You are on page 1of 204

Research Report

AP-R640-20

Development of Design Procedures


for Lightly Bound Cemented Materials
in Flexible Pavements
Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Prepared by Publisher

James Grenfell, Geoffrey Jameson and Phil Hunt Austroads Ltd.


Level 9, 287 Elizabeth Street
Sydney NSW 2000 Australia
Project Manager Phone: +61 2 8265 3300
austroads@austroads.com.au
Les Marchant
www.austroads.com.au

Abstract About Austroads


There is a need to improve Austroads procedures for the structural Austroads is the peak organisation of Australasian road
design of pavements containing lightly bound cemented (LBC) transport and traffic agencies.
materials.
Austroads’ purpose is to support our member organisations to
This report describes the findings of an investigation of the deliver an improved Australasian road transport network. To
performance of Queensland pavements with LBC bases. It describes succeed in this task, we undertake leading-edge road and
the construction and performance monitoring of two trial sections transport research which underpins our input to policy
from which in situ LBC moduli were estimated. Laboratory testing development and published guidance on the design,
was undertaken to improve understanding of the cracking construction and management of the road network and its
characteristics of LBC materials compared to heavily bound associated infrastructure.
cemented (HBC) materials.
Austroads provides a collective approach that delivers value
Unconfined compressive strength requirements for LBC materials are for money, encourages shared knowledge and drives
proposed. A structural design method was developed for pavement consistency for road users.
containing LBC materials and HBC materials in the post-cracking
phase of life. Austroads is governed by a Board consisting of senior
executive representatives from each of its eleven member
organisations:
Keywords • Transport for NSW
Pavement design, rehabilitation, cement, cementitious, bound, • Department of Transport Victoria
stabilisation, recycling, mix design, field performance, wheel-tracking, • Queensland Department of Transport and Main Roads
fatigue, cracking
• Main Roads Western Australia
• Department for Infrastructure and Transport South Australia
ISBN 978-1-922382-35-1
• Department of State Growth Tasmania
Austroads Project No. TT1897
• Department of Infrastructure, Planning and Logistics
Austroads Publication No. AP-R640-20 Northern Territory

Publication date November 2020


• Transport Canberra and City Services Directorate,
Australian Capital Territory
Pages 202 • Department of Infrastructure, Transport, Regional
Development and Communications
© Austroads 2020 • Australian Local Government Association
• New Zealand Transport Agency.
This work is copyright. Apart from any use as permitted under the
Copyright Act 1968, no part may be reproduced by any process
without the prior written permission of Austroads.

This report has been prepared for Austroads as part of its work to promote improved Australian and New Zealand transport outcomes by
providing expert technical input on road and road transport issues.
Individual road agencies will determine their response to this report following consideration of their legislative or administrative
arrangements, available funding, as well as local circumstances and priorities.
Austroads believes this publication to be correct at the time of printing and does not accept responsibility for any consequences arising from
the use of information herein. Readers should rely on their own skill and judgement to apply information to particular issues.
Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Summary
The addition of a small amount of cementitious binder to non-standard granular materials may result in a
fit-for-purpose base or subbase at a significantly lower cost than crushed rock complying with standard
specifications. Such lightly bound cemented (LBC) materials have particular use in pavement rehabilitation
and heavy patching. LBC are unbound granular materials which have been cementitiously stabilised to
produce a low strength material that is less susceptible to block (ladder) cracking and crocodile cracking than
cemented materials. The binder content and strength of LBC materials are significantly lower than heavily
bound cemented (HBC) materials.

The performance of selected Queensland pavements was reviewed. Long lengths of LBC bases showed
little or no signs of block cracking or crocodile cracking, despite the considerable passage of time and
loading since construction. It was concluded that it is possible to design and construct a low-strength material
that develops fine micro-cracking with sufficient base thickness and subbase support to limit the extent that
micro-cracking leads to macro-cracking.

An Austroads pavement structural design method was developed from:


• field deflection testing to estimate the in situ moduli of LBC bases
• laboratory testing and analysis that confirmed that LBC materials are very susceptible to fatigue and it
would not be appropriate to develop a design method to inhibit fatigue of LBC.

It is proposed that Austroads adopt a 28-day unconfined compressive strength (UCS) range of 1.0–2.0 MPa
for LBC materials based on the advice of the Queensland Department of Transport and Main Roads (TMR)
which has the most experience with these materials. This UCS range is the same as currently used in
Austroads guides. TMR has recommended that additional requirements be included in the definition, namely
that there needs to be a ‘steady’ increase in UCS between 7 and 28 days along with a minimum 7-day UCS
of 1.0 MPa or a degree of saturation requirement if the 7-day UCS < 1.0 MPa. These additional requirements
need consideration in the revision of the Austroads guides. TMR will also consider the use of modified
compaction in the future, possibly with adjusted UCS requirements in the longer term.

A structural design method was developed for pavements containing LBC materials and HBC materials in
the post-fatigue cracking phase of life, including:
• A new elastic characterisation method, applicable to LBC materials and HBC materials in the fatigue
cracked state, including methods to vary the design modulus according to the design modulus of the layer
supporting the cracked material and the thickness and modulus of the overlying bound materials.
• Design charts to select LBC base thicknesses to inhibit the development of block cracking and crocodile
cracking, with the minimum thickness varying with design traffic loading and the support provided by the
layer under the LBC base.

This design method was developed considering the use and performance of LBC for moderate-to-heavily
trafficked roads. Its applicability to the design of lightly trafficked roads needs consideration in the revision of
Austroads guides.

The research highlighted a need for a national test method for the mixing, compacting and curing of UCS
specimens to improve consistency of UCS results across jurisdictions. Given that the use of LBC materials
may increase with the improved design method, there is an increased need for this method lest HBC bases
are used as LBC bases.

Austroads 2020 | page i


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Contents
Summary ......................................................................................................................................................... i
1. Introduction ............................................................................................................................................ 1
1.1 Background ...................................................................................................................................... 1
1.2 Purpose ............................................................................................................................................ 2
1.3 Scope ............................................................................................................................................... 2
1.4 Methodology ..................................................................................................................................... 3
2. Research Needs and Project Scope .................................................................................................... 1
2.1 Introduction ....................................................................................................................................... 1
2.2 Road Agency Needs ........................................................................................................................ 1
2.2.1 Queensland Department of Transport and Main Roads ..................................................... 1
2.2.2 Main Roads Western Australia ............................................................................................ 1
2.2.3 Transport for NSW............................................................................................................... 1
2.2.4 Department of Transport Victoria ........................................................................................ 2
2.2.5 Department of Planning, Transport and Infrastructure South Australia .............................. 2
2.2.6 Department of State Growth Tasmania ............................................................................... 2
2.3 Discussion at Austroads PSWG meeting October 2014 .................................................................. 2
2.3.1 Type of Lightly Bound Material ............................................................................................ 2
2.3.2 Project Objectives................................................................................................................ 3
3. Use of Cement Modified and Lightly Bound Cemented Materials .................................................... 5
3.1 Introduction ....................................................................................................................................... 5
3.2 Queensland Department of Transport and Main Roads .................................................................. 5
3.2.1 New Constructions .............................................................................................................. 6
3.2.2 Pavement Rehabilitations .................................................................................................... 8
3.3 Main Roads Western Australia ......................................................................................................... 9
3.3.1 Rural Floodways .................................................................................................................. 9
3.3.2 Hydrated Cement-treated Crushed Rock Base ................................................................. 10
3.4 Department of Transport Victoria ................................................................................................... 11
3.5 Transport for NSW ......................................................................................................................... 13
3.6 Department of Planning, Transport and Infrastructure, South Australia ........................................ 13
3.7 Department of State Growth Tasmania .......................................................................................... 13
3.8 New Zealand Transport Agency ..................................................................................................... 13
4. Literature Review ................................................................................................................................. 14
4.1 Introduction ..................................................................................................................................... 14
4.2 Distress Modes of Cement-treated Materials ................................................................................. 14
4.2.1 Distress Modes .................................................................................................................. 14
4.2.2 Description of Cracking ..................................................................................................... 15
4.3 Shrinkage Cracking ........................................................................................................................ 17
4.3.1 Introduction ........................................................................................................................ 17
4.3.2 Mechanism of Shrinkage ................................................................................................... 18
4.3.3 Measures to Reduce Shrinkage Cracking of Cemented Materials ................................... 18
4.3.4 Inducing Early-life Micro-cracking to Reduce Shrinkage Cracking ................................... 19
4.4 Fatigue Cracking ............................................................................................................................ 20

Austroads 2020 | page ii


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

4.4.1 Introduction ........................................................................................................................ 20


