
3 900 150 10,000
4 1500 250 10,000

7.4 Results and Discussion

7.4.1 Accelerated Pavement Model Test Results

The APM test results were analysed to obtain the permanent and resilient behaviours of
each stress sequence and characterise their responses. The resilient responses are later
used to back-calculate the modulus in finite element modelling. Also, the moisture and
pressure measurements collected during testing are discussed in the following sections.

325
Chapter 7 Characterising Deformation Behaviour using APM Test

7.4.1.1 Permanent Deformation Responses

Figure 7.18 shows the permanent deformations measured varying with the number of
loading cycles for all four stress sequences at the top of the base (surface deformation)
and top of the subgrade. Their accumulation rates of permanent deformation were
higher in the first 10,000 cycles and then decreased to lower values with increasing
numbers of load cycles. As the surface deformation measured at the top of the base
represents the deformation of both the base and subgrade, as shown in Figure 7.18, the
subgrade deformation was substantial and the difference between these two
measurements gives the deformations of the base layer which is shown in Figure 7.19.
Table 7.8 shows the percentages of deformations in each stress sequence calculated at
the end of 80,000 cycles and indicates that the rut depth in the subgrade was three times
that of the base which is very common in a typical pavement. Although the component
of the wheel load transferred to the subgrade is minimal compared with the pressure
experienced by the base materials, the substantial difference in its elastic modulus
makes the subgrade more vulnerable. Although this explains the reason for the subgrade
being considered more critical than the base, it does not mean that the latter can be
neglected. The deformation characteristics of pavement layers are interdependent. In
this case, the back-calculated elastic modulus (discussed in the next section) showed
that the base was approximately 15 to 20 times higher than the subgrade which resulted
in the subgrade deforming three times more than the base. For different combinations of
the modulus and load values, the deformation characteristics could be different. This is
one of the advantages of APM over laboratory RLT testing as, in the latter, different
pavement layer interactions cannot be simulated.

326
Chapter 7 Characterising Deformation Behaviour using APM Test

Figure 7.18: Surface and subgrade permanent deformations varying with number
of load cycles

Figure 7.19: Base deformations varying with number of load cycles


327
Chapter 7 Characterising Deformation Behaviour using APM Test

7.4.1.2 Resilient Deformation Responses

The vertical resilient responses obtained from APM testing (Figure 7.20) show that the
resilient deformations decreased slightly with the number of load cycles after an initial
increase up to 10,000 load cycles which means that the resilient modulus increased with
the number of load cycles in the latter stages. Also, the resilient deformation increased
with an increasing cyclic loading pressure similar to permanent deformation, as shown
in Table 7.8 and Figure 7.21, subgrade deformation was a much greater proportion of
the total surface resilient deformation. Both the permanent and resilient deformations
show that the responses in the 10,000th load cycle were better representations that can be
used for back-calculations which are discussed later in this chapter.

The test was terminated after approximately 80,000 load cycles except for the last
loading cycle which is less common in APM testing. However, continuing the test, let’s
say up to 1 million cycles, would not have revealed anything new if the applied pressure
was substantially lower than the failure load of the material. The Austroads (2010a)
classification of failure is when the resilient deformation increases to double of its initial
value. The experimental results from this test did not show any sign of an increase in
resilient deformation after 80,000 cycles but, in fact, it decreased with the number of
cycles.

Moreover, graphs relating the permanent strain to the permanent strain per cycle are
likely to provide more insight into the material’s behaviour. As shown in Figure 7.22,
for all four stress sequences, the material showed a higher strain rate per cycle for the
first few thousand load cycles but reduced to a very low level after 80,000 cycles.
According to Werkmeister et al. (2005), this is Range B behaviour, i.e., the material will
fail after reaching its design limit. Therefore, as continuing the test until failure was not
feasible, it could be terminated at a relatively lower number of load cycles and a new
stress level (higher than the previous one) introduced; for example, in this test, although
the last test sequence was continued up to 700,000 cycles, the resilient and permanent
deformation responses did not reveal any additional information. Therefore, a multi-
stage test would be likely to save a lot of time as the number of tests required could be
significantly reduced. Collecting data and classifying responses regarding resilient and

328
Chapter 7 Characterising Deformation Behaviour using APM Test

permanent deformation behaviours while the test was running could help to decide
whether to terminate it for a particular cyclic load.

Table 7.8: Proportion of deformation of base and subgrade

Permanent deformation Resilient deformation


% of % of
Stress % of Base % of Base
Subgrade Subgrade
sequence deformation deformation
deformation deformation
1 73 27 78 22
2 78 22 79 21
3 72 28 77 23
4 72 28 76 24

Resilient deformation_top sufrace and subgrade


Top surface (σa_750 kPa) Subgrade (σa_750 kPa)
1 Top surface (σa_1000 kPa) Subgrade (σa_1000 kPa)
Top surface (σa_1250 kPa) Subgrade (σa_1250 kPa)
Top surface (σa_1500 kPa) Subgrade (σa_1500 kPa)

0.8

0.6

0.4

0.2

0 20000 40000 60000 80000


Number of loading cycles

Figure 7.20: Surface and subgrade resilient deformations varying with number of
load cycles

329
Chapter 7 Characterising Deformation Behaviour using APM Test

Figure 7.21: Base resilient deformation varying with number of load cycles
Vertical permanent strain rate (10-3/load cycle)

Figure 7.22: Vertical permanent strain rate varying with vertical permanent strain

330
Chapter 7 Characterising Deformation Behaviour using APM Test

7.4.1.3 Moisture Content Readings

As mentioned in section 7.2.3, the moisture contents in four locations, two on top of the
base layer, one at the interface (buried in the subgrade) and another 100 mm from the
top of the subgrade, were measured. The moisture content measurements began after 28
days curing that is at the beginning of the first loading stage and continued until the end
of testing which was approximately 21 days.

As shown in Figure 7.23, at the beginning of testing, moisture contents of 6.3% to 6.4%
were recorded on top of the base layer. It can be recalled that the CM was compacted
with its OMC of 8%, and it was found from other element testings that it retained a 7%
to 7.2% moisture content after 28 days of curing in the standard curing chamber. In
APM testing, curing was performed under ambient conditions which could have
resulted in an additional moisture loss of 0.7% to 1.2%.

Moreover, as shown in Figures 7.23 and 7.24, the base recorded a decline in moisture
content of around 0.2% to 0.4%, and the base and subgrade interface (top of the
subgrade) showed a reduction of 4.5% from their initial values over the 21 days of
testing. Moisture movements in the base and subgrade interface could be for the
following three reasons.

(1) Capillary action - surface tension between the soil particles and air in the soil matrix
creates capillary pores and pulls the water which could have happened at the
base/subgrade interface.

(2) Evaporation of water – could have occurred from the base which was exposed to the
environment.

(3) Gravitational water flow - the water that moves in or out of a soil by gravity depends
on the permeability of the material and takes place only if the soil is fully saturated.
This was not possible because neither the base nor subgrade was fully saturated but
compacted at their OMCs.

The decrease in moisture content at the top of the subgrade could have been due to
moisture movement through capillary action from the subgrade, which had plenty of

331
Chapter 7 Characterising Deformation Behaviour using APM Test

water, to the base which was relatively dry. It can be recalled that, in the tube suction
test in Chapter 4, the moisture content of the same material exceeded 7% after 10 days
of capillary infiltration. Therefore, there was a good possibility of moisture moving
from the subgrade to base but, as previously mentioned, moisture measurements showed
that the base materials lost 0.2% to 0.4%. This could have been due to an excess
evaporation of water caused by the base being exposed to air during testing which could
have negated the infiltrated water. It is important to note that moisture measurements of
the base material were actually representative of the top half of the base layer because
the moisture probes were reading the moisture contents at depths of 60 to 70 mm as the
needle of each MP was 60 mm and it was planted/ 10 mm from the surface. If the
infiltrated water did not reach the top half (to the level of the MP306s), it would not
have been accounted for in the moisture readings. Moreover, moisture readings at a 100
mm depth from the subgrade showed only 0.5% to 1% increases, means that it was
unaffected by any moisture movements.

In summary, as the moisture readings gave some indications of moisture movements,


the MP306s could be used to measure the moisture contents in APM testing by improve
precision using soil-specific calibrations.

6.6
Moisture reading_top surface
MP 2
MP 1

6.4
Moisture content (%)

6.2

5.8

0 100 200 300 400 500


Time (hours)

Figure 7.23: Moisture content in top surface varying with time

332
Chapter 7 Characterising Deformation Behaviour using APM Test

33
Moisture reading_top surface
MP 100 within the subgrade
MP Interface
32

31
Moisture content (%)

30

29

28

27

0 100 200 300 400 500


Time (hours)

Figure 7.24: Moisture contents at interface and 100 mm in subgrade varying with
time

7.4.2 Finite Element Modelling and Back Calculation of Mechanical


Properties

7.4.2.1 Introduction

As far as the pavement design is concerned, the traditional pavement design methods
are empirical in nature and have been based on some full-scale pavement loading tests.
However, as each pavement configuration, and its loadings and environmental
conditions are generally different, the extrapolation of results, most often lead to
inaccurate designs. On the other hand, mechanistic design is based on a rational
approach which considers the stress-strain behaviours of the pavement materials.
Currently a design comprising both mechanical and empirical elements (M-E) is in
practice.

333
Chapter 7 Characterising Deformation Behaviour using APM Test

One major mechanistic aspect of M-E design in the pavement analysis process is to find
the critical stress and/or strain in any of the materials of the pavement structure. The
Pavement analysis can be done using different techniques and each has merits and
demerits. It was initially started with the simple elastic layer model, but now, detailed
models are being employed and analysed using numerical methods (Burmister 1945;
Kim 2007). However, still design standard Austrods 2010 recommends the use of a
multi-layer program CIRCLY for the pavement design due to its simplicity.

To accommodate non-linear elastic behaviour of pavement materials, specific analysis


programs GT-PAVE, ILLI-PAVE, etc., were developed but their inability to implement
user defined material models made them less attractive. The general-purpose finite
element analysis (FEA) programs such as ABAQUSTM, ANSYSTM, ADINATM, etc.,
have been successfully used for the analysis using both user defined and built-in models
(Sukumaran 2004, Uddin 2005, Kim 2007). In these FEA programs, the pavement
layers can be modelled as 3-D, plain strain or axisymmetric. Though 3-D analysis can
be used to represent more precisely the field condition it takes huge amount of
computing hours/resources.

It has been well proved that unbound granular materials and subgrade materials exhibit
non-linear elasto-plastic material behaviour and linear elastic approximation wouldn’t
yield accurate solution. To deal with the non-linearity, different researchers used
different models for subgrade and pavement base/sub-base layers in their FEA models.
For example, Sukumaran (2004) used elasto-plastic Drucker-Prager model to model
clay subgrade and gravel base respectively; Zaghloul (1993) used Drucker-Prager model
for unbound granular base layer and Cam-Clay model for subgrade. As far as the lightly
stabilized material is concerned, not much work has been done to incorporate elasto-
plastic models in FEA model.

A typical pavement analysis and design is performed knowing the mechanical


properties of the material. These mechanical properties are inputted in a finite element
model or multi-layer elastic programs to obtain critical strains. The critical strains are
then used to obtain allowable number of axle. However, there are circumstances where
the material properties remain unknown and need to determine from non-destructive

334
Chapter 7 Characterising Deformation Behaviour using APM Test

test. Particularly in pavement rehabilitation, to estimate the remaining life or to


understand the

Falling weight deflectometer (FWD) is a good example for a non-destructive testing


which utilises the back calculation technique to determine the stiffness of pavement
layers. FWD test involves applying dynamic loads to a pavement surface and measures
vertical deflections at different locations using sensors. The measured deflections and
the load are then used to determine the modulus of elasticity using back calculation.
Several methods are available to back calculate the mechanical properties of flexible
pavement. FWD data bases are being used to arrive at better techniques and algorithms
which could be used for back calculation (Sangghaleh et al. 2014).

Back calculation in a pavement system is a process of determining the moduli of


materials using numerical simulations. Back calculated moduli are used to determine
remaining life of a pavement. Once obtained layer moduli, pavement analysis can be
performed to obtain critical design parameters such as tensile strain or permanent
deformations which can then be used to arrive at remain life. Back calculation is an
iterative process where different combinations of layer moduli are tried to fit the
deformation responses. It is generally performed using computer programs, however,
manual iteration is possible particularly in-cases where only two layers involved similar
to the involved in this study.

There are different back calculation software are available for determining the elastic
layer moduli. Detail evaluations of six selected programs such as ELCON, ELI-BACK,
ISSEM4, MODCOMP3, MODULUS and WDEF is available in (Program 1993b).
However, in this study back calculations is carried out by performing manual iterative
checks on base and subgrade moduli. The layers were assumed as elastic layers and
different combinations of base and subgrade modulus were tried to match the deflection
obtained from finite element model to the one obtained in the experimental study. As
mentioned earlier, finite element program ABAQUS® was used to generate a model
representative of APM testing.

335
Chapter 7 Characterising Deformation Behaviour using APM Test

7.4.2.2 Development of Finite Element Model

The first step in a FEA of a pavement is selecting a geometry consisting of different


pavement layers and their thicknesses. As previously mentioned, in this FEA, the
pavement configuration used for the APM testing, which has a 150 mm lightly
stabilised base and 600 mm subgrade, was selected as shown in Figure 7.25. The
loading was applied using a 184 mm diameter circular steel plate which had its centre
lying on the top of the centre of the cylindrical tank to ensure that the pavement loaded
symmetrically and the stresses were distributed evenly throughout its layers.