4.4.2 Laboratory Fatigue Characterisation ................................................................................. 20
4.4.3 Fatigue under Accelerated Loading .................................................................................. 23
4.5 Investigation of New Zealand Stabilised Pavements ..................................................................... 26
4.6 Summary ........................................................................................................................................ 27
5. Performance Review of Selected Queensland Pavements ............................................................. 28
5.1 Introduction ..................................................................................................................................... 28
5.2 Queensland Field Site Investigation ............................................................................................... 28
5.2.1 Methodology ...................................................................................................................... 28
5.2.2 Field Site Investigation: Preliminary Findings ................................................................... 30
5.3 Review of Network Pavement Condition Data ............................................................................... 30
5.3.1 Introduction ........................................................................................................................ 30
5.3.2 Cracking Type ................................................................................................................... 31
5.3.3 Factors Associated with Cracking ..................................................................................... 31
5.4 Detailed Pavement Investigations .................................................................................................. 35
5.4.1 Introduction ........................................................................................................................ 35
5.4.2 Site 2/2A Cunningham Highway (Inglewood – Goondiwindi) ............................................ 35
5.4.3 Site 16 Dawson Highway (Gladstone – Biloela) ................................................................ 42
5.5 Conclusions .................................................................................................................................... 45
6. In situ Modulus of LBC Bases ............................................................................................................ 47
6.1 Introduction ..................................................................................................................................... 47
6.2 Field Trial Site A – Bruce Highway, Collinson’s Lagoon ................................................................ 47
6.2.1 Location ............................................................................................................................. 47
6.2.2 Historical Data and Design Traffic ..................................................................................... 48
6.2.3 Existing Pavement Before Treatment................................................................................ 48
6.2.4 Site Works ......................................................................................................................... 50
6.2.5 Surface Deflection Measurements .................................................................................... 53
6.2.6 FWD Back-calculation of Modulus Values ........................................................................ 56
6.2.7 Trial Site A Conclusions .................................................................................................... 59
6.3 Field Trial Site B – Bruce Highway Barratta Creeks ...................................................................... 60
6.3.1 Location ............................................................................................................................. 60
6.3.2 Existing Pavement Before Treatment................................................................................ 61
6.3.3 Site Works ......................................................................................................................... 62
6.3.4 UCS Testing ...................................................................................................................... 64
6.3.5 Surface Deflection Measurements .................................................................................... 65
6.3.6 Back-calculation of Modulus Values.................................................................................. 68
6.3.7 Trial Site B Conclusions .................................................................................................... 72
7. Conventional Laboratory Characterisation Testing ......................................................................... 73
7.1 Introduction ..................................................................................................................................... 73
7.2 Unconfined Compressive Strength (UCS) Testing ........................................................................ 73
7.3 Indirect Tensile Modulus Testing ................................................................................................... 77
7.4 Indirect Tensile Strength Testing ................................................................................................... 78
7.5 Flexural Modulus Testing ............................................................................................................... 78
7.6 Flexural Strength Testing ............................................................................................................... 80
7.7 Summary ........................................................................................................................................ 82

Austroads 2020 | page iii


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

8. Wheel-Tracking Fatigue Testing of Cemented Materials ................................................................. 83


8.1 Introduction ..................................................................................................................................... 83
8.2 Development of Test Method ......................................................................................................... 83
8.2.1 Introduction ........................................................................................................................ 83
8.2.2 Equipment ......................................................................................................................... 83
8.2.3 Method to Prepare Test Slabs ........................................................................................... 85
8.2.4 Fatigue Test Method.......................................................................................................... 86
8.2.5 Deflection Instrumentation and Monitoring ........................................................................ 87
8.2.6 Fatigue Damage Characterisation..................................................................................... 88
8.3 Application of Test Method ............................................................................................................. 91
8.3.1 Introduction ........................................................................................................................ 91
8.3.2 Test Materials .................................................................................................................... 91
8.3.3 Compaction of Mixtures ..................................................................................................... 93
8.3.4 Curing ................................................................................................................................ 94
8.3.5 Results ............................................................................................................................... 95
8.4 Summary of Findings ..................................................................................................................... 98
9. Structural Design of Pavements Containing Cracked Cemented Materials ................................ 100
9.1 Introduction ................................................................................................................................... 100
9.2 Fatigue Life of LBC Layers ........................................................................................................... 101
9.3 Definition of LBC .......................................................................................................................... 103
9.3.1 Current Austroads Definitions ......................................................................................... 103
9.3.2 TMR Definitions ............................................................................................................... 104
9.3.3 Proposed Austroads Definition of Lightly Bound Cemented Materials ........................... 104
9.3.4 Austroads Test Method for Preparation of UCS Test Specimens ................................... 105
9.4 Elastic Characterisation for Pavement Design ............................................................................. 105
9.4.1 Introduction ...................................................................................................................... 105
9.4.2 Factors Affecting Cracked Cemented Material Moduli .................................................... 106
9.4.3 Proposed Austroads Elastic Characterisation ................................................................. 111
9.5 Inhibiting Macro-cracking of LBC bases ....................................................................................... 112
9.5.1 Introduction ...................................................................................................................... 112
9.5.2 Minimum Support to LBC Base ....................................................................................... 113
9.5.3 Minimum Base Thicknesses ............................................................................................ 113
9.5.4 Design Moduli of Underlying Granular Layers ................................................................ 116
9.6 Use of Mechanistic – Empirical Design Method ........................................................................... 117
9.7 Proposed Changes to the Guide to Pavement Technology ......................................................... 117
10. Summary and Conclusions............................................................................................................... 118
References ................................................................................................................................................. 122
Appendix A Performance Review of Selected Queensland Pavements: Inspections................. 125
A.1 Introduction ................................................................................................................................... 125
A.2 Site 1: Cunningham Highway (Ipswich – Warwick) ...................................................................... 126
Overview .......................................................................................................................... 126
Performance Summary ................................................................................................... 127
A.3 Site 2 and 2A: Cunningham Highway (Inglewood – Goondiwindi)............................................... 129
Overview .......................................................................................................................... 129
Performance Summary ................................................................................................... 129

Austroads 2020 | page iv


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

A.4 Sites 5 and 5A: Malanda – Atherton Road ................................................................................... 131


Overview .......................................................................................................................... 131
Performance Summary ................................................................................................... 131
A.5 Site 6: Flinders Highway (Hughenden – Richmond) .................................................................... 132
Overview .......................................................................................................................... 132
Performance Summary ................................................................................................... 133
A.6 Sites 7: Flinders Highway (Hughenden – Richmond) .................................................................. 135
Overview .......................................................................................................................... 135
Performance Summary ................................................................................................... 136
A.7 Site 8: Bruce Highway (Bowen – Ayr) .......................................................................................... 136
Overview .......................................................................................................................... 136
Performance Summary ................................................................................................... 137
A.8 Site 9: Bruce Highway (Bowen – Ayr) .......................................................................................... 137
Overview .......................................................................................................................... 137
Performance Summary ................................................................................................... 138
A.9 Site 10: North Townsville Road .................................................................................................... 140
Overview .......................................................................................................................... 140
Performance Summary ................................................................................................... 140
A.10 Site 11: Douglas – Garbutt Road (Duckworth St) ........................................................................ 141
Overview .......................................................................................................................... 141
Performance Summary ................................................................................................... 142
A.11 Site 12: Bruce Highway (Townsville – Ingham)............................................................................ 143
Overview .......................................................................................................................... 143
Performance Summary ................................................................................................... 144
A.12 Site 13: Bruce Highway (Townsville – Ingham)............................................................................ 145
Overview .......................................................................................................................... 145
Pavement Summary ........................................................................................................ 145
A.13 Sites 14 & 14A: Bundaberg – Miriam Vale Road ......................................................................... 146
Overview .......................................................................................................................... 146
Performance Summary ................................................................................................... 147
A.14 Site 15 Western Yeppoon – Emu Park Road ............................................................................... 148
Overview .......................................................................................................................... 148
Performance Summary ................................................................................................... 148
A.15 Site 16: Dawson Highway (Gladstone – Biloela) ......................................................................... 149
Overview .......................................................................................................................... 149
Performance Summary ................................................................................................... 150
A.16 Sites 17 & 17A: Dawson Highway (Biloela – Banana) ................................................................. 151
Overview .......................................................................................................................... 151
Performance Summary ................................................................................................... 153
Appendix B Surface Deflections Bruce Highway Collinson’s Lagoon ......................................... 155
Appendix C Surface Deflections Bruce Highway Barratta Creeks ................................................ 158
Appendix D Material Splitting and Mixing Procedure ..................................................................... 161
D.1 Introduction ................................................................................................................................... 161
D.2 Splitting ......................................................................................................................................... 161
D.3 Mixing Procedure ......................................................................................................................... 161

Austroads 2020 | page v


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Appendix E Compaction of Wheel-tracking Slabs .......................................................................... 163


Appendix F Wheel-tracking Fatigue Test Method........................................................................... 165
Appendix G Wheel-tracking Test Deflection Plots .......................................................................... 166
Appendix H Vibrating Table Results ................................................................................................ 177

Tables

Table 3.1: Presumptive values for elastic characterisation of LBC materials ................................................. 7
Table 3.2: Minimum thickness and strength of unbound granular pavement materials required
below the LBC layer ....................................................................................................................... 8
Table 3.3: Target UCS to be used to select the optimum content of cementitious binder ............................11
Table 4.1: Typical crack patterns observed at deep-lift stabilised sites ........................................................16
Table 4.2: Pavement materials subject to accelerated loading .................................................................... 23
Table 5.1: Composition of Sites 2 and 2A ..................................................................................................... 36
Table 5.2: Summary of Coring Investigation at Site 2 ................................................................................... 38
Table 5.3: Summary of Coring Investigation at Site 2A ................................................................................ 38
Table 5.4: Summary of UCS data of cores ................................................................................................... 41
Table 5.5: Composition of Sites 16 ............................................................................................................... 42
Table 5.6: Summary of Coring Investigation at Site 16 ................................................................................. 44
Table 6.1: Design traffic summary ................................................................................................................ 48
Table 6.2: Summary of the construction phases ........................................................................................... 50
Table 6.3: TMR UCS results of laboratory cylinders at Bruce Highway Collinson’s Lagoon ........................52
Table 6.4: Contractor’s UCS results at Bruce Highway Collinson’s Lagoon.................................................52
Table 6.5: Pavement structure and seed moduli with sprayed seal surface.................................................56
Table 6.6: Pavement structure and seed moduli with an asphalt surfacing ..................................................56
Table 6.7: 28-day back-calculated moduli between wheelpaths .................................................................. 57
Table 6.8: 90-day back-calculated moduli between wheelpaths .................................................................. 57
Table 6.9: 1 year back-calculated moduli between wheelpaths.................................................................... 58
Table 6.10: 28-day back-calculated moduli outer wheelpath .......................................................................... 58
Table 6.11 90-day back-calculated moduli outer wheelpath .......................................................................... 59
Table 6.12: 1 year back-calculated moduli outer wheelpath ........................................................................... 59
Table 6.13: Field Trial Site B design traffic summary ..................................................................................... 61
Table 6.14: UCS results obtained by TMR at Bruce Highway – Barratta Creeks ...........................................64
Table 6.15: Summary of UCS test results measured by the contractor at Barratta Creeks ...........................65
Table 6.16: Pavement structure and seed moduli .......................................................................................... 68
Table 6.17: Back-calculated moduli between wheelpaths after 28 days ........................................................68
Table 6.18: Back-calculated moduli between wheelpaths after 90 days ........................................................69
Table 6.19: Back-calculated moduli between wheelpaths after 1 year ...........................................................69
Table 6.20: Back-calculated moduli in outer wheelpath after 28 days ............................................................ 70
Table 6.21: Back-calculated moduli in outer wheelpath after 90 days ............................................................ 71
Table 6.22: Back-calculated moduli in outer wheelpath after 1 year .............................................................. 71
Table 7.1: UCS results .................................................................................................................................. 74
Table 8.1: Parameters determined from the deflection magnitude signals ...................................................91
Table 8.2: Properties of selected materials ................................................................................................... 92
Table 8.3: Slabs tested for fatigue damage characteristics .......................................................................... 93
Table 8.4: Maximum dry densities and optimum moisture contents of the stabilised crushed hornfels .......93
Table 8.5: Maximum dry densities and optimum moisture contents for the stabilised
Barratta Creeks gravel ................................................................................................................. 93
Table 8.6: Density results for test slabs ........................................................................................................ 94
Table 8.7: Vibrating table results ................................................................................................................... 95
Table 8.8: Summary of deflection data from trafficked slabs ........................................................................ 96
Table 8.9: Summary of average deflection data trafficked slabs of comparable cement contents
and curing times ........................................................................................................................... 96

Austroads 2020 | page vi


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Table 9.1: Summary of road agency the minimum strength requirements of HBC materials .....................103
Table 9.2: Suggested maximum vertical modulus of cracked cemented material bases and subbases ...112
Table E 1: Typical layer compaction sequence ........................................................................................... 164

Figures

Figure 2.1: Fatigue cracking of thick cemented material bases ....................................................................... 4


Figure 3.2: Great Eastern Highway floodways ................................................................................................. 9
Figure 3.3: Minor cracking at Site 2 observed 8 years after construction ......................................................10
Figure 3.4: Unconfined compressive strength (UCS) of unsoaked samples after 28 days curing.................12
Figure 4.1: Use of chemical binders ............................................................................................................... 14
Figure 4.2: Distress modes of chemically-treated pavements ....................................................................... 15
Figure 4.3: Causes of cracking ....................................................................................................................... 15
Figure 4.4: Illustrations of cracking pattern types ........................................................................................... 17
Figure 4.5: Shrinkage mechanism of cement paste ....................................................................................... 18
Figure 4.6: Micro-cracking of soil-cement ...................................................................................................... 20
Figure 4.7: Beam flexural fatigue test results ................................................................................................. 21
Figure 4.8: Typical pavement cross-section of pavements subject to accelerated loading ...........................22
Figure 4.9: Pavement fatigue lives predicted for the CAPTIF pavements using 2000 MPa modulus ...........22
Figure 4.10: Change in measured vertical subgrade strains with loading cycles ............................................24
Figure 4.11: Change in back-calculated moduli with loading cycles ................................................................ 25
Figure 4.12: Site 12 – Hirini Street Gisborne.................................................................................................... 26
Figure 4.13: Variation in back-calculated moduli of 3% cemented materials ...................................................27
Figure 5.1: Locations of field sites .................................................................................................................. 29
Figure 5.2: Crack behaviour in the different wheelpath locations .................................................................. 31
Figure 5.3: Crack percentage as a function of maximum deflection D0 .........................................................32
Figure 5.4: Crack percentage as a function of curvature (D0-D200) ................................................................ 32
Figure 5.5: Total crack length versus LBC base thickness ............................................................................ 33
Figure 5.6: Longitudinal crack length versus LBC base thickness ................................................................. 34
Figure 5.7: Transverse crack length versus LBC base thickness .................................................................. 34
Figure 5.8: Crocodile crack length versus LBC base thickness ..................................................................... 35
Figure 5.9: 7-day UCS testing used to select the binder content ................................................................... 36
Figure 5.10: Examples of cracking at Site 2 ..................................................................................................... 37
Figure 5.11: Coring rig set-up ........................................................................................................................... 37
Figure 5.12: Core extracted .............................................................................................................................. 37
Figure 5.13: Successfully extracted core ......................................................................................................... 39
Figure 5.14: Example core hole ....................................................................................................................... 40
Figure 5.15: Example of disintegrated of the LBC base observed after coring................................................40
Figure 5.16: Example of LBC base layer not able to be extracted in tact ........................................................41
Figure 5.17: Dawson Highway inspection in 2016 ........................................................................................... 43
Figure 5.18: Core hole showing large aggregate size ...................................................................................... 44
Figure 5.19: Broken down material extracted from core hole .......................................................................... 44
Figure 5.20: Material extracted from core hole................................................................................................. 45
Figure 5.21: Partial core extracted with layers of intact material...................................................................... 45
Figure 6.1: Aerial view of Bruce Highway at Collinson’s Lagoon ................................................................... 47
Figure 6.2: Pavement composition before in situ stabilisation ....................................................................... 49
Figure 6.3: Maximum deflections and curvatures under 40 kN FWD load in 2014 ........................................49
Figure 6.4: Pavement composition after in situ stabilisation and surfacing ...................................................51
Figure 6.5: Comparison of maximum deflections between the wheelpaths ...................................................53
Figure 6.6: Comparison of maximum deflections in the outer wheelpath ......................................................54
Figure 6.7: Comparison of curvatures between the wheelpaths .................................................................... 55
Figure 6.8: Comparison of curvatures in the outer wheelpath ....................................................................... 55
Figure 6.9: Aerial view of the trial section....................................................................................................... 60
Figure 6.10: Pavement composition before in situ stabilisation ....................................................................... 61
Figure 6.11: Measured maximum deflections and curvatures at Bruce Highway – Barratta Creeks
prior to stabilisation ...................................................................................................................... 62

Austroads 2020 | page vii


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 6.12: Pavement composition after stabilisation ..................................................................................... 63


Figure 6.13: Maximum deflections between wheelpaths on Bruce Highway – Barratta Creeks section .........66
Figure 6.14: Maximum deflections in outer wheelpath: Bruce Highway – Barratta Creeks section.................66
Figure 6.15: Curvature between wheelpaths: Bruce Highway – Barratta Creeks section ...............................67
Figure 6.16: Curvatures in the outer wheelpath: Bruce Highway – Barratta Creeks section ...........................67
Figure 6.17: LBC base modulus between wheelpaths ..................................................................................... 70
Figure 6.18: LBC base modulus in outer wheelpath ........................................................................................ 72
Figure 7.1: UCS of stabilised crushed hornfels .............................................................................................. 75
Figure 7.2: UCS of stabilised Barratta Creeks gravel .................................................................................... 75
Figure 7.3: Stress-strain relationship during 28-day UCS testing .................................................................. 76
Figure 7.4: Stress-strain relationship during 90 day UCS testing .................................................................. 76
Figure 7.5: Stabilised crushed hornfels indirect tensile modulus values ........................................................77
Figure 7.6: Stabilised crushed hornfels indirect tensile strength values ........................................................78
Figure 7.7: Stabilised crushed hornfels flexural modulus values ................................................................... 79
Figure 7.8: Stabilised Barratta Creeks gravel flexural modulus values ..........................................................79
Figure 7.9: Flexural beam testing in the laboratory ........................................................................................ 80
Figure 7.10: Stabilised crushed hornfels flexural strengths ............................................................................. 81
Figure 7.11: Stabilised Barratta Creeks gravel flexural strengths .................................................................... 81
Figure 8.1: Extra-large wheel tracking device ................................................................................................ 84
Figure 8.2: Schematic view of test mould....................................................................................................... 84
Figure 8.3: Rolling wheel loading ................................................................................................................... 85
Figure 8.4: Planetary concrete mixer used to mix stabilised materials ..........................................................85
Figure 8.5: Austrack Extra Large Wheel-tracking device ............................................................................... 86
Figure 8.6: Schematic view and photographs of the deflection monitoring system developed
for wheel tracking slab ................................................................................................................. 87
Figure 8.7: Examples of visual condition at the end of testing ....................................................................... 88
Figure 8.8: Vibrating table with support cradle ............................................................................................... 89
Figure 8.9: Example of measured slab elastic and permanent deflections ....................................................89
Figure 8.10: Example of variation in deflections per loading cycles with number of loading cycles ................90
Figure 8.11: Analysing peak to peak deflections from the centre sensor under the wheelpath .......................90
Figure 8.12: Particle size distribution of size 20 mm crushed hornfels ............................................................ 92
Figure 8.13: Particle size distribution of Barratta Creeks gravel ...................................................................... 92
Figure 8.14: Overview of XL-WT testing for stabilised crushed hornfels .........................................................97
Figure 8.15: Overview of XL-WT testing for stabilised Barratta Creeks gravel ................................................97
Figure 8.16: Stabilised crushed hornfels slab deflections ................................................................................ 98
Figure 9.1: Variation in 7-day flexural strength with cement content ...........................................................101
Figure 9.2: Example of LBC base thickness requirement to inhibit fatigue cracking ...................................102
Figure 9.3: Modulus reduction of 3% cement-treated crushed rock subbase on a sand subbase ..............107
Figure 9.4: Experiment 3310 back-calculated HBC materials at the end of the loading ..............................107
Figure 9.5: Proposed cracked cemented material modulus reduction factors with overlaying
thickness of bound materials...................................................................................................... 108
Figure 9.6: Modulus reduction of 3% cement-treated crushed rock subbase on clay subgrade .................109
Figure 9.7: Limits on design moduli cracked cemented materials ............................................................... 110
Figure 9.8: Concept of load transfer at joints in concrete pavements ..........................................................114
Figure 9.9: Example variation in allowable traffic loading in terms of erosion with concrete base
thickness .................................................................................................................................... 114
Figure 9.10: Minimum LBC base thicknesses for pavements with thin bituminous surfacings ......................115
Figure 9.11: Minimum LBC base thickness for granular subbase with E = 150 MPa ....................................116
Figure 9.12: Suggested maximum vertical modulus of top sublayer of normal standard base material........117
Figure A 1: Locations of field monitoring sites .............................................................................................. 126
Figure A 2: Site 1 Cunningham Highway ...................................................................................................... 128
Figure A 3: Cunningham Highway (Inglewood – Goondiwindi) .................................................................... 129
Figure A 4: Site 2 inspection 2016 ................................................................................................................ 130
Figure A 5: Site 5 Malanda – Atherton Road ................................................................................................ 131
Figure A 6: Site 6 Flinders Highway ............................................................................................................. 132
Figure A 7: Site 6 UCS testing ...................................................................................................................... 133