ABAQUSTM, a commercially available FE software, which allows the development of


3-D, plane strain or axisymmetric models, was selected for the numerical analysis. Due
to the double symmetry of the geometry and load about the y axis, as shown in Figure
7.26 and 7.27, an axisymmetric model was chosen, with the thickness of each layer and
loading diameter the same as to those in the APM testings.

Dia = 184 mm

Lightly stabilised granular base 150 mm

600 mm
Subgrade

1175 mm

Figure 7.25: Configuration of APM

336
Chapter 7 Characterising Deformation Behaviour using APM Test

D C
Base

Subgrade
150 mm

Plane of axis symmetry


600 mm
A B

Axis of symmetry

Figure 7.26: Schematic of axisymmetric model selected for finite element modelling

Loading

Part module_Base

Part module_Subgrade

Figure 7.27: Part module of axisymmetric model

337
Chapter 7 Characterising Deformation Behaviour using APM Test

As shown in Figure 7.31, the axisymmetric model was discretised using the element
type C4X4R, a 4-node bilinear axisymmetric quadrilateral with reduced integration.

Figure 7.28: FE mesh used for axisymmetric model

7.4.2.3 Boundary Conditions and Loading

The following boundary conditions were applied to the selected axisymmetric model, as
shown in Figure 7.29.

 The vertical and horizontal movements of the bottom surface were fully
constrained (U1=U2=U3=0).
 The edges were prevented from moving in horizontal directions and allowed to
move only vertical directions (U2=U3=0). In ABAQUS, the ‘symmetric along y
axis’ option was used to define this boundary condition which allowed
measurements of vertical deformations in the y(U2) direction.

338
Chapter 7 Characterising Deformation Behaviour using APM Test

The loading for simulating a tyre pressure is often a constant uniform circular or
rectangular sinusoidal pulse or haversine load because most FEAs of flexible pavements
were based on a two-dimensional (2-D) axisymmetric analysis (Hadi and Bodhinayake
2003; Saad et al. 2005; Schwartz 2002). However, for a comprehensive FEA, 3-D
model could be used with spatially varying tyre pressures (Wang 2011). As the scope of
this study was limited to back-calculation, a constant tyre pressure was considered
which was also better simulating the loading type used in the experimental study. This
pavement configuration was subjected to circular loadings of 750 kPa, 1000 kPa, 1250
kPa and 1500 kPa pressures using a radius of 92 mm, as in the APM tests. The 92 mm
radius covered an area of 0.027 m2 which transferred the load from 20 kN for 750 kPa
to 40 kN for 1500 kPa. A harvesine load pulse at a frequency of 2 Hz was applied at
these selected pressures and the deformations and stress values measured at the peak
pressure.

y(2)

x(1)

Figure 7.29: Boundary conditions used for FE analysis

339
Chapter 7 Characterising Deformation Behaviour using APM Test

7.4.2.4 Back-Calculation Considering Isotropic Elastic Properties

Back-calculations were performed matching the experimental resilient deformations in


the 10,000th cycle obtained from the APM test of each loading pressure to those
obtained from an ABAQUS FEA. As the permanent and resilient deformations attained
a stable state after 10,000 cycles, as discussed in section 7.4.1, the material’s
deformation responses could be well represented by the deformation response obtained
after 10,000 loading cycles.

Initially, a back-calculation using isotropic analysis, was performed by assigning


isotropic elastic properties to both the subgrade and base materials. The Poisson’s ratios
of the base and subgrade materials were kept constant at 0.25 and 0.35 respectively
while their elastic moduli were changed until the FE and experimental curves matched.
As previously mentioned, three deformation readings were measured in the APM
testing, two of which, the surface and interface, were used in the matching process.
After obtaining the elastic moduli by matching two of these deformations, the middle
deformations of both the experimental and FEs were compared and found to be a good
match. Similar steps were followed for all four stress sequences to obtain the elastic
properties shown in Table 7.9. The matching curves obtained using this elastic modulus
for the four stress sequences are shown in Figures 7.30 to 7.37.

The back-calculated elastic moduli of the subgrade did not change considerably
between different cyclic stress values, with those of 42 MPa and 37 MPa obtained for
cyclic loadings of 750 kPa and 1250 kPa respectively. However, those of the base
differed significantly and decreased with increases in the cyclic stress. The maximum
and minimum elastic moduli of 950 MPa and 750 MPa were back-calculated for 750
kPa and 1500 kPa cyclic stresses respectively which clearly demonstrated the
dependency of the elastic modulus on the stress values.

340
Chapter 7 Characterising Deformation Behaviour using APM Test

Matching surface deformations (750 kPa)


Exp. surface 
0.4
Abaqus surface 

0.3

Deformation (mm)

0.2

0.1

0 0.2 0.4 0.6


Time (s)

Figure 7.29: Matching ABAQUS surface vertical resilient deformations and


measured experimental deformations (loading pressure 750 kPa)

0.4 Matching interface, middle deformations (750 kPa)


Exp. surface  Exp. middle 
Abaqus surface  Abaqus surface 

0.3
Deformation (mm)

0.2

0.1

0 0.2 0.4 0.6


Time (s)

Figure 7.30: Matching ABAQUS middle and interface vertical resilient


deformations and measured experimental deformations (loading pressure 750 kPa)

341
Chapter 7 Characterising Deformation Behaviour using APM Test

Matching surface deformations (1000 kPa)


0.6
Exp. surface 
Abaqus surface 

0.4

Deformation (mm)

0.2

0 0.2 0.4 0.6


Time (s)

Figure 7.31: Matching ABAQUS surface vertical resilient deformations and


measured experimental deformations (loading pressure 1000 kPa)

Matching interface,middle deformations (1000 kPa)


Exp. interface  Exp. middle 
0.5 Abaqus interface  Abaqus middle 

0.4
Deformation (mm)

0.3

0.2

0.1

0 0.2 0.4 0.6


Time (s)

Figure 7.32: Matching ABAQUS middle and interface vertical resilient


deformations and measured experimental deformations (loading pressure 1000
kPa)

342
Chapter 7 Characterising Deformation Behaviour using APM Test

Matching surface deformations (1250 kPa)


0.8
Exp. surface 
Abaqus surface 

0.6

Deformation (mm)

0.4

0.2

0 0.2 0.4 0.6


Time (s)

Figure 7.33: Matching ABAQUS surface vertical resilient deformations and


measured experimental deformations (loading pressure 1250 kPa)

Matching interface,middle deformations (1250 kPa)


Exp. interface  Exp. middle 
Abaqus interface  Abaqus middle 

0.6
Deformation (mm)

0.4

0.2

0 0.2 0.4 0.6


Time (s)

Figure 7.34: Matching ABAQUS middle and interface vertical resilient


deformations and measured experimental deformations (loading pressure 1250
kPa)

343
Chapter 7 Characterising Deformation Behaviour using APM Test

Matching surface deformations (1500 kPa)


1
Exp. surface 
Abaqus surface 

0.8

Deformation (mm)
0.6

0.4

0.2

0 0.2 0.4 0.6


Time (s)

Figure 7.35: Matching ABAQUS surface vertical resilient deformations and


measured experimental deformations (loading pressure 1500 kPa)

Matching interface,middle deformations (1500 kPa)


Exp. interface  Exp. middle 
Abaqus interface  Abaqus middle 
0.8

0.6
Deformation (mm)

0.4

0.2

0 0.2 0.4 0.6


Time (s)

Figure 7.36: Matching ABAQUS surface vertical resilient deformations and


measured experimental deformations (loading pressure 1500 kPa)

344
Chapter 7 Characterising Deformation Behaviour using APM Test

Table 7.9: Back-calculated elastic moduli matching resilient deformations

Cyclic loading
Base elastic modulus Subgrade elastic
Loading sequence pressure
(MPa) modulus (MPa)
(kPa)
1 750 950 MPa 42
2 1000 900 MPa 39
3 1250 850 MPa 37
4 1500 725 MPa 39

7.4.2.5 Back-Calculation Considering Cross-anisotropic Elastic Properties

Pavement materials are generally not isotropic in nature mainly because their layers are
compacted in one plane (the x-y plane in Figure 7.38) and stacked on each other in the
orthogonal direction (z) which makes them cross-isotropic. As the response of such a
material in either the x or y direction is the same but differs in the z direction, Austroads
(2010a) suggests the use of cross-anisotropy for subgrade, unbound and modified
granular materials. Therefore, in this analysis, subgrade and lightly stabilised granular
materials are defined as cross-isotropic and their resilient modulus values back-
calculated using a similar approach to that in the previous section.

To define cross-anisotropy in the elastic range, the five parameters needed were two
elastic moduli, two Poisson’s ratios and one shear modulus. The Poisson’s ratio was
assumed to be the same in both directions and, similar to the isotropic analysis, 0.25 and
0.35 were selected for the base and subgrade respectively. The ratio of the vertical to
horizontal moduli was assumed to be 2 and the shear modulus calculated based on
equation 7.1. As shown in Table 7.10, all these parameters were inserted in ABAQUS
the ‘engineering constant’ tab in the elastic category for defining a cross-anisotropic
material.

In this analysis, for a particular vertical modulus, the horizontal and shear moduli
needed to be calculated and inserted manually for each iteration. Similar to the isotropic

345
Chapter 7 Characterising Deformation Behaviour using APM Test

analysis, matching of the surface and interface resilient deformations of the


experimental and FE models was performed. The back-calculated parameters are shown
in Table 7.11 in which the vertical elastic modulus of the base reduced with an increase
in the cyclic load and, similar to the isotropic analysis, no such trend was noticed for the
subgrade material.

Figures 7.39 and 7.40 show comparisons of the moduli obtained from the isotropic and
cross-anisotropic analyses of the base and subgrade respectively. Although the base
material’s vertical elastic moduli obtained from the cross-anisotropic analysis were
marginally higher than those from the isotropic analysis, the subgrade showed
considerable increase. In terms of the horizontal moduli, in the isotropic analysis, they
remained the same and were half in the cross-anisotropic analysis. Therefore, more
significant differences were noticeable in the horizontal than vertical moduli which
would likely have a considerable effect on the deformation and stress values.

Figure 7.37: Layered compaction makes cross-anisotropic pavement layers

346
Chapter 7 Characterising Deformation Behaviour using APM Test

Ex E y
G xy  Eq. (7.1)
E x (1  2 xy )  E y
where:
Gxy = Shear modulus (Pa)
E = Elastic modulus (Pa)

n= Poisson’s ratio

Table 7.10: Input parameters required for cross-anisotropic analysis

Material nxy/nxz/nyz Eh Ev Gf(x,y,z)


The ratio of vertical to
Lightly stabilised Equation
0.25 horizontal modulus was
base layer 7.1
assumed to be 2
The ratio of vertical to
Equation
Subgrade 0.35 horizontal modulus was
7.1
assumed to be 2

Table 7.11: Back-calculated cross-anisotropic properties

E1/E3 E2
Material (MPa) (MPa)
V G 12 G 13 G 23

Base 475 950 0.25 380 136 238


750 kPa
Subgrade 25 50 0.35 19 7 11

Base 475 950 0.25 380 136 238


1000 kPa
Subgrade 22.5 45 0.35 17 6 10

Base 450 900 0.25 360 129 225


1250 kPa
Subgrade 21 42 0.35 16 6 10

Base 375 750 0.25 300 107 188


1500 kPa
Subgrade 22.5 45 0.35 17 6 10

347
Chapter 7 Characterising Deformation Behaviour using APM Test

Isotropic
1000 Cross-anisotropic

800

Back-calculated elastic modulus


600

400

200

0
Stress Stress Stress Stress
sequene 1 sequene 2 sequene 3 sequene 4
Stress sequences

Figure 7.38: Comparison of moduli of base materials

Isotropic
50 Cross-anisotropic

40
Back-calculated elastic modulus

30

20

10

0
Stress Stress Stress Stress
sequene 1 sequene 2 sequene 3 sequene 4

Stress sequences

Figure 7.39: Comparison of moduli of subgrade materials

348
Chapter 7 Characterising Deformation Behaviour using APM Test

7.4.2.6 Comparison of Vertical Pressure Distributions

As previously discussed, four EPCs were used to measure the pressure at four different
locations. Figure 7.41 shows typical responses measured at the base-subgrade interface
using EPC 1 which demonstrates that the pressure remained stable after increasing its
value in the initial 5,000 to 10,000 loading cycles. An increase in the cyclic loading
stress from 750 kPa to 1500 kPa resulted in increase in the vertical pressure values, with
those measured in the middle of the layer higher (as expected) for the first three
sequences but registered lower in the last loading stage. EPC 2 could have gone out of
the plane because of the larger deformation in the last loading stage and produced
erroneous results. Therefore, measuring pressure values using an EPC when a material
undergoes a large deformation should be avoided. These measured values were used to
plot the pressure distribution along the depth of the base layer and, as shown in Figure
7.42, the pressures obtained from the middle EPC should have shown higher values.

EPC 1_bottom
250 s a_750 kPa s a_1250 kPa
s a_1000 kPa s a_1500 kPa

200
Pressure measured EPCs (kPa)

150

100

50

0 20000 40000 60000 80000


Number of loading cycles

Figure 7.40: Variations in EPC measurements at base-subgrade interface with


number of cycles

349
Chapter 7 Characterising Deformation Behaviour using APM Test

Top of the base layer

Figure 7.41: Vertical pressure distributions measured using EPCs at different


depths

The similar vertical pressure distributions in the middle obtained from the numerical
isotropic and cross-anisotropic analyses are shown in Figures 7.43 and 7.44 respectively
and Figure 7.45 shows the vertical pressure distribution obtained for 1500 kPa axial
pressure. The cross-anisotropic pressure distributions skewed closer to the surface than
the isotropic pressure distribution which could have been due to the differences in their
horizontal moduli.. The pressure values obtained at the top, middle and interface of two
of these FE models and measured using EPCs are shown in Table 7.12. The vertical
pressure values obtained from isotropic and cross cross-anisotropic analysis varied
significantly. Pressures measured at the bottom surface in the experiment and obtained
from the cross-anisotropic analysis matched very well, except for the first load
sequence, while errors in the middle EPC readings could clearly be seen by comparing
them with those obtained from the FE models.

350
Chapter 7 Characterising Deformation Behaviour using APM Test

Top of the base layer

Figure 7.42: Vertical pressure distributions obtained from isotropic analysis

Top of the base layer

Figure 7.43: Vertical pressure distributions obtained from cross-anisotropic


analysis

351
Chapter 7 Characterising Deformation Behaviour using APM Test

Table 7.12: Comparison of stress values at interface and mid-depth of base


material

Base subgrade interface Mid-depth of base


From From
From From
From cross- From cross-
isotropic isotropic
EPC anisotropic EPC anisotropic
analysis analysis
analysis analysis
Sequence
42 51 91 90 370 463
1
Sequence
102 67 115 109 493 614
2
Sequence
137 84 143 143 616 767
3
Sequence
202 111 193 163 740 933
4

Figure 7.44: Vertical pressure distributions obtained from cross-anisotropic


analysis for 1500 kPa cyclic loading

352
Chapter 7 Characterising Deformation Behaviour using APM Test

7.4.2.7 Comparison of Horizontal Pressure Distributions

As previously explained, in the APM testing, horizontal pressures were measured by


EPCs at 184 mm from the centre (twice the length of the loading plate) and the
boundary wall. Both these EPCs were placed for measuring horizontal pressure readings
at the mid-depth of the base layer (points X and Y in Figure 7.46). Similar to the
vertical pressures, the horizontal pressures increased with the number of load cycles and
reached stable values after 10,000 cycles.

Measuring the pressures at the boundary wall (EPC 4) was important because of a
concern regarding the arrangement of the circular tank, that is, the possible confining
pressure emanating from the boundary wall. As previously mentioned, the base and
subgrade materials were compacted within the reinforced concrete cylinder and
boundary wall of the tank which could potentially act as a barrier and prevent the
material (particularly that adjacent to the boundary wall) from deforming in lateral
directions. In contrast, in an in-service pavement, its materials are allowed to strain in
all directions under wheel loadings. Therefore, pressure measurements were important
for determining any confinement caused by the cylindrical tank and. As shown in Table
7.13, those measured pressures at the boundary were only 3 to 4 kPa for all stress
sequences. Therefore, for the selected diameter of the loading plate and the stress values
used for loadings, the effect of the boundary wall could be neglected. This table also
shows that the horizontal pressures obtained from the cross-anisotropic analysis were
less than those at the boundary wall.

Moreover, the horizontal pressure measured in the APM test and the value obtained
from the cross-anisotropic analysis at 184 mm from the centre agreed well for a higher
cyclic stress (sequence 4) but showed a maximum 40% difference in sequence 1, with
all the experimental readings less than the results from the FEA.

353
Chapter 7 Characterising Deformation Behaviour using APM Test

Table 7.13: Horizontal pressures at boundary wall and 184 mm from centre

Pressure at 184 mm from


Pressure at boundary wall
centre
Experiment Cross- Experiment Cross-
(EPC 3) anisotropic (EPC 3) anisotropic
(kPa) analysis (kPa) (kPa) analysis (kPa)
Sequence 1 21 36 3 1
Sequence 2 34 48 3 2
Sequence 3 48 59 4 2
Sequence 4 66 73 4 1

Although previously, the stress values at two particular elements (the locations at which
EPCs were installed) were discussed, the horizontal pressure distribution along the
depth of the base layer provided a better understanding of the distribution pattern.
Figures 7.47 and 7.48 show pressure distributions obtained at B-B, which was 184 mm
from the middle, and C-C, which was at the boundary wall, from the cross-anisotropic
analysis for 1500 kPa.

EPC 4
EPC 3
A B C

150 mm
X Y

A B C
Lightly stabilised granular base
578.5 mm

Subgrade

Figure 7.45: Locations at which horizontal pressure distributions obtained from


cross-anisotropic analysis

354
Chapter 7 Characterising Deformation Behaviour using APM Test

From the horizontal pressure distributions at B-B for all four stress sequences, it was
noted that the pressure changed its direction at around a depth of 110 mm. This was not
surprising as the lightly stabilised material was expected to have a tensile capacity
because of its increased stiffness. The stress patterns were very similar to those obtained
from a typical flexural beam test in which the top and bottom parts from the neutral axis
would be subjected to compressive and tensile forces respectively. It can be recalled that
the pressure measured using EPC 3 at the mid-depth simply represented the pressure at
that particular location and not the distribution of the vertical axis passing through that
point.

Similarly horizontal distribution was obtained at the boundary wall is shown in Figure
7.48 and Figure 7.49 shows the distribution for the whole base layer. Unlike in B-B, the
top half subjected to tension and the bottom subjected to compression and at the middle
the pressure is almost zero. As previously discussed, the EPCs readings were obtained
at the middle of the layer which measured 3 to 4 kPa as they lied at the neutral axis. But
the FE analysis showed that there is actually a pressure distribution.

Stress sequence 1 (750 kPa)


0 Stress sequence 2 (1000 kPa)
Stress sequence 3 (1250 kPa)
Stress sequence 4 (1500 kPa)
-20
Dpeth of the stabilised layer (mm)

-40

-60

mid-depth
-80

-100

-120

-140

-160

-80 -40 0 40 80
Horizontal pressure obtained from cross-anisotropic
analysis at B-B (kPa)

Figure 7.46: Horizontal pressure distributions obtained at B-B from cross-


anisotropic analysis

355
Chapter 7 Characterising Deformation Behaviour using APM Test

0
Stress sequence 1 (750 kPa)
Stress sequence 2 (1000 kPa)
-20 Stress sequence 3 (1250 kPa)
Stress sequence 4 (1500 kPa)

Dpeth of the stabilised layer (mm)


-40

-60
mid-depth
-80

-100

-120

-140

-160

-100 0 100
Horizontal pressure obtained from cross-anisotropic
analysis at C-C (kPa)

Figure 7.47: Horizontal pressure distributions obtained at boundary wall from


cross-anisotropic analysis

Tensile region

Neutral axis
Compressive region

150 mm

Compressive region

Tensile region Base and subgrade


interface

587.5 mm

Figure 7.48: Horizontal pressure distributions in base layer obtained from cross-
anisotropic analysis for 1500 kPa cyclic loading

356
Chapter 7 Characterising Deformation Behaviour using APM Test

7.4.3 Repeated Load Triaxial Test Results

As mentioned in section 7.3, two RLT tests were performed matching most of the
conditions prevailed in the APM test and the results obtained from the tests are
discussed in the next section.

7.4.3.1 Permanent and Resilient Response of Repeated Load Triaxial Test

Figure 7.50 shows a typical permanent and resilient deformation response obtained
from RLT for each stress sequence. This particular figure obtained for second RLT last
stress sequence. After analysing the data, it was found that first sequence test results
from both tests did not produce consistence outcome with the other sequences, hence
needed to be eliminated from the remaining calculations.

0.28 0.17

0.24

0.16
Permanent deformation (mm)

Resilient deformation (mm)


0.2

0.16 Permanent deformation 0.15


Resilient deformation

0.12

0.14

0.08

0.04 0.13

0 2000 4000 6000 8000 10000


Number of loading cycle

Figure 7.49: Typical RLT permanent and resilient deformation response

357
Chapter 7 Characterising Deformation Behaviour using APM Test

7.4.3.2 Selecting Stress Values from Experiments

The permanent deformations obtained from the RLT tests showed good correlations
with the number of loading cycles. The model shown in equation 7.2 relating
accumulated permanent deformation with number of load cycles was used to obtain the
regression parameters which are tabulated in Table 7.14. As shown in Figure 7.51, the
obtained regression parameters a and b and bulk stress have strong exponential and
power relationships respectively. If a bulk stress for a particular stress combination is
known, a and b can be calculated and then used to obtain the permanent deformation at
a given number of load cycles or the accumulated permanent deformation after a
number of load cycles.

Applying this outcome to the APM tests was not straightforward because, unlike the
RLT tests, their confining or vertical pressure distributions are not uniform. Therefore,
as explained in the next section, the equivalent pressure distribution was obtained as

ep = aNb Eq. (7.2)

where:

ep = Accumulated permanent deformation (mm)


N = Number of load cycles
a, b = Regression parameters

Table 7.14: RLT test regression coefficients

Maximum
Confining
deviatoric Bulk stress
Test pressure R2 a b
stress (kPa)
(kPa)
(kPa)
592 100 892 0.84 0.0002 0.4245
Test 1 794 150 1244 0.80 0.0012 0.2964
1500 250 2250 0.93 0.0245 0.2349
357 40 480 0.98 0.0002 0.6953
Test 2 447 50 597 0.97 0.0002 0.7078
501 69 708 0.96 0.0009 0.5822

358
Chapter 7 Characterising Deformation Behaviour using APM Test

0.8 0.03

Y=5E-5 e0.0027X
R2 =0.88

Regression parameter a
Regression parameter b

0.6 0.02

Y=98.84 (X)-0.794
R2=0.95

0.4 0.01

0.2 0

0 500 1000 1500 2000 2500


Bulk Stress (kPa)

Figure 7.50: Regression parameters a and b varying with bulk stress

7.4.3.3 Selection Stress Values from ABAQUS

From the ABAQUS model, an imaginary RLT specimen size cylindrical shape was
considered, as shown in Figure 7.52, and the vertical and horizontal pressure
distributions at A-B and C-D respectively obtained. The cylinder’s 50 mm width
matched the radius of the RLT specimen and the thickness (height) of its base 150 mm
is same as height used for measuring deformation in the RLT specimen.

The vertical and horizontal pressure distributions obtained from the cross-anisotropic
analysis are shown in Figures 7.53 and 7.54 respectively. As they were not uniform, a
method for obtaining uniform pressure distributions that could be used with the RLT
test data was required. A vertical non-linear pressure distribution was approximated and
considered a linear distribution between the top and bottom pressure readings and
converted to a uniform pressure distribution by equating the area. As the horizontal

359
Chapter 7 Characterising Deformation Behaviour using APM Test

pressure distribution changed from compression to tension with an increasing depth, the
resultant area was considered to obtain uniform pressure distribution.

Table 7.15 shows the equivalent uniform stress values obtained for four stress
sequences which were then used to calculate, bulk stress and the respective a, b using
the relationship established in the previous section. Moreover, using equation 7.2, for
the APM test permanent deformation data, a and b regression parameters was obtained
similar to the RLT test. Figures 7.55 shows the relationship between obtained regression
parameters and it can be seen that they are related very well with the regression
coefficient showing 0.98 and 0.84.

The objective of this analogy is to demonstrate that the RLT and FEM can be used to
predict the permanent deformation outcome from APM after establishing relationships
or models using all three of them. The quality of the predictions can be improved by
gathering more RLT and APM test data and further improving the established
relationships. However, the limitation of the analysis is that it may not perform well for
Range C type behaviours in which the equation 7.2 is less predicative. As most of the
pavements exhibits Range A or B, it is likely to cover most of the stress combinations.

Another important remark is that the simulations (confining and deviatoric stresses) take
place in the RLT is very different from the in-situ pressure distribution and RLT does
not account for the pavement layer interaction as well which explains the superiority of
APM tests over RLT tests. Having said that, the analogy similar to the one presented
above with certain meaning full assumptions, still can be used to obtain approximate
predictions.

360
Chapter 7 Characterising Deformation Behaviour using APM Test

Equivalent area of triaxial specimen

A C

150 mm Base

B D
50 mm

Subgrade

Figure 7.51: An imaginary RLT specimen size cylindrical in the accelerated


pavement model

Figure 7.52: Vertical pressure distribution along A-B

361
Chapter 7 Characterising Deformation Behaviour using APM Test

Mid-depth

Figure 7.53: Horizontal pressure distribution along C-D

Table 7.15: Representative uniform pressure distributions from cross-anisotropic


analysis and comparison of regression coefficients

Rep.
Rep. Bulk a b
Sequen confining
deviatoric stress a b (APM (APM
ce pressure
stress (kPa) (kPa) test) test)
(kPa)

1 401 100 725 0.00008 0.55 0.001 0.366

2 534 132 960 0.00025 0.42 0.007 0.247

3 667 162 1191 0.00072 0.34 0.015 0.256

4 806 196 1444 0.00230 0.29 0.034 0.227

362
Chapter 7 Characterising Deformation Behaviour using APM Test

Regression parameter b (APM test)


0.1 0.2 0.3 0.4

0.0025 0.6

Y=0.0691X-0.0001
R2=0.98
0.002
0.5

Regression parameter b
Regression parameter a

0.0015 mid-depth Y=1.6551X-0.0551


R2=0.84
0.4

0.001

0.3
0.0005 a
b
Regression line_ a
Regression line_b

0 0.2

0 0.01 0.02 0.03 0.04 0.05


Regression parameter a (APM test)

Figure 7.54: Relationship between regression coefficients

7.5 Summary

This chapter focussed on an APM test, back-calculation of elastic moduli using FEM
and RLT tests to mainly characterise the deformation behaviour and to develop
relationship between them. APM testing was performed on a pavement model
constructed having a subgrade and a lightly stabilised granular base having thickness of
600 mm and 150 mm respectively. During the test, vertical deformations, pressure and
moisture measurements were taken in addition to the axial loading pressures. Four
cyclic loading pressures, 750 kPa, 1000 kPa, 1250 kPa and 1500 kPa were applied to
the base using 184-mm diameter loading plate a frequency of 2 Hz and rest period of
0.5 second. Approximately 80,000 load cycles were applied for the first three load
sequences and 700,000 for the last before terminating the test. In the RLT test, two tests
having different combinations of confining and deviatoric pressures were carried out
while keeping most of the conditions similar to the APM test. The collected

363
Chapter 7 Characterising Deformation Behaviour using APM Test

deformation, moisture and pressure distributions were processed for obtaining


meaningful outcomes.