Austroads 2020 | page viii


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure A 8: Site 6 inspection ......................................................................................................................... 134


Figure A 9: Geotextile reinforced seal to inhibit cracking ............................................................................. 135
Figure A 10: Sites 7 Flinders Highway ............................................................................................................ 135
Figure A 11: Site 7 block cracking .................................................................................................................. 136
Figure A 12: Site 8 Bruce Highway Bowen to Ayr .......................................................................................... 137
Figure A 13: Site 8 inspection ......................................................................................................................... 137
Figure A 14: Site 9 Bruce Highway Bowen to Ayr .......................................................................................... 138
Figure A 15: Site 9 inspection ......................................................................................................................... 139
Figure A 16: Site 10 North Townsville Road ................................................................................................... 140
Figure A 17: Site 10 inspection ....................................................................................................................... 141
Figure A 18: Site 11 Douglass – Garbutt Road .............................................................................................. 142
Figure A 19: Site 11 inspection ....................................................................................................................... 143
Figure A 20: Site 12 Bruce Highway Townsville to Ingram ............................................................................. 143
Figure A 21: Site 12 inspection ....................................................................................................................... 144
Figure A 22: Site 13 Bruce Highway Townsville to Ingram ............................................................................. 145
Figure A 23: Site 13 inspection ....................................................................................................................... 146
Figure A 24: Sites 14 and 14A Bundaberg-Miriam Vale Road ....................................................................... 147
Figure A 25: Site 14A inspection .................................................................................................................... 147
Figure A 26: Site 15 Western Yeppoon-Emu Park Road ............................................................................... 148
Figure A 27: Site 15 inspection ....................................................................................................................... 149
Figure A 28: Site 16 Dawson Highway Gladstone to Biloela .......................................................................... 149
Figure A 29: Site 16 inspection ....................................................................................................................... 150
Figure A 30: Site 17 and 17A Dawson Highway Biloela to Banana ............................................................... 151
Figure A 31: Project construction records sub-sections 1A to 1G .................................................................. 152
Figure A 32: Project construction records sub-sections 2A to 3G .................................................................. 152
Figure A 33: Site 17 inspection ....................................................................................................................... 154
Figure B 1: 28-day measured deflections ..................................................................................................... 155
Figure B 2: 90-day measured deflections ..................................................................................................... 156
Figure B 3: 1-year measured deflections ...................................................................................................... 157
Figure C 1: 28-day measured deflections ..................................................................................................... 158
Figure C 2: 90-day measured deflections ..................................................................................................... 159
Figure C 3: 1-year measured deflections ...................................................................................................... 160
Figure D 1: Motorised rotary splitter and splitting process............................................................................ 161
Figure D 2: Planetary concrete mixer used to mix cemented material .........................................................162
Figure F 1: Location of rut depth measuring points ...................................................................................... 165
Figure G 1: Slab 5555 deflection change with loading cycles....................................................................... 166
Figure G 2: Slab 5568 deflection change with loading cycles....................................................................... 166
Figure G 3: Slab 5688 deflection change with loading cycles....................................................................... 167
Figure G 4: Slab 4833 deflection change with loading cycles....................................................................... 167
Figure G 5: Slab 4866 deflection change with loading cycles....................................................................... 168
Figure G 6: Slab 4932 deflection change with loading cycles....................................................................... 168
Figure G 7: Slab 5029 deflection change with loading cycles....................................................................... 169
Figure G 8: Slab 5383 deflection change with loading cycles....................................................................... 169
Figure G 9: Slab 5492 deflection change with loading cycles....................................................................... 170
Figure G 10: Slab 5694 deflection change with loading cycles....................................................................... 170
Figure G 11: Slab 5922 deflection change with loading cycles....................................................................... 171
Figure G 12: Slab 5988 deflection change with loading cycles....................................................................... 171
Figure G 13: Slab 6028 deflection change with loading cycles....................................................................... 172
Figure G 14: Slab 5945 deflection change with loading cycles....................................................................... 172
Figure G 15: Slab 5951 deflection change with loading cycles....................................................................... 173
Figure G 16: Slab 4811 deflection change with loading cycles....................................................................... 173
Figure G 17: Slab 4876 deflection change with loading cycles....................................................................... 174
Figure G 18: Slab 5469 deflection change with loading cycles....................................................................... 174
Figure G 19: Slab 5541 deflection change with loading cycles....................................................................... 175
Figure G 20: Slab 5928 deflection change with loading cycles....................................................................... 175
Figure G 21: Slab 6095 deflection change with loading cycles....................................................................... 176

Austroads 2020 | page ix


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

1. Introduction

1.1 Background
Austroads member agencies face significant challenges in the provision of a safe, resilient and productive
road network to meet increasing road user demands. The scarcity of quality, cost-effective road-building
materials is an increasing national challenge, particularly in the context of the lack of funding for road
maintenance and rehabilitation. In many instances, the use of alternative approaches, such as treating
granular materials with chemical binders, is becoming commonplace as reserves of high-quality
manufactured granular materials are exhausted or hauling such materials over long distances is
cost-prohibitive.

One output from Austroads project TT1664 Cemented Materials Characterisation was a recommendation to
improve design procedures for granular materials stabilised with cementitious binder contents of 3% or more
(Austroads 2014). To complement that work, road agencies and industry identified a need for improved
guidance for the design of pavements using materials with lower binder contents. The addition of a small
amount of cementitious binder to non-standard granular materials may result in a fit-for-purpose base or
subbase at a significantly lower cost than the use of crushed rock complying with standard specifications.
Cementitious binders may also be added to gravels complying with standard specifications improve shear/rut
resistance in certain traffic and/or high rainfall environments. Such lightly bound cemented (LBC) materials
have particular use in pavement rehabilitation and heavy patching.

The definition of LBC materials, given in Austroads (2017) p55, is as follows:


Lightly bound materials are granular materials to which moderate amounts of stabilising
binder have been added to improve modulus, and where an increase in tensile capacity
may occur. Lightly bound granular materials may exhibit behaviour between modified
granular materials and more heavily bound cemented materials. While the properties of
lightly bound materials and their response to loading is the subject of further research, it is
currently common practice to categorise materials with a 28-day UCS of 1.0 to 2.0 MPa as
lightly bound.

This UCS criteria relates to test specimens prepared using 100% standard Proctor compactive effort at
100% standard optimum moisture content, normal curing for a minimum 28 days in moist condition without
soaking in water (Austroads 2019b).

As distinct from LBC materials, the definition of modified granular materials (Austroads 2017a p54) is as
follows:
Modified granular materials are granular materials to which small amounts of stabilising
agents have been added to improve modulus or to correct other deficiencies in properties
without causing a significant increase in tensile capacity (which could lead to crack
propagation). Modification of poor or marginal materials can yield improved material
properties at significantly less cost than the use of premium materials. Modified materials
have a maximum 28-day compressive strength (UCS) of 1.0 MPa, tested after moist curing
but without soaking at 100% standard Proctor maximum dry density and optimum moisture
content.

This UCS criteria also relates to test specimens prepared using 100% standard Proctor compactive effort at
100% standard optimum moisture content, normal curing for a minimum 28 days in moist condition without
soaking in water (Austroads 2017a).

Austroads 2020 | page 1


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

This report is concerned with a particular type of lightly bound material in which only cementitious binders are
used. In this report these are referred to as LCM materials, although, as discussed in this report, it is
proposed to change the 28-day UCS range to 1.0–2.0 MPa consistent with Austroads (2017). In addition,
when significant amounts of binder are used, it is proposed to refer to these as heavily bound cemented
(HBC) materials. Both LBC and HBC are types of cemented materials.

While both LBC and HBC materials may be susceptible to fatigue cracking their condition in the
post-cracking phase of life may differ. As described in Section 8.2.4 (p114) of Austroads (2017):
A post-cracking phase of the design life can only be considered in the mechanistic
thickness design calculations if cracking from the fatigued cemented material does not
reflect through to the surface. To reduce the risk of reflection cracking the pavement should
provide a minimum cover equivalent to 175 mm of asphalt over the cemented material.