The permanent deformation response of base and subgrade of the APM test showed that
the accumulation of permanent deformation achieved a stable value after showing an
increase for the first 10,000 cycles and it was found that the subgrade was responsible
for nearly 75% of the total deformation. Moreover, the accumulated permanent
deformation increased with increasing applied pressure.

Resilient deformations decreased slightly with the number of load cycles after an initial
increase up to 10,000 load cycles and resilient modulus increased with increase cyclic
loading pressure. Similar to the permanent deformation response, subgrade deformed
significantly compared with base. It is found that based on the resilient and permanent
deformation responses classifying the behaviour would help to decide the termination
point. Performing multi-cyclic pressures at ‘reasonable’ numbers of loading cycles
gives more revelations of the material behaviour than performing single stage test for
several millions of cycles.

The moisture contents at different locations were measured using four moisture probes
calibrated for base and subgrade. It was shown that the soil-specific calibration is
essential to improve the accuracy of the measurements and MP 306 moisture sensors
proved that they can be used for continuous measurements. Measurement at different
locations gave some indications regarding the moisture movement due to various
influencing factors.

A FE model was developed in ABAQUS to back-calculate the elastic properties of base


and subgrade. An axisymmetric model simulating APM test was developed using
‘assumed’ elastic properties. A harvesine load pulse at a frequency of 2 Hz was applied
at four loading pressures. Back-calculations were performed matching the experimental
resilient deformations in the 10,000th cycle obtained from the APM test of each loading
pressure to those obtained from an ABAQUS FEA. Back-calculation using isotropic
and cross-anisotropy analysis, were performed by assigning isotropic and cross-
anisotropic elastic properties to both the subgrade and base materials. It was found that,
between these two analysis, the vertical modulus of base did not vary considerably.

364
Chapter 7 Characterising Deformation Behaviour using APM Test

However, subgrade’s modulus from cross-anisotropic analysis showed significantly


higher values than the values obtained from isotropic analysis. Furthermore, it was
found that the moduli of base decreased with increasing cyclic loading pressure.

Vertical and horizontal pressure distributions of experimental and FE model were


analysed and it was found that EPCs can be used to measure pressures to a reasonable
accuracy. It was found that that measuring pressures using an EPC in places where a
material undergoes a large deformation should be avoided, because the EPC needs a
firm base to fit in.

Finally, using RLT test data and FEA, relationship between regression coefficients was
developed in terms of bulk-stress. Similar regression coefficients obtained from APM
showed strong correlations that can be used to predict permanent deformation response
of different combinations of stresses.

365
8 Chapter 8

Summary, Conclusions and Recommendations

8.1 General

The durability of lightly stabilised granular material against f-t and w-d effects was
examined using different laboratory experiments focussing on various mechanical
properties. Laboratory tests such as unconfined compressive strength (UCS), repeated
load triaxial (RLT), flexural beam test, vacuum saturation test (VST) and tube suction
test (TST) were included in the testing program. Along with the different element
testings, variables such as the type of granular material, number of curing days, binder
percentage, curing method and type of f-t procedures were considered for obtaining
better laboratory test to evaluate the durability of lightly stabilised granular materials.
Apart from the durability study, characterising the permanent deformation (PD)
behaviour of lightly stabilised material using RLT and APM testings were presented in
the last two chapters. The outcomes from both these tests were analysed, primarily
using shakedown theory, for the characterising the material’s response. Conclusions
drawn from the durability studies and PD characterisations are presented in the next
section while the last section is dedicated for recommendations for future work.

8.2 Conclusions

8.2.1 Evaluating Suitability of Direct and Indirect Laboratory Tests for


Assessing Freeze-Thaw and Moisture Effects on Lightly Stabilised
Granular Materials

Two granular materials, Canberra materials (CM) and Queensland materials (QM),
stabilised with 1.5% of cement-flyash (CF) were used to prepare cylindrical specimens
Chapter 8 Summary, Conclusions and Recommendations

having dimensions of standard UCS. Moulding moisture content was selected as a


primary variable for assessing the f-t durability using four tests such as UCS after 7-day,
12 f-t cycles (f-t durability test), vacuum saturation and TST. The assessment of
individual test results showed that these four tests showed dependency to the moulding
moisture content in which the UCS increased with decreasing moisture content before
showing decline at the lowest moisture content. A comparison with f-t durability test
and 7-day UCS revealed that f-t cycles caused destructive effects on the stabilised
material. A comparison with all four test results indicated that the 7-day UCS, f-t
durability test and VST showed a good correlation among them. As the testing time for
f-t durability is nearly a month, VST or 7-day UCS tests can be used to assess the
durability of the material. However, as the relationship between DVs and UCS values
from the other there tests did not show good correlation, TST is not suitable for
assessing the f-t durability.

However, additional tests performed using TST indicated following outcomes that gives
more insight about the materials behaviour.

 It showed the dependency to moulding moisture content and binder content.


 It identified that the coarser CM material showed lessor DV, therefore it’s less
susceptible to moisture and f-t effects than finer QM.
 It showed that the rate of increase in DVs and the moisture content measured in
top and bottom of the specimen showed that the reading up to10 days is enough
terminate the test.
 A scattered relationship between DVs and moisture content well explained that
the latter is not a mere measure of moisture content.
 It also revealed that the quality of the porous disk and smoothness of the top
surface of specimens were important to obtain precise and repeatable
measurements.

367
Chapter 8 Summary, Conclusions and Recommendations

8.2.2 Assessing Durability of Lightly Stabilised Granular Material against


Freeze-Thaw and Wet-Dry Cycles Using Different Element Testings
and Testing Procedures

In this study, three different element testings were performed to assess the durability of
lightly stabilised granular materials against f-t and w-d cycles while also considering the
factors causing changes in their mechanical properties. Different factors that could
influence a laboratory f-t durability tests were investigated using UCS, repeated load
triaxial (RLT) test and flexural beam test to obtain best combinations of variable that
could be used to assess the f-t durability of the lightly stabilised material.

The preliminary investigation suggested that the standard Mr test available for the
unbound granular material needed to be altered to suit the increased stiffness of lightly
stabilised granular material. As the lower axial stresses was found to be difficult to
control at 10 Hz frequency and caused inconsistent deformation measurements, higher
deviatoric stress values were selected for the Mr test. The dependency of Mr to stress
and the loading frequency was investigated and it was found that the loading frequency
has minimal effect on Mr with bulk stress model, k-θ, describing the stress dependency.

UCS and RLT f-t durability tests showed that, irrespective of the testing procedure or
variables, lightly stabilised material showed enough resistance against destructive f-t
actions. UCS test results comparison with two tested granular material suggested that
the material with the finer particle size distribution affected more due to f-t actions than
coarser material which suggests that by changing the grading, the effects of f-t can be
minimized.

There were many similarities in the outcome of Mr test results compared with UCS test
results. The analysis of overall test results suggested that irrespective of the level of
stress, the absolute differences in Mr between stress sequences were found to be same.
However, in few series of tests, the Mr test failed to clearly identify the destructive
nature of f-t cycles. The RLT permanent deformation tests revealed that the f-t effects in
rutting was noticeable at higher deviatoric stress levels and selecting a lower stress
value and limiting the number of load cycles to 10,000 would not help comparison
between the specimens.

368
Chapter 8 Summary, Conclusions and Recommendations

The UCS and RLT experiments showed that the effect of f-t is highly dependent on the
available moisture within the soil matrix. Supplying moisture during the thaw cycle
generated more destructive effects than not supplying. The strength gain due to
continuous hydration and pozzalanic reaction also influences the number of f-t cycles
required notice the f-t effects. Controlling the strength gain is important to shorten the
required number of days to complete the study which can be done by accelerated curing.

The w-d tests performed on QM showed that the lightly stabilised material is very
vulnerable to w-d cycles than to f-t cycles. Not a single tested lasted 12 w-d cycles for
the selected w-d testing procedure questioning the durability against the w-d cycles.
Further studies needed investigating suitability of the w-d testing procedure and
evaluating influence of different factors involved with the w-d testing procedure.

The flexural beam test suggested that the f-t and w-d cycles have utmost effect on MoR
and SSM than any other properties. It emphasised that the lightly stabilised granular
material behaves more of unbound than cemented material and characterisation based
on its compressive properties are more meaningful than flexural properties.

Out of the material property tested, it was found that the either UCS or RLT permanent
deformation test could be better choice for the f-t durability study. Also, the tests should
adopt testing procedure that supply moisture during the thaw cycle and limits the
strength gain during the f-t cycles.

8.2.3 Characterising Permanent Deformation Behaviour `of Lightly


Stabilised Granular Material using Repeated Load Triaxle (RLT)
Test

Characterising the PD behaviour using RLT test was performed on CM stabilised with
0.5% and 1.5% of CF and cured for 7 days. Specimens stabilised with 0.5% of CF were
tested for moisture contents of 8% and 6.5% for four and three different confining
pressures respectively and three moisture contents of 8%, 6.5% and 5% were selected
for specimens stabilised with 1.5% of CF binder at one confining pressure of 50 kPa.

369
Chapter 8 Summary, Conclusions and Recommendations

The PD behaviour from each stress sequences were classified based on the accumulated
permanent strain from 3000 to 5000 load cycles to range A, B or C shakedown
behaviour. Based on this classification, shakedown limit calculations were performed
to obtain plastic creep line (PCL) and plastic shakedown line (PSL). The established
shakedown limit equations can be used to predict the material’s performance for a given
stress combinations while considering the impact of moisture content.

Furthermore, two existing models, log and power based models, which relate permanent
strain with number of loadings were used to analyse the permanent deformation
obtained for each stress sequence. These two equations were fitted with the
experimental data to obtain the regression coefficients and based on the regression
constant, log model showed better fit to the experimental data than the power model.

The regression coefficient was further analysed to obtain its dependency on stress
values and to establish an equation to predict PD in terms of bulk stress and confining
stress.

Finally, analyses of resilient deformation responses generally showed the characteristics


of the three shakedown ranges. Mr of specimens stabilised with 8% and 6.5%
compaction moisture contents were studied using k-θ model which showed a good fit
with the data. Mr of specimen compacted with 6.5% showed marginally higher value
than specimen compacted with 8% moisture content particularly the differences were
high at the lower bulk stress and difference reduced with increasing bulk stress.

8.2.4 Characterising Deformation Behaviour of Lightly Stabilised


Granular Material using Accelerated Pavement Model (APM) Test

Characterising the permanent deformation (PD) behaviour using APM test was
performed on a pavement model constructed in a cylindrical tank having 600 mm of
subgrade and 150 mm of lightly stabilised granular base. CM stabilised with 1.5% CF,
cured for 28 days was subjected to four axial cyclic pressures of 750 kPa, 1000 kPa,
1250 kPa and 1500 kPa in which the number of load cycles for the first three loadings
sequences limited to 80,000 and for the last loading sequence 0.7 million load cycles

370
Chapter 8 Summary, Conclusions and Recommendations

were applied. The model pavement instrumented with linear variable differential
transformers (LVDTs), moisture probes (MPs) and earth pressure cells (EPCs) for
measuring vertical deformations, moisture and pressures respectively. The moisture
content readings showed that the soil-specific calibration is essential to improve the
accuracy of the measurements and MP 306 moisture sensors proved that they can be
used for continuous measurements.
The permanent deformation (PD) measurement showed that the accumulation
rates were higher in the first 10,000 cycles and then decreased with increasing
numbers of load cycles. Although the measured PD increased with increasing
axial pressure, rate of accumulation almost ceased after 10,000 cycles showing
that the deformation behaviour is resilient afterwards. The rut depth measured in
the subgrade was found to be three times that of the stabilised base contributing
approximately 75% of the overall PD of the pavement. Furthermore, resilient
deformation also showed that nearly 80% of the resilient deformation comes
from the subgrade emphasizing the importance of strong foundation for a well-
functioning pavement. The study also emphasize the need to characterise the
deformation behaviour for every stress sequence using the data of 50,000 to
100,000 load cycles to avoid continuing the test for higher number of load
cycles which would reveal very little about the material’s behaviour.
An axisymmetric FE model was developed in ABAQUS to back-calculate the
elastic properties of base and subgrade. Back-calculation using isotropic and
cross-anisotropy analysis was performed and it was found that between these
two analyses, the vertical modulus of base did not vary considerably but
subgrade’s modulus from cross-anisotropic analysis showed significantly higher
values than the values obtained from isotropic analysis. Vertical and horizontal
pressure distributions of experimental and FE model were analysed and it was
found that EPCs can be used to measure pressures to a reasonable accuracy and
the measured values matched well with results from cross-anisotropic analysis
than isotropic analysis. The PD obtained from two RLT tests, results from APM
and outcome from FE analysis were used to obtain relationship between
regression coefficients in terms of bulk-stress. Strong correlations between the

371
Chapter 8 Summary, Conclusions and Recommendations

regression coefficients showed that it can be used to predict PD response of a


pavement under different combinations of stresses.