The above text relates to HBC materials which have sufficient binder to produce a bound layer with
significant tensile strength. Such cemented materials will commonly have 28-day UCS values above 2 MPa
but it is beyond the scope of this report to develop a minimum UCS value for HBC materials. Research was
required as to whether LBC materials can be designed to have cracking severities markedly less than for
HBC materials. Such LBC bases may have a useful service life in the post-cracking phase of life without the
need for a minimum 175 mm cover: periodic placement of a strain alleviating membrane (SAM) sprayed seal
can be sufficient to provide a serviceable pavement. Austroads (2019a) provides guidance on thin surfacing
treatments which inhibit reflection cracking.

Unlike pavements with HBC bases, pavements with LBC bases would not be designed to inhibit fatigue
because uneconomic thicknesses would be required due to their low strength. Accordingly, LBC bases would
be thinner than HBC bases with an associated lower initial construction cost.

Whilst such LBC materials are currently used throughout Australasia, design procedures are not
well-founded. In some instances this lack of knowledge has led to cracking despite the design intent as
highlighted in a New Zealand Transport Agency report (Gray et al. 2011) and as reported by some Australian
road agencies (Section 3). Road agencies advised (Section 2) that LBC bases would be more widely used if
the cracking could be more reliably limited.

The development of an Austroads mechanistic-empirical (ME) thickness design method would provide a
quantifiable means of comparing low-cost designs incorporating LBC materials to more traditional pavement
configurations. These developments would inform appropriate risk management strategies in the use of LBC
bases in relation to surface cracking.

Accordingly, in 2014 Austroads contracted the Australian Road Research Board (ARRB) to undertake
research project TT1897 Increasing the Use of Low-cost Modified Granular Materials in New and
Rehabilitated Pavements.

1.2 Purpose
The purpose of project TT1897 was to improve understanding of the mechanisms of crack formation
associated with LBC materials and develop Austroads guidance in terms of the pavement design. This report
describes the research undertaken and the proposed thickness design method.

1.3 Scope
There are a number of possible reasons why LBC materials crack in road pavements, including shrinkage
cracking and reflective cracking from underlying expansive subgrades. The principal focus of the research
undertaken in project TT1897 was load-induced fatigue cracking.

Austroads 2020 | page 2


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

1.4 Methodology
The research included:
• identifying the research needs and project scope based on discussions with Australian and New Zealand
road agencies (Section 2)
• summarising what was known about the use and performance of LBC materials by road agencies at the
commencement of the project (Section 3)
• reporting and reviewing previous research findings (Section 4)
• reviewing the performance of selected Queensland LBC bases with thin bituminous surfacings (Section 5)
• monitoring and analysing the findings of field trials addressing the early life in situ moduli of LBC bases
(Section 6)
• undertaking laboratory testing to improve understanding of the fatigue characteristics of LBC materials
(Section 7 and Section 8)
• developing a framework for the design of flexible pavements containing LBC materials (Section 9)
• summarising the research findings and reporting the conclusions (Section 10).

Austroads 2020 | page 3


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

2. Research Needs and Project Scope

2.1 Introduction
To assist in defining the project scope and tasks, road agency views were sought. These are reported below
together with the outcomes of a meeting of the Austroads Pavement Structures Working Group (APSWG) in
October 2014.

APSWG agreed that there was potential to increase the use of bases stabilised with low quantities of
cementitious binders to improve the performance of nonconforming granular materials used as bases in rural
areas and to provide higher-modulus layers under thin asphalt surfacings in urban roads. However, the field
performance of these LBC bases throughout Australia and New Zealand had been variable. Hence the
importance of this research project.

2.2 Road Agency Needs

2.2.1 Queensland Department of Transport and Main Roads

TMR identified two key research needs for its LBC, both plant-mixed and in situ stabilised:
• improved structural design procedures, if necessary, including an ability to design for fatigue cracking
• improved mix design procedures, particularly the influence of unbound granular material properties on
LBC performance.

TMR considered that there was potential for an increased use of LBC for moderately trafficked urban
pavements in wetter climates where the use of unbound granular bases was risky and premature thin asphalt
fatigue an issue. Thin asphalt-surfaced LBC bases would provide a lower-cost pavement than full-depth
asphalt. The use of LBC bases on rural highways would increase if the overall pavement thickness could be
reduced compared to an unbound granular pavement.

TMR did not consider there was an application for hydrated cement-treated crushed rock base (HCTCRB)
(Section 2.2.2) in Queensland.

The performance of Queensland LBC bases was investigated as part of the project as detailed in Section 5.

2.2.2 Main Roads Western Australia

Main Roads Western Australia (MRWA) were satisfied with the performance of their process of in situ
modification using 1.5–2% type LH (low heat) binder (65% slag/35% cement) and further research related to
this treatment was a low priority.

MRWA’s principal research need related to improving the design processes for thin asphalt-surfaced
HCTCRB pavements for heavily trafficked urban road applications. Such pavements have considerably lower
whole-of-life costs than full-depth asphalt, deep-strength asphalt or concrete pavement types.

2.2.3 Transport for NSW

The main concerns of Transport for NSW (TfNSW) with respect to in situ cementitious stabilisation related to
construction practices, including spread rate control and variable depth of mixing.

Austroads 2020 | page 1


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

However, in some parts of NSW, TfNSW asset managers were seeking to develop treatments for marginal
materials that result in improved resistance to shoving without shrinkage cracking. Methods to assess the
minimum or maximum amount of linear shrinkage have yet to be developed.

TfNSW did not consider there was an application for HCTCRB (Section 2.2.2) in New South Wales.

2.2.4 Department of Transport Victoria

The Department of Transport (DoT) Victoria considered there was potential for an increase in the use of LBC
materials if the design processes were improved. In terms of binder types, over recent years there has been
increased interest in proprietary products including the use of bitumen emulsion. Whilst DoT would like these
binders to be considered further, DoT staff understood that the project funding was not sufficient to
investigate all binder types.

DoT did not consider there was an application for HCTCRB (Section 2.2.2) in Victoria.

2.2.5 Department of Planning, Transport and Infrastructure South Australia

The Department of Planning, Transport and Infrastructure’s (DPTI) main interest was in situ stabilisation of
granular pavements in remote rural locations, particularly with marginal materials. If there were more
confidence in design/performance, then it would lead to greater use in other locations.

DPTI considered the main research need was to improve design processes that reduce the risk of shrinkage
and fatigue cracking.

2.2.6 Department of State Growth Tasmania

The Department of State Growth (DSG) considered there was potential for an increase in the use of LBC
materials if design processes could be improved to risk the shrinkage, and fatigue cracking reflecting through
thin bituminous surfacings.

DSG did not consider there was an application for HCTCRB (Section 2.2.2) in Tasmania.

2.3 Discussion at Austroads PSWG meeting October 2014


The scope of the project was discussed at the October 2014 meeting of the APSWG.

2.3.1 Type of Lightly Bound Material

APSWG agreed that the most common reason to chemically treat granular pavement materials was to
enhance the shear resistance of unbound granular bases with thin bituminous surfacings. Such treatments
include adding moderate amounts of stabilising binders to unbound granular materials to improve modulus,
and where an increase in tensile capacity may occur.

There are a wide range of binders that have been used, including cementitious binders, lime, bituminous
binders, and a wide range of proprietary products. As the project resources were limited, advice was sought
from road agencies regarding priorities for research needs that would lead to an increase in the use of lightly
stabilised materials. The consensus was that the highest priority was stabilisation with cementitious binders,
which in this report are referred to as LBC materials.

Austroads 2020 | page 2


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

2.3.2 Project Objectives

APSWG agreed that concerns about surface cracking were limiting the use of LBC materials. The most likely
causes of cracking are:
• shrinkage of the treated pavement material after construction
• shrinkage of reactive subgrades
• load-induced fatigue cracking.

Shrinkage cracking of cementitiously treated granular materials has also been extensively researched for
HBC materials. As discussed in Section 4.3, Austroads (2017) provides advice regarding measures to
reduce the severity and intensity of shrinkage cracking of HBC materials and also surfacing treatments to
inhibit reflection cracking. It is likely that much of this guidance is also applicable to granular materials lightly
bound with 1–2% cementitious binders. Also, existing surface treatments that have proved adequate for HBC
bases should be equally or more effective for LBC bases.

APSWG agreed that the project scope should not be extended to cover the cracking of LBC bases due to
reactive subgrades. This subject is better addressed in a separate research project across all pavement
types.

Accordingly, APSWG proposed that limited research in relation to shrinkage cracking not be undertaken. It
was agreed that the applicability of this existing guidance for HBC materials be reviewed by:
• collating and reviewing road agency performance data
• if necessary, undertaking short-term field trials to validate guidance on surfacing treatments
(e.g. effectiveness of SAM and geotextile reinforced seals).

Consideration would be given to whether specifications should include drying shrinkage limits.

It was agreed that the major focus of the research in this project should be load-induced cracking of LBC.
Due to their low strengths, it was anticipated that the research would confirm that LBC pavements have very
short fatigue lives. Based on advice of practitioners in Queensland, Western Australia and New Zealand, it
seems that, although the strength/modulus of these LBC materials quickly reduces under traffic loading, the
provision of thin bituminous surfacing treatments is usually sufficient to provide a useful service life after the
modulus has reduced. In this regard, the nature of the fatigue cracking of LBC materials after a reduction in
strength/modulus appears different from the block and ladder cracking typically associated with thick,
cemented materials (Figure 4.4). Consequently, a key research objective was to gain a better understanding
of the condition of LBC post-fatigue cracking. APSWG agreed that a desirable project outcome would be the
development of a design method that reduced the risk of fatigue of LBC base and the resulting unacceptable
surface cracking.