8.3 Recommendations for Further Research

Although the experimental work conducted in this thesis has enhanced the
understanding of the durability and permanent deformation behaviour of lightly
stabilised granular materials, the following recommendations are suggested for further
research studies.

 The DVs measured using the TST did not produce good correlations with the
mechanical properties. Additional studies can be performed with different
granular materials and binders for further validating the outcomes.

 VST, f-t durability test and 7-day UCS test showed good correlations between
their outcomes. The variables used in this investigation were moulding moisture
content and two types of granular materials. Studies with different binder
contents, type of binders and other granular materials are recommended for
further validation of the results.

 TST on specimens subjected to f-t cycles can be performed and compared


against the DVs of specimens that are not subjected to f-t cycles. This will
further help to determine the ability of TST to assess the f-t durability of lightly
stabilised granular materials.

 The shakedown limit equations can be improved by performing more number of


stress sequences particularly at lower stress states.

 The part of the classification of shakedown limits were based on the permanent
strain limits suggested for unbound granular materials. The validity of the limits
to lightly stabilised granular materials is further checked by performing series of
single stage RLT tests for larger number of load cycles.

372
Chapter 8 Summary, Conclusions and Recommendations

 The flexural beam test showed increased vulnerability to f-t and w-d cycles. The
testing program can be extended to include indirect diametrical test (IDT) to
further improve the understanding of the material’s tensile properties against the
f-t and w-d cycles.

 Durability against w-d cycles’ study is extended with additional experiments to


evaluate the impacts of different variables associated with the w-d procedure
including temperature of drying phase and the duration of wetting and drying
phases.

 The flexural tests were performed for monotonic loading conditions in this
research. However, the effects of f-t and w-d can be further investigated for
dynamic loadings with different stress ratios similar to the Mr test.

 In the APM test, MPs and EP cells were used for obtaining the moisture and
pressure distribution. The number of MPs and EPs can be increased to obtain
more accurate moisture and pressure distribution.

 In the FE analysis, the materials can be modelled as non-linear elastic model


using stress dependent Mr. The function user material sub routine available in
ABAQUS can be used to input the constitutive models.

 The APM testing was performed for particular moisture content and the
moisture movement did not vary considerably throughout the test. The study
could be extended to include the effects of flooding on the moisture changes and
deformation characteristics using the same APM testing facility.

373
References

References

AASHTO-T22 2014. Standard Method of Test for Compressive Strength of Cylindrical


Concrete Specimens, Washingtan D.C.
AASHTO-T307-99 2007. Determining the Resilient Modulus of Soils and Aggregate
Materials. Washington, D. C.
AASHTO 1993. Guide for Design of Pavement Structures, Washington DC.
AASHTO 2008. Mechanistic-Empirical Pavement Design Guide - A Manual of
Practice (Interim Edition), Wasington DC.
ACI 1997. Prediction of Creep, Shrinkage, and Temperature Effects in Concrete
Structures. In 209R-92. Part 1: Materials and General Properties of Concrete,
American Concrete Institute, Farmington Hills Michigan.
ACI. 2010. ACI Concrete Terminology [online]. Available from
http://www.ciccp.es/ImgWeb/Castilla%20y%20Leon/Documentaci%C3%B3n%
20T%C3%A9cnica/ACI%20Concrete%20Terminology.pdf.
Ahmed Z., Marukic I., Zaghloul S., and Vitillo N. 2005. Validation of Enhanced
Integrated Climatic Model Predictions with New Jersey Seasonal Monitoring
Data. Transportation Research Record: Journal of the Transportation Research
Board, 1913: 148-161.
Aitchison G. D., and Richards B. G. 1965. A broad scale study of moisture conditions
in pavement subgrades througout
Australia - the selection of design values for soil suction equilibri and soil suction
changes in pavement subgrades, Reference has to be found.
Arnold G., and Morkel C.2012. Development of tensile fatigue criteria for bound
materials New Zealand Institute of Highway Technology.
AS-3582.1 1998. Supplementary cementitious materials for use with portland and
blended cement - Fly ash, Sydney Australia.
AS-3582.2 2001. Supplementary cementitious materials for use with Portland and
blended cement: Part 2: Slag - Ground granulated iron blast-furnace.
AS-3972 2010. General purpose and blended cements, Sydney, Australia.
AS-5101.4 2008. Methods for Preparation and Testing of Stabilised Materials -
Unconfined Compressive Strength of Compacted Materials, Sydney Australia.
ASTM-C469 2014. Testing Methods for Static Modulus of Elasticity and Poisson’s
Ratio of Concrete in Compression, ASTM International, West Conshohocken,
PA.

374
References

ASTM-C593 2011. Standard Specification for Fly Ash and Other Pozzolans for Use
With Lime for Soil Stabilization, ASTM International, West Conshohocken, PA.
ASTM-C1609 2010. Standard Test Method for Flexural Performance of Fiber-
Reinforced Concrete, West Conshohocken,United States.
ASTM-D559 2007. Standard Test Methods for Wetting and Drying Compacted Soil-
Cement Mixtures, West Conshohocken,United States.
ASTM-D560 2003. Standard Test Methods for Freezing and Thawing Compacted Soil-
Cement Mixtures, West Conshohocken,United States.
ASTM-D593 2011. Standard Specification for Fly Ash and Other Pozzolans for Use
With Lime for Soil Stabilization, West Conshohocken,United States.
ASTM-D1241-07 2007. Standard Specification for Materials for Soil-Aggregate
Subbase, Base, and Surface Courses.
ASTM-D1632 2007. Standard Practice for Making and Curing Soil-Cement
Compression and Flexure Test Specimens in the Laboratory, West
Conshohocken,United States.
ASTM-D1633 2000. Standard Test Method for Compressive Strength of Molded Soil-
Cement Cylinders, ASTM International, West Conshohocken, PA.
ASTM-D5102-09 2009. Standard Test Method for Unconfined Compressive Strength
of Compacted Soil-Lime Mixtures, ASTM International, West Conshohocken,
PA.
Ateş A. 2013. The Effect of Polymer-Cement Stabilization on the Unconfined
Compressive Strength of Liquefiable Soils. International Journal of Polymer
Science, 2013.
Austin A. 2009. Fundamental characterisation of unbound base course materails under
cyclic loading. Master of Science in Civil Engineering, Louisiana State
University.
Austroads 2002. Guide to Stabilisation in Roadworks, Sydney Aystralia.
Austroads 2010a. Guide to Pavement Technology Part 2: Pavement Structural Design,
Sydney, Australia.
2010b. Guide to Pavement Technology.
Austroads.2010c. Cost Effective Structural Treatments for Rural Highways: Cemented
Materials.
AustStab T. N. 2007. Stabilisation using insoluble dry powdered polymers.
AustSTAB T. N. 2010. What is Lime.

375
References

Ba. M., Fall. M., Sall. O. A., and Samb. F. 2012. Effect of Compaction Moisture
Content on the Resilient Modulus of Unbound Aggregates from Senegal
Geomaterials, 2: 19-23.
Barbu B., McManis K., and Nataraj M.2004. Study of Silts Moisture Susceptibility
Using the Tube Suction Test, Transportation Research Board 2004 Annual
Meeting, CD-ROM Publication, Transportation Research Board, National
Research Council, Washington D. C. .
Barbu B. G., and Scullion T.2005. Repeatability and Reproducibility Study for Tube
Suction Test, Implementation of TubeSuction Test in TxDOT Districts.
Barksdale R. D. Laboratory evaluation of rutting in basecourse materials. In 3rd
Internationl Conference on Structural Design of Asphalt Pavements1972, pp.
161-174.
Barksdale R. G. 1971. Compressive Stress Pulse Times in Flexible Pavements for Use
in Dynamic Testing. Highway Research Record, 345(345): 32-44.
Basma A. A., and Al-Suleiman 1991. Climatic consideration in new AASHTO flexible
pavement design. JOURNAL OF TRANSPORTATION ENGINEERING,
177(2): 210-223.
Bishop A. W. 1959. The principle of effective stress. Tek. Ukeblad, 106(39): 859–863.
Bulut R., Muraleetharan K. K., Zaman M., Yue E., Soltani H., and Hossain Z.2013.
Evaluation Of The Enhanced Integrated Climatic Model For Modulus-Based
Construction Specification For Oklahoma Pavements, Oklahoma Transportation
Center.
Burmister D. M. 1945. The General Theory of Stresses and -Displacements in Layered
Soil Systems. Journal of applied physics, 16.
Carvalho R. L., and Schwartz C. W. 2006. Comparisons of Flexible Pavement Designs
(AASHTO Empirical Versus NCHRP Project 1-37A Mechanistic–Empirical).
Transportation Research Record: Journal of the Transportation Research Board,
1947: 167-174.
Chakrabatri S., Kodikara J. K., and Pardo L. 2002. Survey results on stabilisation
methods and performance of local government roads in Australia. Road and
Transport Research, 11.
Chu T. Y. 1977. Soil Moisture as a Factor in Subgrade Evaluation. ransportation
Engineering Journal, 103(Has to be found...).
Craciun O., and Lo S.-C. R. 2010. Matric Suction Measurement in Stress Path Cyclic
Triaxial Testing of Unbound Granular Base Materials. Geotechnical Testing
Journal, 33(1).
Dempsey, and Thompson 1968. Durability Properties of Lime – Soil Mixtures.
Highway Research Record, 235: 61-75.

376
References

Dempsey B. J., and Thompson M. R.1973. Vacuum saturation method for predicting
freeze-thaw durability of stabilised materials, Transportation Engineering Series
no 6, Iilinois cooperative highway research program series no 143.
Drumm E. C., Reeves J. S., Madget M. R., and Trolinger W. D. 1997. Subgrade
Resilient Modulus Correction for Saturation Effects. Journal of Geotechnical &
Geoenvironmental Engineering, 123(7).
Ekblad J. 2006. Influence of Water on Coarse Granular Road Material Properties.
EN_13286-7 2004. Unbound and Hydraulically Bound Mixtures - Part 7:Cyclic Load
traxial Test for Unbound Mixtures.
Fredrick Lekarp U. I. a. A. D. 2000. RESILIENT RESPONSE OF UNBOUND
AGGREGATES. JOURNAL OF TRANSPORTATION ENGINEERING.
Gandara J. A., Kancherla A., Alvarado G., Nazarian S., and Scullion T.2005. Impact of
Aggregate Gradation on Base Material Performance Research Project TX-0-
4358 Materials, Specifications, and Construction Techniques for High
Performance Flexible Bases Texas Department of Transportation.
Ganne V. K. 2009. Long-Term Durability Studies On Chemically Treated Reclaimed
Asphalt Pavement (Rap) Materials Master Of Science In Civil Engineering, The
University Of Texas At Arlington.
Gaskin G. J., and Miller J. D. 1996. Measurement of of soil water content using a
simplified impedance measuring technique. Journal of aggrecultural
engineering, 63: 153-160.
Gnanendran C. T., and Piratheepan J. 2008. Characterisation of a lightly stabilised
granular material by indirect diametrical tensile testing. International Journal of
Pavement Engineering, 9(6): 445-456.
Gnanendran C. T., and Piratheepan J. 2010. Determination of Fatigue Life of a
Granular Base Material Lightly Stabilized with Slag Lime from Indirect
Diametral Tensile Testing. JOURNAL OF TRANSPORTATION
ENGINEERING, 136(8).
González A., Jameson G., Carteret R. d., and Yeo R. 2013. Laboratory fatigue life of
cemented materials in Australia. Road Materials and Pavement Design, 14(3):
518-536.
Gutherie W. S., Roper M. B., and Eggett D. L. 2008. Evaluation of Laboratory
Durability Tests for Stabilized Aggregate Base Materials In Transportation
Research Board 87th Annual Meeting (CD-ROM). Transportation Research
Board, National Research Council, Washinton, D.C.
Guthrie W. S., and Scullion T.2003. Interlaboratory Study Ofthe Tube Suction Test,
Research Report 0-4114-2, Texas Transportation Institute, College Station, TX.
Guthrie W. S., and Shambaugh 2009. Ruggedness Evaluation of the Tube Suction Test.
Cold Regions Engineering.