Similar to the findings of New Zealand research (Section 4.4.3), a key objective of the Austroads project was
also to gain a better understanding of the continuum from cement-modified materials to LBC materials to
HBC materials. In particular, an improved understanding was required of the variation in the extent and
severity of surface cracking with binder content. As shown in Figure 2.1, thin bituminous-surfaced HBC
bases at their end of fatigue life may have large blocks of cracked cemented materials. Such cracking has
not been reported by practitioners for LBC bases (see Section 5) which are thinner, have lower flexural
strengths, and shorter fatigue lives.

Austroads 2020 | page 3


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 2.1: Fatigue cracking of thick cemented material bases

Source: Austroads (2008).

Austroads 2020 | page 4


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

3. Use of Cement Modified and Lightly Bound


Cemented Materials

3.1 Introduction
In scoping the project in 2014–15, advice was sought from Australian and New Zealand road agencies on
their use of cement modified granular materials and LBC materials. Of particular interest was whether they
could identify a list of LBC projects that could be investigated during the project. Rural highway projects with
the following characteristics were requested:
• in situ stabilised granular bases under thin bituminous surfacings
• stabilisation using 1–2% of lime or medium-to-slow-setting cementitious binders
• with three or more years of moderate-to-heavy trafficking.

Road agency uses of cementitiously modified and LBC materials are summarised below.

3.2 Queensland Department of Transport and Main Roads


TMR has adopted cement stabilisation of pavement materials on an intermittent basis for many years. In
recent years, the use of LBC bases has been more frequent in some regions, in particular North
Queensland.

In relation to LBC materials the following key applications and drivers were identified:
1. New constructions
Pavements in both wet environments and dry environments (with lower-level drainage) have experienced
problems in terms of the specification of unbound pavement base materials. In conjunction with good
cross-section design (drainage, full-width homogenous materials, etc.) a LBC is often utilised to
eliminate/minimise the moisture sensitivity of unbound gravels. Low amounts of cement have been used
to minimise cracking effects. For design purposes, the LBC is treated as a good-quality unbound
pavement layer.
Some experienced practitioners and designers also made the unproven (or un-tested) comment that
pavements designed in accordance with the current material specifications do not appear to be
withstanding the loading from the current legal heavy vehicles. It is suspected that pavements are
suffering from both overloading and a high number of quick axle repetitions from B-doubles, road trains,
and other approved configurations. With the failure mode often being a rut/shove type failure there is a
suspicion that shear failure from heavy loads could be a major contributor. It is often difficult to determine
whether elevated pavement moisture levels have pre-existed (migrated into the pavement somehow
before failure) or have subsequently entered via rainfall post a loading shear failure.
New construction projects typically use a controlled mixing process in a plant (pugmill) with the mixed
material placed and compacted as part of a designed pavement structure. Base quality gravel is typically
mixed with 1.5–2.0% of type general blend (GB) cement and used as a base layer (on top of a
substructure designed in accordance with Austroads (2017) and TMR (2018). Generally, construction is
conducted according to MRTS10 Plant-mixed Lightly Bound Pavements (TMR 2019). Some district
variations exist, in particular the slurry requirements between cementitious layers, where applicable.
The underlying pavement (subbase) layer materials vary, with unbound granular (gravel),
cement-stabilised subbase (Category 2) and cement-modified all in use.

Austroads 2020 | page 5


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

In certain clay subgrade areas, lime-stabilised subgrades have also been noted and rock treatments over
swampy ground sometimes apply. In areas where the grade line is elevated with embankment works
required, select fill subgrade materials (300 mm total layer) with a minimum soaked CBR of 10–15% often
apply. Otherwise, natural materials are compacted to 97% standard maximum dry density.
2. Rehabilitation projects
Where sufficient funds allow, projects are designed for a 20-year design traffic loading. This often
includes widening existing pavements to meet current geometric standards or widening to cater for
additional gravel overlay material.
Typical applications, after widening works to match the existing pavement, include:
• the existing pavement remains ‘as-is’ and is overlaid with new plant-mixed LBC
• the existing pavement is in situ modified with cementitious additive and overlaid with a new unbound
base (gravel) or plant-mixed LBC (that is, the existing pavement becomes a lightly bound subbase)
• the existing pavement is overlaid with gravel to a defined depth and then mixed with the in situ
pavement material to provide a total LBC depth of between 150 mm and 250 mm.
The driver for the use of LBC is either as per (1) above and/or to maximise the utilisation of existing in situ
materials to provide adequate and cost-effective pavement structures for the largest possible length.

3.2.1 New Constructions

TMR’s pavement design supplement (TMR 2018) uses terminology and definition that is broadly in alignment
with Austroads (2017) as follows:
Lightly bound granular materials are typically specified to have a UCS between 1.0 and
2.0 MPa at 28 days when used as LBC bases, and 1.0–2.0 MPa at seven days when used
as improved layers/subbase.
While this approach may result in a material that is more prone to fatigue and/or shrinkage
cracking than cement-modified materials (with maximum 28-day UCS of 1.0 MPa, as
defined in Austroads (2017)), it has a number of benefits, including:
• reduced moisture sensitivity
• higher strength and stiffness
• reduced permeability
• reduced erodibility
• reduced sensitivity to variations in grading and plasticity
• higher binder content is more readily and consistently achieved.
To alleviate some of the concerns relating to cracking when used in basecourses, TMR
typically adopts additional controls such as:
• SAM/SAMI seals over the LBC bases
• minimum 200 mm total thickness of LBC base
• minimum support conditions, as detailed in Section 8.2.8 of Austroads (2017).

The TMR specification MRTST10 (TMR 2019) for lightly bound base and subbase requires a 28-day UCS
minimum of 1.0 MPa and a maximum of 2.0 MPa, tested after moist curing but without soaking at
100% standard Proctor maximum dry density (MDD) and optimum moisture content (OMC). The
cementitious binder most commonly used is 75% cement/25% fly ash (complying with type GB cement).

Austroads 2020 | page 6


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Over the last 15 years TMR has placed over 100 lane-km of plant-mixed lightly bound base (PM-LB) and
carried out a significant amount of in situ stabilisation. Binder contents were generally in the range of
1.5–3.0%. Figure 3.1 is an example of a PM-LBC trial in Brisbane placed in 2011 using 1.5% type GB
(general blend) cement. The lightly bound base was 300 mm thick, placed in two 150 mm layers. Several
other projects are described in Section 5.

Figure 3.1: LBC trial on Frank Mallan Drive Lytton about 2.5 years after construction

Source: Lam and Bryant (2014).

In terms of the TMR structural design method, the lightly bound layer is modelled as unbound granular
material as follows (TMR 2018):
• sub-layered into five sublayers using Austroads (2017) sub-layering
• cross-anisotropic (with a degree of anisotropy of 2)
• Poisson’s ratio of 0.35
• modulus of top granular sublayer in accordance with Table 3.1.

Table 3.1: Presumptive values for elastic characterisation of LBC materials

MRTS10 material type(1) Presumptive vertical modulus of top sublayer (MPa)


Lightly bound base(2) Refer to Table 6.5 of Austroads (2017)
Lightly bound improved layer/subbase 210

1 Transport and Main Roads Specification MRST10 Plant-mixed lightly bound pavements (TMR 2019).
2 Lightly bound base layers are typically covered with either a thin asphalt or sprayed bituminous treatment, with the
lightly bound pavement material forming the main structural pavement layer. These pavements would typically be
specified in moderately trafficked pavement applications or where moisture control is required during construction or
throughout the service life of the pavement.

Source: TMR (2018).

Due to the sensitivity of the performance of lightly bound bases to underlying support conditions, TMR (2018)
provides the following minimum requirements:
For lightly bound base pavements, the base is typically supported on a subbase with a
thickness of at least 150 mm and which achieves a vertical design modulus of at least
150 MPa at the top of the subbase (determined using the procedures detailed in
Sections 8.2.2 and 8.2.3 of AGPT02). This may be achieved by increasing the thickness of
the subbase, and/or including additional select fill or unbound granular material beneath
the subbase.

Austroads 2020 | page 7


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

TMR (2018) provides the following advice concerning the risk of cracking:
When using lightly bound granular materials, the risks of both shrinkage and fatigue
cracking should be recognised and accepted, and associated maintenance interventions
over the life of the pavement should be anticipated.

3.2.2 Pavement Rehabilitations

TMR has used cement stabilisation of pavement materials as a rehabilitation treatment on an intermittent
basis for many years, including for the treatment of flood-affected roads. In Section 5, the performance of
LBC bases in Queensland is described in detail, particularly the performance of in situ stabilisation of rural
roads.

In the design of pavement rehabilitation treatments (TMR 2020), the definition of LBC and the thickness
design methods differ from those used in the design of new pavements. For rehabilitation projects, LBC
materials have a target 7-day UCS of 1.5 MPa within an allowable 7-day range of 1–2 MPa. The 28-day UCS
values could be 30–50% higher for a type GB cement. These criteria relate to specimens compacted to
100% standard Proctor maximum dry density (MDD) at standard optimum moisture content (OMC), then
moist cured without soaking.

Where LBC are used, the following TMR requirements apply to in situ stabilisation:
• The minimum thickness and strength of unbound granular pavement material specified in Table 3.2 shall
be provided under the modified layer.
• The thickness of the modified layer shall be between 200–300 mm.
• Must complete direct measurement of the characteristic modulus as per AGPT02.