377
References

Hazell D., Jameson G., Vuong B., Moffatt M., Martin A., and Lourensz S.2010.
Assessment of Rut-resistance of Granular Bases using the Repeated Load
Triaxial Test AP–R360/10, Austroads.
Hicks R. G. 1970. Factors influencing the resilient properties of granular materials. PhD
thesis, University of California, Berkeley, Berkeley, Calif.
2004. Pavement Analysis and Design.
Jameson G. W., Sharp K. G., and R Y.1992. Cement-treated crushed rock pavement
fatigue under accelerated loading, The Mulgrave (Victoria) ALF trial
1989/1991(ARR 229). , Vermont South, Vic: Australian Road Research Board
(ARRB).
Johnson Plastic Flow, residual stresses and SHAKEDOWN in rolling contact. In 2nd
International Conference on Contact Mechanics and Wear of Rail/Wheel
Systems, Uni-versity of Rhode Island, Waterloo Ontario1986.
Kalankamary G. P., and Donald D. T. 1963. Development of Freeze-Thaw Test for
Design of Soil-cement. Highway Research Board, 36: 77-96.
kamal M. A., Dawson A. R., Farouki O. T., Hughes D. A. B., and A S. A. 1993. Field
and laboratory evaluation of the mechanical behaviro of unbound granular
materials in pavements. Transportation Research Record Washington D C,
1406: 88-97.
Kannemeyer L., and Plessis L. D. Ultra-Thin CRCP: Modeling, Testing Under
Accelerated Pavement Testing and Field Application for Roads. In Proceedings
of the 2nd CROW workshop, University of Delft, Belgium2006.
Karasahin M., Dawson A. R., and Holden J. T. 1993. Applicability of resilient
constitutive models of granular materials for unbound base layers Transportation
Research Record, National research coucil, Washington D C, 1406: 98-107.
Khalid A. T. A. H. A. 2008. Characterizing the resilient behaviour of treated municipal
solid waste bottom ash blends for use in foundations.
Khoury N. N. 2005. Durability of cementitiously stabilized aggregate bases for
pavement application.
Khoury N. N., and Zaman M. M. 2002. Effect of wet-Dry Cycles on Resilient Modulus
of Class C Coal fly Ash-Stabilised Aggregate Base. Transportation Research
Record: Journal of Transportation Research Board, 1787: 13-21.
Khoury N. N., and Zaman M. M. 2004. Correlation between resilient modulus,moisture
variation, and soil suction for subgrade soils. Journal of the Transportation
Research Board, 1874: 99-107.
Khoury N. N., and Zaman M. M. 2006. Durability Effects on Flexural Behavior of fly
Ash Stabilised Limestone Aggregate. ASTM Journal of testing and evaluation,
34(3).

378
References

Khoury N. N., and Zaman M. M. 2007. Environmental Effects on Durability of


Aggregates Stabilized with Cementitious Materials. Journal of Materials in Civil
Engineering, 19(1).
Kim D. S., and Drabkin S. 1994. Accoracy improvement of external resilient modulus
measurement using specimen grouting to end platens. 73rd Annual meeting
meeting of the transportation research board, Washington DC.
Kim I. T., and Tutumluer E. 2006. Field Validation of Airport Pavement Granular
Layer Rutting Predictions. Transportation Research Record, 1952: 48-57.
Kim M. 2007. Three-Dimensional Finite Element Analysis of Flexible Pavements
considering Non-linear Pavement Foundation Behavior.
Kolisoja P. 1997. Resilient deformation characteristics of granular ma-terials.
Leib B., Jabro J., and Matthews G. 2003. Field evaluation and performance comparison
of soil moisture sensors. Soil science, 168: 396-408.
Lekarp F., Isacsson U., and Dawson A. 2000. State of the Art II: Permanent Strain
Response of Unbound Aggregates. Journal of Transportation Engineering,
126(1): 76-83.
Li L., Liu J., Zhang X., and Saboundjian S. 2010. Laboratory Investigation of
SeasonalVariations in Resilient Modulus of Alaskan Base Course Material In
GeoShanghai 2010 International Conference.
Lim S., and Zollinger D. G. 2003. Estimation of Compressive Sterngth and Modulus of
Elasticity of Cement Treated Aggregate Base Material.
Little D. N.2000. Evaluation of structural properties of lime stabilized soils and
aggregates, National lime association.
Little D. N., Males E. H., Prusinski J. R., and Steward B.2002. Cementitious
Stabilisation, Millennium Papers A2J01, Trasportation Research Board.
Litwinowicz A. 1986. Characterisation of cement stabilised crushed rock pavement
materials. Master of Engineering thesis, University of Queensland, Brisbane,
Australia.
Lo S. C. R., Chu J., and Lee I. K. 1989. A Technique for Reducing Memebrane
Penetration and Bedding Errors. Geotechnical Testing Journal, 12(4): 159-161.
Loulizi A., Al-Qadi I. L., Lahouar S., and Freeman T. E. 2002. Measurement of
Vertical Compressive Stress Pulse in Flexible Pavements. Transportation
Research Record, 1816(2376): 125-135.
2013. Pavement Engineering: Principles and Practice.
Marradi A., and Fausto L. 2008. Performance of cement stabilized recycled crushed
concrete. In First International Conference on Transport Infrastructure ICTI,
Beijing, China.

379
References

Miller G. A., and Zaman M. 2000. Field and Laboratory Evaluation of Cement Kiln
Dust as Soil Stabilizer. Transportation Research Record: Journal of
Transportation Research Board, 1714: 25-32.
Mittelbach H., Lehner I., and Seneviratne S. I. 2012. Comparison of four soil moisture
sensor types under field conditions in Switzerland. Journal of Hydrology, 430-
431: 39-49.
Monismith C. L., Seed H. B., Mitry F. G., and Chan C. K. Prediction of pavement
deflections from laboratory tests. In Proc., 2nd International Conference Struct.
Des. of Asphalt Pavements, 109–1401967.
Morgan J. R. The response of granular materials to repeated loading. In Proc., 3rd
Conf., ARRB, 1178–11921966.
N K. N., and Zaman 2007. Durability of stabilized base courses subjected to wet – dry
cycles.
N K. N., and R B. 2010. Performance of a Stabilized Aggregate Base Subject to
Different Durability Procedures. Journal of Materials in Civil Engineering 22(5).
Nazarian S., and Yuan D. 2008. Variation in Moduli of Base and Subgrade with
Moisture. In GeoCongress 2008.
Nazarian S., Pezo R., and Picornell M.1996. Testing methodology for resilient modulus
of base materials, Research Project 0-1336, Texas department of transportation
in cooperation with federal highway administration.
Nazzal M. D., and Mohammad L. N. 2010. Estimation of Resilient Modulus of
Subgrade Soils for Design of Pavement Structures. Journal of Material in Civil
Engineering, 22.
NCHRP 2004. Guide for mechanistic-emprical design, Part 2. Design Inputs.
NCHRP 2008. Calibration and Validation of the Enhanced Integrated Climatic Model
for Pavement Design.
2010. Concrete Technology.
Ohiduzzaman M., Lo S. C. R., and Craciun O. 2012. Matric Suction and Deformation
Behavior of Unbound Granular Base (UGB) Materials under Constant and
Variable Cell Pressure. Unsaturated Soils: Research and Applications, 1: 347-
354.
Parker J. W. 2008. EVALUATION OF LABORATORY DURABILITY TESTS FOR
STABILIZED SUBGRADE SOILS. M.Sc, Brigham Young University.
Parsons R. L., and Milburn J. P. 2002. Engineering Behavior of Stabilized Soils. TRB
2003 Annual Meeting.
Paul D., and Gnanendran C. T. 2012a. Prediction of Nonlinear Stress-Strain
Relationship of Lightly Stabilized Granular Materials from Unconfined

380
References

Compression Testing. Journal of Materials in Civil Engineering, 24(8): 1118-


1124.
Paul D. K. 2012. Charaterisation of Lightly Stabilised Material using various laboratory
tests. PhD Thesis, University of New South Wales, Australian Defence Force
Academy, Canberra, Australia.
Paul D. K., and Gnanendran C. T. 2012b. Characterisation of Lightly Stabilised
Granular Base Materials by Flexural Beam Testing and Effects of Loading Rate.
Geotechnical Testing Journal, 35(5).
Paul D. K., and Gnanendran C. T. 2013. Stress–strain behaviour and stiffness of lightly
stabilised granular materials from UCS testing and their predictability.
International Journal of Pavement Engineering, 14(3): 291-308.
Paul D. K., and Gnanendran C. T. 2015. Characterization of Lightly Stabilized
Granular Base Materials Using Monotonic and Cyclic Load Flexural Testing.
JOURNAL OF MATERIALS IN CIVIL ENGINEERING, 27.
Paul D. K., Theivakularatnam M., and Gnanendran C. T. 2015. Damage Study of a
Lightly Stabilised Granular Material Using Flexural Testing. Indian
Geotechnical Journal, Published online.
Paute J. L., Hornych P., and Benaben J. P. Repeated Load Triaxial Testing of Granular
Materials in the French Network of Labo-Ratoires des Ponts et Chausse´es. In
Flexible Pavements, Euroflex 1993, Balkema, Rotterdam, The Netherlands1996,
pp. 53-64.
Perera Y.Y P. E., Claudia E. Zapata, William N. Houston, and Houston. S. L. Long-
Term moisture conditions under highway pavements. In Geo Trans 20042004.
Pezo R. F., German C., Hudson W. R., and Stroke K. H. 1991. A reliable resilient
modulus testing system. Transportation Research Record, 1307(Natioanl
Research council, Washington): 90-98.
Ping P. E., Xiong W., and Yang Z. 2003. Implementing Resilient Modulus Test for
Design of Pavement Structures in Florida.
Piratheepan J., Gnanendran C. T., and Lo S. R. 2009. Characterization of
Cementitiously Stabilised Granular Materials for Pavement Design Using
Unconfined Compression and IDT Testings with Internal Displacement
Measurements. ASCE’s Journal of Materials in Civil Engineering, 22(5): 495-
505.
Piratheepan J. G. C. T. a. L. S. R. 2010. Characterizationof cementitiously stabilized
granular materials for pavement design using unconfined compression and IDT
testing with internal displacement measurements. Journal of Materials in Civil
Engineering, 22(5): 495-505.
Plauborg F., Iversen B. V., and Lærke P. E. 2005. In situ comparison of three dielectric
soil moisture sensors in drip irrigated sandy soils. Vadose Zone Journal, 4:
1037-1047.

381
References

Program S. H. R.1993a. Distress identification manual for the long term pavement
performance project, SHRP-P-338, Washington, D.C.
Program S. H. R.1993b. Layer Moduli Backcalculation Procedure: Software Selection
SHRP-P-65 1, National Research Council.
Puppala A., Mohammad L., and Allen A. 1999. Permanent Deformation
Characterization of Subgrade Soils from RLT Test. Journal of Materials in Civil
Engineering, 4(11).
Qian J. 2010. In-situ Testing and Evaluation of Moisture content in Existing Sub-grade.
In GeoShanghai 2010 International conference.
Quintero N. M. 2007. Validation of the Enhanced Integrated Climatic Model (EICM)
for the Ohio SHRP Test Road at U.S. 23. Master of Science.
Roper M. B. 2007. Evaluation of Laborat or y Durability Tests forStabilized Aggregate
Base Materials. Master of Science, Brigham Young University
Rust F. C., Grobler J. E., and Myburgh P. A. Towards analytical mix design for large
stone asphalt mixes. In Proceedings of the 7th International Conference on
Asphalt Pavements, Nottingham, UK1992.
Saarenketo T.2000. Tube suction test - results of round robin tests on unbound
aggregates, Finnish National Road Adminstration.
Saarenketo T., and scullion T.1996. Using Electrical Properties to classify the strength
properties of base course aggregates, Texas Transportation Institute.
Sahu V., and Gayathri V. 2014. The use of stabilized fly ash as a green material in
pavement substructure: A review International journal of civil and structural
engineering, 4(3): 306-314.
Salour F., Erlingsson S., and Zapata C. E. 2014. Modelling resilient modulus seasonal
variation of silty sand subgrade soils with matric suction control. Canadian
Geotechnical journal, 51(12): 1413-1422.
Sangghaleh A., Pan E., Green R., Wang R., Liu X., and Cai Y. 2014. Backcalculation
of pavement layer elastic modulus and thickness with measurement errors.
International Journal of Pavement Engineering, 15(6): 521-531.
Scullion T., and Saarenketo T. 1997. Using Suction and Dielectric Measurements as
Performance Indicators for aggregate base material. Transportation Research
Record, 1577: 37-44.
Shihata S. A., and Baghdadi Z. A. 2001. Simplified method to assess freeze-thaw
durability of soil cement. Journal of Material in Civil Engineering, 13(4).
Si Z., and Herrera C. H. 2007. Laboratory and Field Evaluation of Base Stabilization
Using Cement Kiln Dust. Transportation Research Record: Journal of the
Transportation Research Board, 1989(2): 42-49.

382
References

Simonsen E., Janoo V. C., and Isacsson U. 2002. Resilient Properties of Unbound Road
Materials during Seasonal Frost Conditions.
Sireesh Saride M., Someshwar Rao M., SarathChandra Prasad. J S. M., and Babu R. D.
2014. EVALUATION OF FLY-ASH-TREATED RECLAIMED ASPHALT
PAVEMENT FOR DESIGN OF SUSTAINABLE PAVEMENT BASES - AN
INDIAN PERSPECTIVE. In Geo-Congress 2014.
Smith W., and Hansen B. 2003. Detailed Investigation of the Performance of GB
Cements in Pavement Stabilisation Works. In 21st ARRB Conference.
Smith W., and Vorobieff G. 2007. Recognition of Sustainability by using stabilisation
in road rehabilitation In ASA Sustainability & Slag Conference, Sydney.
Smith W. S., and Nair K. 1973. Development of procedures for characterization of
untreated granular base coarse and asphalt treated base course materials. Rep.
No. FHWA-RD-74-61, Federal Highway Administration, Washington, D.C.
Solanki P., Zaman M. M., and Khalife R. 2011. Tube Suction Test for Evaluating
Durability of Cementitiously Stabilized Soils.
Steyn W. J. v. M., Beer M. d., and Visser A. T. Thin Asphalt and Double Seal
Rehabilitated Lightly Cemented Pavements: Evaluation of Structural Behaviour
and Life Cycle Costs. In 8th International Conference on Asphalt
Pavements1997.
Sweere G. T. H. 1990a. Unbound granular bases for roads.
Sweere G. T. H. 1990b. Unbound granular bases for roads.
Syed I., Scullion T., and Randolph R. B. 2000. Tube Suction Test for Evaluating
Aggregate Base Materials in Frost- and Moisture-Susceptible Environments.
Transportation Research Record: Journal of the Transportation Research Board,
1709: 78-90.
Syed I., Scullion T., and Smith R. E.2003. Recent development in charaterising
durability of stabilised materials, Transportation Research Board 82nd Annual
Meeting, CD-ROM Publication, Transportation Research Board, Washington,
DC.
Theivakularatnam M., and Gnanendran C. T. Effects of Freeze-thaw on Lightly
Stabilized Material?s Flexural Properties. In Geo-Congress 2014 Technical
Papers, Atlanta,USA2014, pp. 465 - 3475.
Thompson M. R., and L S. K. 1990. Repeated Triaxial Characterization of Granular
Bases. Transportation Research Board, 1278: 7-17.
Tom Wilmot B. R. Stabilised Pavements - selecting the additive:cementitious, polymer
or bitumen. In Australian Civil Engineering Transactions, Vol. CE42, 20002000.
Vorobieff G. 1997a. Cementitious Binder Options.

383
References

Vorobieff G.1997b. Cementitious Binder Options.


Vorobieff G. 2000. Chemical Binders used in Australia.
Walker J., Willgoose G., and Kalma J. 2004. In situ measurement of soil moisture: A
comparison of techniques. Journal of Hydrology, 293: 85-99.
Wang H., and Zhao Y. Measuring Water Content Variations in Stems by Standing
Wave Ratio Principle In International Conference on Mechatronics and
Automation, Xi'an, China2010, pp. 124-128.
Werkmeister S. Shakedown Analysis of Unbound Granular Materials using Accelerated
Pavement Test Results from New Zealand's CAPTIF Facility. In GeoShanghai
2006, Shanghai2006, pp. 220-228.
Werkmeister S., Dawson A. R., and Wellner F. 2005. Permanent Deformation
Behaviour of Granular Materials. Road Materials and Pavement Design.
White G. W., and Gnanendran C. T. R. 2005. The influence of compaction method and
density on the strength and modulus of cementitiously stabilised pavement
materials. The International Journal of Pavement Engineering, 6(2): 97-110.
Wilson J. H. D. J., Henning T. F. P., and Alabaster D. 2014. Comparing Results
between the Repeated Load Triaxial Test and Accelerated Pavement Test on
Unbound Aggregate. Journal of Material in Civil Engineering, 26(3).
Yang S. R., Huang W. H., and Tai Y. T. 2005. Variation of resilient modulus with soil
suction for compacted subgrade soils. Journal of the transportation board (1913):
99-106.
Yeo Y., and Nikraz H. 2012. Evaluation of water ingress in cement treated material for
durability assessments. Australian Journal of Civil Engineering, 10(2).
Yeo Y. S., Jitsangiam P., and Nikraz H. 2011. Dynamic Effects on Fatigue Life of
Cement Treated Crushed Rock. In International Conference on Advances in
Geotechnical Engineering, Perth, Australia.
Yeo Y. S., Nikraz H., and Jitsangiam P. 2012. Tube Suction Test to Measure Moisture
Susceptibility of Australian Pavements. Engineering journal, 16(4).
Yin M. I.2011. Using local gravel sources and granular stabilisation for the production
of pavement gravels in the Toowoomba Region, University of Southern
Queensland, Faculty of Engineering and Surveying.
Yuan D., and Nazarian S. 2003. Variation in Moduli of Base and Subgrade with
Moisture.
Zaman. M. M., Zhu. J. H., and Laguros. J. G. 1999. Durability effects on resilient
Moduli of Stabilised Aggregate Base. Transportation Research Record, 1687:
29-38.

384
References

Zhang Z., and Tao M. 2008. Durability of Cement Stabilized Low Plasticity Soils.
Journal of geotechnical and geoenvironmental engineering, 134(2).

385
Appendix A

Appendix A
0.45

0 f-t cycle
0.4
Specimen 1
Specimen 2
0.35 Specimen 3
Mean stress-strain curve

0.3
Flexural Stress (MPa)

0.25

0.2

0.15

0.1

0.05

0 200 400 600 800 1000 1200 1400 1600 1800


Strain (10 -3)

Figure A. 1: Flexural stress-strain relationship of specimen stabilised with 1.5%


CF binder and subjected to 0 f-t cycle
0.3

1 f-t cycle
Specimen 1
0.25 Specimen 2
Specimen 3
Mean stress-strain curve

0.2
Flexural Stress (MPa)

0.15

0.1

0.05

0 200 400 600 800 1000 1200 1400 1600 1800


Strain (10 -3)

Figure A. 2: Flexural stress-strain relationship of specimen stabilised with 1.5%


CF binder and subjected to 1 f-t cycle

A-1
Appendix A

0.3
2 f-t cycles
Specimen 1
Specimen 2
0.25 Specimen 3
Mean stress-strain curve

0.2
Flexural Stress (MPa)

0.15

0.1

0.05

0 200 400 600 800 1000 1200 1400 1600 1800


Strain(10 -3)

Figure A. 3: Flexural stress-strain relationship of specimen stabilised with 1.5%


CF binder and subjected to 2 f-t cycles

4 f-t cycles
Specimen 1
Specimen 2
0.25 Specimen 3
Mean stress-strain curve

0.2
Flexural Stress (MPa)

0.15

0.1

0.05

0 200 400 600 800 1000 1200 1400 1600 1800


Strain (10 -3)

Figure A. 4: Flexural stress-strain relationship of specimen stabilised with 1.5%


CF binder and subjected to 4 f-t cycl

A-2
Appendix A

0.25 8 f-t cycles


Specimen 1
Specimen 2
Specimen 3
Mean stress-strain curve
0.2

Flexural Stress (MPa)


0.15

0.1

0.05

0 200 400 600 800 1000 1200 1400 1600 1800


Strain (10 -3)

Figure A. 5: Flexural stress-strain relationship of specimen stabilised with 1.5%


CF binder and subjected to 8 f-t cycles
0.25

12 f-t cycles
Specimen 1
Specimen 2
0.2 Specimen 3
Mean stress-strain curve
Flexural Stress (MPa)

0.15

0.1

0.05

0 200 400 600 800 1000 1200 1400 1600 1800


Strain (10 -3)

Figure A. 6: Flexural stress-strain relationship of specimen stabilised with


1.5% CF binder and subjected to 12 f-t cycles

A-3
Appendix A

2.5

0 f-t cycle
Specimen 1
Specimen 2
2
Specimen 3
Mean stress-strain curve

Flexural Stress (MPa)


1.5

0.5

0 200 400 600 800 1000 1200 1400 1600 1800


Strain (10 -3)

Figure A. 7: Flexural stress-strain relationship of specimen stabilised with 3% CF


binder and subjected to 0 f-t cycle

0.7
1 f-t cycle
Specimen 1
0.6 Specimen 2
Specimen 3
Mean stress-strain curve
0.5
Flexural Stress (MPa)

0.4

0.3

0.2

0.1

0 200 400 600 800 1000 1200 1400 1600 1800


Strain (10 -3)

Figure A. 8: Flexural stress-strain relationship of specimen stabilised with 3% CF


binder and subjected to 1 f-t cycle

A-4
Appendix A

0.8
2 f-t cycles
Specimen 1
0.7
Specimen 2
Specimen 3
0.6 Mean stress-strain curve

Flexural Stress (MPa) 0.5

0.4

0.3

0.2

0.1

0 200 400 600 800 1000 1200 1400 1600 1800


Strain(10 -3)

Figure A. 9: Flexural stress-strain relationship of specimen stabilised with 3% CF


binder and subjected to 2 f-t cycles

0.7
4 f-t cycles
Specimen 1
Specimen 2
0.6
Specimen 3
Mean stress-strain curve

0.5
Flexural Stress (MPa)

0.4

0.3

0.2

0.1

0 200 400 600 800 1000 1200 1400 1600 1800


Strain (10 -3)

Figure A. 10: Flexural stress-strain relationship of specimen stabilised with 3% CF


binder and subjected to 4 f-t cycles

A-5
Appendix A

0.7

8 f-t cycles
0.6
Specimen 1
Specimen 2
Specimen 3
0.5
Mean stress-strain curve

Flexural Stress (MPa)


0.4

0.3

0.2

0.1

0 200 400 600 800 1000 1200 1400 1600 1800


Strain (10 -3)

Figure A. 11: Flexural stress-strain relationship of specimen stabilised with 3% CF


binder and subjected to 8 f-t cycles

0.8

12 f-t cycles
0.7
Specimen 1
Specimen 2
0.6 Specimen 3
Mean stress-strain curve
Flexural Stress (MPa)

0.5

0.4

0.3

0.2

0.1

0 200 400 600 800 1000 1200 1400 1600 1800


Strain (10 -3)

Figure A. 12: Flexural stress-strain relationship of specimen stabilised with 3% CF


binder and subjected to 12 f-t cycles

A-6
Appendix B

Appendix B
0.84 1.2

0.8
1.1
Resilient strain (10-3)

Resilient strain ( 10-3)


0.76 Shakedown range B

Shakedown range A 1

0.72

0.9
0.68

0.64 0.8
0 2000 4000 6000 8000 10000 0 2000 4000 6000 8000 10000
Number of load cycles Number of load cycles

(a) (b)

Figure B. 1: Resilient strain measured in RLT tests varying with number of cycles of specimens compacted at 8% moisture content with
50 kPa confining pressure for (a) stress sequence 1 and (b) stress sequence 2

B-1
Appendix B

1.5 2.4

2.2
1.4

Resilient strain ( 10-3)


Resilient strain ( 10-3)

Shakedown range B 2

1.3
Shakedown range C
1.8

1.2
1.6

1.1 1.4
0 2000 4000 6000 8000 10000 0 1000 2000 3000 4000
Number of load cycles Number of load cycles

(a) (b)

Figure B. 2: Resilient strain measured in RLT tests varying with number of cycles of specimens compacted at 8% moisture content with
50 kPa confining pressure for (a) stress sequence 3 and (b) stress sequence 4

B-2
Appendix B

0.95 1.2

0.9 1.15

Resilient strain ( 10-3)


Resilient strain ( 10-3)

0.85 1.1

Shakedown range B

0.8 Shakedown range B 1.05

0.75 1

0.7 0.95

0 2000 4000 6000 8000 10000 0 2000 4000 6000 8000 10000
Number of load cycles Number of load cycles

(a) (b)

Figure B. 3: Resilient strain measured in RLT tests varying with number of cycles of specimens compacted at 8% moisture content with
100 kPa confining pressure for (a) stress sequence 1 and (b) stress sequence 2

B-3
Appendix B

1.5
2

1.4
1.8
Resilient strain ( 10-3)

Resilient strain (10-3)


Shakedown range C
1.3
Shakedown range C
1.6

1.2 1.4

1.1 1.2
0 2000 4000 6000 8000 10000 0 2000 4000 6000 8000 10000
Number of load cycles Number of load cycles

(a) (b)

Figure B. 4: Resilient strain measured in RLT tests varying with number of cycles of specimens compacted at 8% moisture content with
100 kPa confining pressure for (a) stress sequence 3 and (b) stress sequence 4

B-4
Appendix B

2.5

Resilient strain ( 10-3)


1.5

Shakedown range C

0.5

0
0 2000 4000 6000 8000
Number of load cycles

Figure B. 5: Resilient strain measured in RLT tests varying with number of cycles of specimens compacted at 8% moisture content and
tested with with 100 kPa confining pressure for stress sequence 5

B-5
Appendix B

1 1.15

1.1

0.9

1.05

Resilient strain ( 10-3)


Resilient strain ( 10-3)

0.8 1

Shakedown range A/B Shakedown range B

0.95

0.7

0.9

0.6 0.85

0 2000 4000 6000 8000 10000 0 2000 4000 6000 8000 10000
Number of load cycles Number of load cycles

(a) (b)

Figure B. 6: Resilient strain measured in RLT tests varying with number of cycles of specimens compacted at 8% moisture content with
140 kPa confining pressure for (a) stress sequence 1 and (b) stress sequence 2

B-6
Appendix B

1.5
1.3

1.4
1.2

Resilient strain ( 10-3)


Resilient strain ( 10-3)

1.3 Shakedown range C


1.1 Shakedown range C

1.2
1

1.1
0.9
0 2000 4000 6000 8000 10000
0 2000 4000 6000 8000 10000 Number of load cycles
Number of load cycles
(a) (b)

Figure B. 7: Resilient strain measured in RLT tests varying with number of cycles of specimens compacted at 8% moisture content with
140 kPa confining pressure for (a) stress sequence 3 and (b) stress sequence 4

B-7
Appendix B

1.5

1.4

Resilient strain (10-3)


Shakedown range C
1.3

1.2

1.1
0 2000 4000 6000 8000 10000
Number of load cycles

Figure B. 8: Resilient strain measured in RLT tests varying with number of cycles of specimens compacted at 8% moisture content with
140 kPa confining pressure for stress sequence

B-8
Appendix B

1.04 1.6

1
1.5
Resilient strain ( 10-3)

Resilient strain ( 10-3)