Table 3.2: Minimum thickness and strength of unbound granular pavement materials required below the LBC
layer

Minimum thickness of unbound Minimum CBR of unbound Subgrade


granular material below the granular material below the design CBR
modified layer(1) (mm) modified layer (%)
(%)
300 ≥3–< 5
150 35 ≥5–< 15
0 ≥15

1 Where there the subgrade CBR is less than 3%, a capping layer is required as per TMR pavement design
supplement.
2 CBR is California Bearing Ratio.

Source: Adapted from TMR (2020).

Modified materials are considered to behave as unbound granular materials with improved stiffness. For the
purpose of mechanistic design, they are modelled with the following properties:
• cross-anisotropic (degree of anisotropy of 2)
• a Poisson’s ratio of 0.35
• sub-layered
• have a maximum potential modulus less than the characteristic modulus determined by direct
measurement as per AGPT02 and 600 MPa.

Austroads 2020 | page 8


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

3.3 Main Roads Western Australia

3.3.1 Rural Floodways

For many years Main Roads Western Australia (MRWA) has used in situ modified granular bases for rural
floodways – MRWA policy is to stabilise all floodway pavements (where bridges are not provided) and repair
of sprayed seal surfaced rural granular pavements. The binder used is 1.5% to 2% type LH binder
(35% cement/65% slag). The UCS requirement for this use is a 7-day value of 0.6–1.0 MPa on laboratory
samples stabilised with type GP cement, compacted to 100% modified Proctor MDD and OMC, and tested
after four hours soaking in water. Layer thicknesses are designed using the Austroads (2017) empirical
design chart for unbound granular materials. The modified granular base is normally surfaced with a cutback
prime and a Class 170 bitumen double/double sprayed seal. MRWA reports these pavements are generally
performing well.

When the granular base is crushed gravel or crushed rock achieving a 7-day UCS below 1.0 MPa is not
always possible at the minimum practical cement contents of 0.75%. In those circumstances, the stabilised
base is surfaced with a geotextile reinforced seal (GRS) to inhibit surface cracking.

In response to the project request for performance data, MRWA identified three floodways on the Great
Eastern Highway between chainages 239.3–240.7 km. All three floodways include a 200 mm thick
cement-modified laterite (PI = 6%) base overlying 300–400 mm thick granular subbase. These pavements
were constructed in 2006.

The three floodways (Figure 3.2) were inspected in December 2014 as part of this research project. After
eight years of trafficking, all the pavements had minimal rutting and good ride quality. At two of three sites
there was no cracking in the traffic lanes. At the other floodway, about 5–10% of the eastbound lane was
cracked with pumping of fines (Figure 3.3). The spacing between cracks was about 500 mm. There was no
cracking in the westbound lane. The difference in cracking between the east and westbound lanes could not
be explained.

Figure 3.2: Great Eastern Highway floodways

Site 1 239.33–239.42 km Site 2 240.59–240.70 km

Site 3 242.39–242.50 km

Austroads 2020 | page 9


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 3.3: Minor cracking at Site 2 observed 8 years after construction

3.3.2 Hydrated Cement-treated Crushed Rock Base

The other significant MRWA application of cement-treated bases is hydrated cement-treated crushed rock
base (HCTCRB). HCTCRB is crushed rock base plant-mixed with 2% cement then stockpiled and reworked
at the quarry before placement in roadbed. Thin asphalt-surfaced HCTCRB was extensively used for Perth
freeways between 2000 and 2009. In recent years, use has been limited due to HCTCRB cracking reflecting
through the thin asphalt surfacing.

HCTCRB currently has a specified maximum 28-day UCS of 1.3 MPa when tested at 100% modified Proctor
MDD and soaked for four hours before testing. HCTCRB test results indicate the 28-day UCS values at
100% standard Proctor MDD without soaking are less than 1 MPa. Despite this UCS value being within in
the Austroads (2017) range for modified materials, MRWA considers HCTCRB susceptible to fatigue
cracking and now requires consideration of this in thickness design.

Due to this risk of HCTCRB cracking, MRWA now requires the following surfacing for HCTCRB in freeway
applications:
• 30 mm open-graded asphalt
• 30 mm size 10 mm dense graded asphalt
• geotextile reinforced seal.

At the October 2014 meeting of APSWG (Section 2.3), it was agreed not to consider the field performance of
HCTCRB in this research project as this material is only used in Western Australia.

Austroads 2020 | page 10


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

3.4 Department of Transport Victoria


The most common reasons for stabilisation of pavements in Victoria are to improve the strength of materials
to give non-standard natural or blends of materials similar strength characteristics to standard quarry
products, and to reduce variability in performance. Due to rehabilitation funding constraints there has been
limited use of in situ stabilisation in DoT projects in recent years.

DoT’s in situ stabilisation specification (VicRoads 2008) allows a wide range of binders to be used in
cementitious modification, from fast-setting type GP cement to slow-setting blends such as slag/lime.
Rehabilitation practices vary between regions; for example, at the commencement of the research project:
• Western Region – a variety of pavement rehabilitation treatments are used including in situ stabilisation
for lower-trafficked rural roads. Slow-setting slag/lime or sometimes triple blends are used at a binder
content of about 2%.
• Northern Region – there is limited use of in situ cement stabilisation; when it is used, slow-setting
slag/lime binders are preferred. They have also treated granular bases with dry powdered polymer.
• Eastern Region – favours stabilising existing granular materials bases for use as a subbase followed by a
200 mm thick granular resheet. Commonly, type General Purpose (GP) cement is used; however, type
GB cement has also been used.

For cement stabilisation, binder contents range from 1.5–3.5% to meet the 7-day UCS requirements in
Table 3.3. Note that for type GB cement, the 7-day UCS of 1.5 MPa equates to a 28-day UCS value of about
2.0 MPa. This value relates to specimens compacted to 100% modified Proctor MDD rather than
100% standard Proctor MDD as used in the Austroads definition of modified materials. Assuming that
standard MDD is 5% lower than modified MDD and assuming a 9% reduction in UCS for a 1% decrease in
density (White 2006), the VicRoads 28-day UCS for type GB cement is about 1.4 MPa.

Table 3.3: Target UCS to be used to select the optimum content of cementitious binder

Cementitious binder
Target 7-day UCS at modified
content
compactive effort (MPa)
(% by mass)
Type of work
Slow(1) Medium(2) Rapid(3)
Min Max
Setting Setting Setting
Material modification 1.5(4) 3.5 1.0 1.5 2.0
Fully bound not
4.0 5.5 2.5 3.5
(deep-lift stabilisation) applicable
Bound subbase for deep strength asphalt
not
pavements where assigned design 3.0 2.5 3.5 5.0
applicable
modulus is not more than 500 MPa

1 Slow-setting: Slag/lime blends, alkali-activated slag and other supplementary cementitious blends.
2 Medium-setting: type GB cements, cement/slag blend (50% to 60% cement content),cement/fly ash blend (70% to
80% cement content), cement/slag/fly ash blend (55% to 65% cement content), slag/lime blends, alkali-activated slag
and other supplementary cementitious blends.
3 Fast-setting: type GP cement.
4 The Superintendent may agree to a cementitious binder content of less than 1.5% but at least 1% by mass if the UCS
requirement at a binder content of 1.5% is more than 50% above the target UCS. If less than 1.5% by mass of
cementitious binder is permitted, a minimum of two mixing runs shall be undertaken after spreading of cementitious
binder and the minimum density ratio specified in Clause 307.13 shall be increased by one percentage point.

Source: VicRoads (2008).

Austroads 2020 | page 11


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Since the specification values shown in Table 3.3 were selected, VicRoads tested the UCS of three
high-quality crushed rocks treated with 1% type GP cement (Figure 3.4). The testing arose from the need to
investigate the practice of adding 1% binder to plant-mixed crushed rock during the construction of unbound
granular pavement material as an aid to achieving dry back. The results indicated that the 28-day UCS
values of specimens compacted to 100% standard Proctor MDD and without soaking ranged from
2.6–3.2 MPa, with an average value of 2.9 MPa.

Figure 3.4: Unconfined compressive strength (UCS) of unsoaked samples after 28 days curing

Source: VicRoads (2011).

In addition, all three stabilised crushed rock subbase materials had UCS values exceeding the DoT target
(Table 3.3). For these high-quality granular materials it seems unlikely that type GP cement can be used to
manufacture a material meeting the proposed Austroads definition of LBC (Section 9.3).

DoT consider that in situ stabilised material should be of a reasonable quality; otherwise, it becomes overly
dependent on the performance of the binder or it may require a higher than desirable binder content, thus
increasing the incidence of cracking. Also, with higher binder contents there is a greater risk that the
specified density may not be achieved. A satisfactory grading and plasticity index also improve resistance to
deformation (rutting).

When using in situ stabilising granular bases, DoT anticipates that special surfacing treatments such as a
SAM seal are required to inhibit surface cracking.

In terms of structural thickness design for in situ stabilisation, DoT uses its unbound granular design chart
(based on Austroads (2017)) to select stabilisation depth and desirable material properties. No reduction is
made in the thickness of cover over subgrade for the increased modulus of the cement-treated base.

Due to the limited use of this type of rehabilitation treatment in recent years, DoT did not identify any projects
for a performance review in this project other than one possible project on Maroondah Highway, Merton. It
appears the Merton project was stabilised with a slow-setting binder, being a mixture of cement, slag and
lime, but it was not inspected as the exact location of the project could not be established.

Austroads 2020 | page 12


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

3.5 Transport for NSW


Transport for NSW (TfNSW) requirements for modified granular materials are specified in R71 Construction
of Unbound and Modified Pavement Course (Roads and Maritime Services (Roads and Maritime) 2018).
This specification includes the following requirements:
• nominated working time (determined in line with Roads and Maritime Test Method T147 (Roads and
Maritime 2012)) in excess of six hours
• UCS (at 28 days normal curing or 7 days accelerated curing) less than 1 MPa, when tested at
100% standard Proctor MDD and tested without soaking. This aligns with the Austroads maximum UCS
requirements for modified materials.