0.96

Shakedown range B Shakedown range B


1.4

0.92

1.3
0.88

0.84 1.2
0 2000 4000 6000 8000 10000 0 2000 4000 6000 8000 10000
Number of load cycles Number of load cycles

(a) (b)

Figure B. 9: Resilient strain measured in RLT tests varying with number of cycles of specimens compacted at 8% moisture content with
180 kPa confining pressure for (a) stress sequence 1 and (b) stress sequence 2

B-9
Appendix B

2.2
2

1.9
2

Resilient strain (10-3)


Resilient strain ( 10-3)

1.8

1.8
Shakedown range C Shakedown range C
1.7

1.6
1.6

1.5 1.4
0 2000 4000 6000 8000 10000 0 400 800 1200
Number of load cycles Number of load cycles
(a) (b)

Figure B. 10: Resilient strain measured in RLT tests varying with number of cycles of specimens compacted at 8% moisture content with
180 kPa confining pressure for (a) stress sequence 3 and (b) stress sequence 4

B-10
Appendix B

0.8 0.88

0.75
0.84
Resilient strain ( 10-3)

Resilient strain ( 10-3)


0.7
Shakedown range B Shakedown range B
0.8

0.65

0.76
0.6

0.55 0.72
0 2000 4000 6000 8000 10000 0 2000 4000 6000 8000 10000
Number of load cycles Number of load cycles

(a) (b)

Figure B. 11: Resilient strain measured in RLT tests varying with number of cycles of specimens compacted at 6.5% moisture content
with 50 kPa confining pressure for (a) stress sequence 1 and (b) stress sequence 2

B-11
Appendix B

1.04 1.6

1 1.5
Resilient strain ( 10-3)

Resilient strain ( 10-3)


Shakedown range B
0.96 1.4

Shakedown range B
0.92 1.3

0.88 1.2

0.84 1.1
0 2000 4000 6000 8000 10000 0 2000 4000 6000 8000 10000
Number of load cycles Number of load cycles

(a) (b)

Figure B. 12: Resilient strain measured in RLT tests varying with number of cycles of specimens compacted at 6.5% moisture content
with 50 kPa confining pressure for (a) stress sequence 3 and (b) stress sequence 4

B-12
Appendix B

2.8

2.4

Resilient strain ( 10-3) 2


Shakedown range C

1.6

1.2
0 40 80 120 160
Number of load cycles

Figure B. 13: Resilient strain measured in RLT tests varying with number of cycles of specimens compacted at 6.5% moisture content
with 50 kPa confining pressure for stress sequence 5

B-13
Appendix B

0.95 1.08

0.9
1.04

0.85
Resilient strain ( 10-3)

Resilient strain ( 10-3)


1
Shakedown range B
0.8
Shakedown range B
0.96
0.75

0.92
0.7

0.65 0.88
0 2000 4000 6000 8000 10000 0 2000 4000 6000 8000 10000
Number of load cycles Number of load cycles

(a) (b)

Figure B. 14: Resilient strain measured in RLT tests varying with number of cycles of specimens compacted at 6.5% moisture content
with 100 kPa confining pressure for (a) stress sequence 1 and (b) stress sequence 2

B-14
Appendix B

1.4
1.8

1.35

1.7
1.3
Resilient strain ( 10-3)

Resilient strain (10-3)


1.25
1.6

Shakedown range C Shakedown range C


1.2

1.5

1.15

1.1 1.4

0 2000 4000 6000 8000 10000 0 2000 4000 6000 8000 10000
Number of load cycles
Number of load cycles

(a) (b)

Figure B. 15: Resilient strain measured in RLT tests varying with number of cycles of specimens compacted at 6.5% moisture content
with 100 kPa confining pressure for (a) stress sequence 3 and (b) stress sequence 4

B-15
Appendix B

1.6

Resilient strain ( 10-3)


1.2

0.8 Shakedown range C

0.4

0
0 2000 4000 6000 8000
Number of load cycles

Figure B. 16: Resilient strain measured in RLT tests varying with number of cycles of specimens compacted at 6.5% moisture content
with 100 kPa confining pressure for stress sequence 5

B-16
Appendix B

0.24 0.76

0.22
Shakedown range B 0.72

Resilient strain ( 10-3)


Resilient strain ( 10-3)

0.2

0.68

0.18 Shakedown range B

0.64
0.16

0.14 0.6
0 2000 4000 6000 8000 10000 0 2000 4000 6000 8000 10000
Number of load cycles Number of load cycles
(a) (b)

B-17
Appendix B

Figure B. 17: Resilient strain measured in RLT tests varying with number of cycles of specimens compacted at 6.5% moisture content
with 140 kPa confining pressure for (a) stress sequence 1 and (b) stress sequence 2

B-17
Appendix B

1.1 1.3

1.2

1
Resilient strain ( 10-3)

Resilient strain ( 10-3)


1.1
Shakedown range B
Shakedown range C
0.9

0.8 0.9
0 2000 4000 6000 8000 10000 0 2000 4000 6000 8000 10000
Number of load cycles Number of load cycles

(a) (b)

Figure B. 18: Resilient strain measured in RLT tests varying with number of cycles of specimens compacted at 6.5% moisture content
with 140 kPa confining pressure for (a) stress sequence 3 and (b) stress sequence 4

B-18
Appendix B

1.35

1.3

Resilient strain ( 10-3)


1.25

Shakedown range C
1.2

1.15

1.1
0 2000 4000 6000 8000 10000
Number of load cycles

Figure B. 19: Resilient strain measured in RLT tests varying with number of cycles of specimens compacted at 6.5% moisture content
with 140 kPa confining pressure for stress sequence 5

B-19
Appendix B

0.2

0.28

0.16
0.24
Resilient strain ( 10-3)

Resilient strain (10-3)


Shakedown range A
0.12
0.2

Shakedown range A
0.08
0.16

0.04 0.12
0 2000 4000 6000 8000 10000 0 2000 4000 6000 8000 10000
Number of load cycles (a) of load cycles
Number

Figure B. 20: Resilient strain measured in RLT tests varying with number of cycles of specimens compacted at 8% moisture content with
50 kPa confining pressure for (a) stress sequence 1 (b) stress sequence 2

B-20
Appendix B

0.52 0.8

0.48 0.76

0.44 0.72
Resilient strain ( 10-3)

Resilient strain (10-3)


0.4 0.68 Shakedown range B

Shakedown range A/B


0.36 0.64

0.32 0.6

0.28 0.56
0 2000 4000 6000 8000 10000 0 2000 4000 6000 8000 10000
Number of load cycles Number of load cycles

(a) (b)

Figure B. 21: Resilient strain measured in RLT tests varying with number of cycles of specimens compacted at 8% moisture content with
50 kPa confining pressure for (a) stress sequence 3 (b) stress sequence 4

B-21
Appendix B

1.3

1.2

Resilient strain (10-3)


1.1 Shakedown range B

0.9
0 2000 4000 6000 8000 10000
Number of load cycles

Figure B. 22: Resilient strain measured in RLT tests varying with number of cycles of specimens compacted at 8% moisture
content with 50 kPa confining pressure for stress sequence 5

B-22
Appendix B

0.44 0.52
MC = 6.5%, Conf.pre = 50 kPa, 1.5% CF MC = 6.5%, Conf.pre = 50 kPa, 1.5% CF
Stress sequence 1 Stress sequence 2

0.4 0.48
Resilient strain ( 10-3)

Resilient strain (10-3)


Shakedown range A
0.36 0.44

Shakedown range A

0.32 0.4

0.28 0.36
0 2000 4000 6000 8000 10000 0 2000 4000 6000 8000 10000
Number of load cycles Number of load cycles

(a) (b)

Figure B. 23: Resilient strain measured in RLT tests varying with number of cycles of specimens compacted at 6.5% moisture content
with 50 kPa confining pressure for (a) stress sequence 1 (b) stress sequence 2

B-23
Appendix B

1
MC = 6.5%, Conf.pre = 50 kPa, 1.5% CF 1.4 MC = 6.5%, Conf.pre = 50 kPa, 1.5% CF
Stress sequence 3 Stress sequence 4

0.9 1.3
Resilient strain (10-3)

Resilient strain ( 10-3)


1.2
Shakedown range B
0.8 Shakedown range B/C

1.1

0.7

0.6
0.9
0 2000 4000 6000 8000 10000
Number of load cycles 0 2000 4000 6000 8000 10000
Number of load cycles
(a) (b)

Figure B. 24: Resilient strain measured in RLT tests varying with number of cycles of specimens compacted at 6.5% moisture content
with 50 kPa confining pressure for (a) stress sequence 3 (b) stress sequence 4

B-24
Appendix B

2.8
MC = 6.5%, Conf.pre = 50 kPa, 1.5% CF
Stress sequence 5

2.4

Resilient strain (10-3)


Shakedown range C

1.6

1.2
0 1000 2000 3000 4000
Number of load cycles

Figure B. 25: Resilient strain measured in RLT tests varying with number of cycles of specimens compacted at 6.5% moisture content
with 50 kPa confining pressure for stress sequence 5

B-25
Appendix B

0.36 0.32
MC = 5%, Conf.pre = 50 kPa, 1.5% CF
MC = 5%, Conf.pre = 50 kPa, 1.5% CF
Stress sequence 1
Stress sequence 2
0.32
0.3

0.28
Resilient strain ( 10-3)

Resilient strain ( 10-3)


Shakedown range A/B 0.28

0.24

0.26
0.2

0.16 0.24 Shakedown range B

0.12 0.22
0 2000 4000 6000 8000 10000
0 2000 4000 6000 8000 10000
Number of load cycles Number of load cycles

(a) (b)

Figure B. 26: Resilient strain measured in RLT tests varying with number of cycles of specimens compacted at 5% moisture content with
50 kPa confining pressure for (a) stress sequence 1 (b) stress sequence 2

B-26
Appendix B

MC = 5%, Conf.pre = 50 kPa, 1.5% CF MC = 5%, Conf.pre = 50 kPa, 1.5% CF


0.7 0.96 Stress sequence 4
Stress sequence 3

0.92

Resilient strain ( 10-3)


0.6
Resilient strain (10-3)

Shakedown range B Shakedown range B


0.88

0.84

0.5

0.8

0.4 0.76

0 2000 4000 6000 8000 10000 0 2000 4000 6000 8000 10000
Number of load cycles Number of load cycles
(a) (b)

Figure B. 27: Resilient strain measured in RLT tests varying with number of cycles of specimens compacted at 5% moisture content with
50 kPa confining pressure for (a) stress sequence 3 (b) stress sequence 4

B-27
Appendix B

MC = 5%, Conf.pre = 50 kPa, 1.5% CF


1.3 Stress sequence 5

1.2

Resilient strain ( 10-3)


Shakedown range B

1.1

0.9
0 2000 4000 6000 8000 10000
Number of load cycles

Figure B. 28: Resilient strain measured in RLT tests varying with number of cycles of specimens compacted at 5% moisture content with
50 kPa confining pressure for stress sequence 5

B-28
Appendix B

Conf.pressure 50 kPa_MC 8% Conf.pressure 100 kPa_MC 8%


Stress sequence 1 Stress sequence 1 Stress sequence 3
5 Stress sequence 2 25 Stress sequence 2 Stress sequence 4
Stress sequence 3

4 20

Permanent strain (10-3)


Permanent strain (10-3)

3 15

2 10

1 5

0 0
0 2000 4000 6000 8000 10000 0 2000 4000 6000 8000 10000
Number of loading cycles Number of loading cycles
(a) (b)

Figure B. 29: Permanent strains measured in RLT tests varying with number of cycles of specimens compacted at 8% moisture content
with (a) 50 kPa confining pressure and (b) 100 kPa confining pressure

B-29
Appendix B

16
25 Conf.pressure 140 kPa_MC 8% Conf.pressure 180 kPa_MC 8%
Stress sequence 1 Stress sequence 4 Stress sequence 1
Stress sequence 2 Stress sequence 5 Stress sequence 2
Stress sequence 3 Stress sequence 3
20
12

Permanent strain (10-3)


Permanent strain (10-3)

15

10

0
0
0 2000 4000 6000 8000 10000
0 2000 4000 6000 8000 10000
Number of loading cycles Number of loading cycles

(a) (b)

Figure B. 30: Permanent strains measured in RLT tests varying with number of cycles of specimens compacted at 8% moisture content
with (a) 140 kPa confining pressure and (b) 180 kPa confining pressure

B-30
Appendix B

Conf.pressure 50 kPa_MC 6.5%


Stress sequence 1 Stress sequence 3 Conf.pressure 100 kPa_MC 6.5%
Stress sequence 2 Stress sequence 4 40 Stress sequence 1 Stress sequence 3
Stress sequence 2 Stress sequence 4
16

30
12

Permanent strain (10-3)


Permanent strain (10-3)

8 20

4 10

0
0
0 2000 4000 6000 8000 10000
Number of loading cycles 0 2000 4000 6000 8000 10000
Number of loading cycles

(a) (b)

Figure B. 31: Permanent strains measured in RLT tests varying with number of cycles of specimens compacted at 8% moisture content
with (a) 50 kPa confining pressure and (b) 100 kPa confining pressure

B-31
Appendix B

Conf.pressure 140 kPa_MC 6.5%


10 Stress sequence 1 Stress sequence 4
Stress sequence 2 Stress sequence 5
Stress sequence 3

Permanent strain (10-3)


6

0 2000 4000 6000 8000 10000


Number of loading cycles

Figure B. 32: Permanent strains measured in RLT tests varying with number of cycles of specimens compacted at 8% moisture content
with 140 kPa confining pressure

B-32

You might also like