The modified material may be plant-mixed or in situ stabilised. LBC bases are not commonly used.

Deep-lift in situ stabilisation has been used over the last 20 years as a strengthening treatment for rural
highways. The high binder content associated with deep-lift stabilisation, typically 4–5%, makes the layer a
HBC base susceptible to cracking. Slow-setting binders such as lime, slag/lime and to a less extent
slag/lime/fly ash and slag/cement are used. These binders have the benefit of minimising shrinkage cracking.

Due to the limited use of LBC materials in NSW, TfNSW was unable to provide performance data for the
project.

3.6 Department of Planning, Transport and Infrastructure, South


Australia
The Department of Planning, Transport and Infrastructure’s (DPTI) main interest in modified materials and
LBC is in situ stabilisation of granular pavements in remote rural locations, particularly for improving the
performance of non-conforming materials. DPTI has utilised cement, lime/fly ash, lime, and some other
proprietary binders over the last 15 years, with variable levels of success. Shrinkage and fatigue cracking
have been the initial distress modes followed by pumping and rutting.

Due to the limited use of LBC materials, DPTI was unable to provide performance data for the project.

3.7 Department of State Growth Tasmania


Currently there is limited use of modified materials and LBC in Tasmania. Cement stabilisation using type GP
cement has been used in the past, but the extent of reflection cracking has been of concern.

Due to the limited use of LBC, the Department was unable to provide performance data for the project.

3.8 New Zealand Transport Agency


As discussed in Section 4.5, in-service performance data (Gray et al. 2011) was collated for
14 cement-stabilised bases in New Zealand.

Austroads 2020 | page 13


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

4. Literature Review

4.1 Introduction
To assist in fulfilling the project objectives previous research on granular materials cementitiously treated
with 1–2% cementitious binders was summarised.

Section 4.2 describes the distress modes of materials treated with cementitious binders. As shrinkage
cracking is common, Section 4.3 describes relevant research findings and current guidance on reducing its
effect on pavement performance. Section 4.4 summarises research that has been undertaken on how LCM
materials respond to loading.

4.2 Distress Modes of Cement-treated Materials

4.2.1 Distress Modes

Chakrabarti, Kodikara and Pardo (2001) reported a survey of Australian local government agencies on the
distress modes of chemical treatments of lightly trafficked pavements (see Figure 4.1). At that time type GP
cement and lime were the most commonly used binders.

Figure 4.1: Use of chemical binders

Source: Chakrabarti, Kodikara and Pardo (2001).

Cracking was the predominant distress type of chemically treated pavements (Figure 4.2). Shrinkage of the
treated material was the most reported reason for the cracking, followed by cracking due to expansive
subgrades and load-induced fatigue cracking (Figure 4.3).

Austroads 2020 | page 14


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 4.2: Distress modes of chemically-treated pavements

Source: Chakrabarti, Kodikara and Pardo (2001).

Figure 4.3: Causes of cracking

Source: Chakrabarti, Kodikara and Pardo (2001).

Over the 15 years since this survey was taken there has been an increase in the use of medium-to-slow
setting binders, improvements in stabilisation equipment and surfacing treatments. This may well have
altered the predominant distress modes; nevertheless, the review of Queensland cracking data
(Section 5.3.2) concluded longitudinal and transverse cracking were the dominant cracking types of LBC
bases with a very minor amount of traffic-induced crocodile cracking.

It should be noted that the binder contents in use were not reported by Chakrabarti, Kodikara and
Pardo (2001), but were most likely in the range 2–4%. By comparison, the objective of the Austroads project
is to investigate granular bases lightly bound with lower contents, commonly around 2% cementitious binder.

As mentioned in Section 2.3.2, the scope of this Austroads research project does not cover cracking of LBC
materials due to expansive subgrades. This subject is better addressed in a separate research project
across all pavement types.

4.2.2 Description of Cracking

Vorobieff, Walter and Anwar (2010) provide a detailed description of the nature of cracking of HBC bases
stabilised using slow-setting cementitious binders (5% binder) in New South Wales. The description was
based on periodic inspections of deep-lift in situ stabilised pavements over 15 years. The observed cracking
patterns of HBC bases provide a useful benchmark to compare and contrast with that for LBC materials
investigated in this research project.

Austroads 2020 | page 15


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Table 4.1 lists the typical crack patterns observed in pavements with HBC bases. The primary cracking is
transverse shrinkage cracking. Vorobieff, Walter and Anwar (2010) attribute the secondary longitudinal
cracking to shrinkage cracking. George (2002) provides an alternative explanation:
In fact, the performance of a cement-bound base is often hardly affected by primary
cracking, which is the occurrence of transverse cracking due to shrinkage and occasionally
thermal effects. Traffic causes further deterioration of primary cracks through shear
movement of the crack edges, in some cases resulting in longitudinal wheel-path cracks.
Transverse cracks widen, with possible reduction in interlocking characteristics of the crack
faces. Little et al. (1995) with calculations, showed that transverse cracks induced by
shrinkage and exacerbated by thermal contraction can increase the intensity of
load-induced flexural stresses by a factor of as much as 2.0, which explains how the
longitudinal cracks evolve with time. This phenomenon, termed secondary cracking, is the
chief cause of a significant reduction in the performance of a cement-bound base and
hence, eventually, to pavement distress. The primary transverse cracks, if reflected
through the wearing surface, may permit ingress of surface water and consequent
degradation of the lower layers by pumping and/or debonding.

Vorobieff, Walter and Anwar (2010) reported that with the development of a weaker subgrade through
moisture ingress, the formation of short sections of wheelpath block (ladder) cracking can result in the larger
block crack pattern (Figure 4.4). Over the 15 year observation period three of the four crack patterns shown
in Figure 4.4 were observed: crocodile fatigue cracking was not observed at any of the sites.

The key difference between the cracking characteristics of HBC base and that observed for Queensland LBC
bases (Section 5) was the absence of block cracking for LBC bases.

Table 4.1: Typical crack patterns observed at deep-lift stabilised sites

Crack patterns Description


Pattern A – Transverse Initial cracking that appears across the traffic lanes – usually within 12 months of
shrinkage cracking construction. These cracks become wider in the winter months and close – up in the
summer months with changes in pavements temperatures.
Not all cracks extend across the full width of the pavement – some cracks may only occur
between in situ stabilisation mixing runs.
Pattern B – Block A combination of transverse and longitudinal shrinkage cracking in large slabs. The
(shrinkage cracking) transverse shrinkage cracks occur at a typical spacing of 3 to 6 m, and the longitudinal
cracks commonly appear at the edges of mixing runs.
Due to the 2.4 m wide mixing runs, the spacing of transverse cracks appears to decrease
to about 2 to 3 m intervals in order for the ‘slabs’ to form with a length to width ratio less
than 2:1.
Pattern C – Wheelpath About 400 mm by 500 mm rectangular blocks appear in either wheelpath. Unless the
block (‘ladder’) cracking cracks are sealed distinct blocks will form and continue to form. The cracks on the
surface appear wider around the perimeter of the blocks.
It is commonly believed that these longitudinal cracks that form the edges of the blocks
are due to fatigue. However, it is more likely that the initial cracks are due to higher loads
applied to a ‘cantilever’ pavement structure and the material not sustaining the tensile
strains in the surface of the stabilised layer.
These blocks may be depressed if the subgrade is weak or wet and erosion of the
subgrade fines may appear on the surface of the road.

Source: Vorobieff, Walter and Anwar (2010).

Austroads 2020 | page 16


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 4.4: Illustrations of cracking pattern types

Transverse shrinkage pattern – Pattern A Block (shrinkage) cracking – Pattern B

Wheelpath block cracking – Pattern C Crocodile fatigue cracking – Pattern D

Source: Vorobieff, Walter and Anwar (2010).

4.3 Shrinkage Cracking

4.3.1 Introduction

Road agency practitioners advised that the shrinkage cracking of LBC bases is of concern as it can lead to
increase surfacing and routine maintenance costs and possibly affect service and structural lives albeit to a
lesser extent than for HBC bases.

Austroads (2017) does not provide any specific guidance on the shrinkage characteristics of LBC. As
described in Section 4.3.3, Austroads (2017) provides advice on measures to reduce shrinkage cracking of
HBC materials, which to some extent applies to LBC materials.

For HBC bases with cement contents of 3% or more shrinkage cracking due to drying and/or thermal
movement with time can reduce the pavement service life considerably by erosion and degradation of
material in the vicinity of these cracks. Shrinkage cracking also increases load-induced stresses and
therefore influences fatigue life.

Austroads 2020 | page 17


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

4.3.2 Mechanism of Shrinkage

Shrinkage of cement-treated material mainly consists of drying shrinkage and autogenous shrinkage. Drying
shrinkage is a result of water from the mixture evaporates to the atmosphere (Figure 4.5). Autogenous
shrinkage is not related to moisture loss to the environment, but it is caused by water consumption through
hydration reaction which is called self-desiccation (Dunlop, Moss & Dodd 1975; Zhang, Hou & Han 2012).

Shrinkage cracking occurs when the shrinking cement-treated layer is restrained in movement by the
underlying layer and/or are self-restrained (e.g. by strain gradient).

Shrinkage cracks are generally fine/narrow during initiation, but subsequently become wider. This is
attributed to continued drying shrinkage and thermal cycling.

Figure 4.5: Shrinkage mechanism of cement paste

Source: Yeo (2011).

4.3.3 Measures to Reduce Shrinkage Cracking of Cemented Materials

The cause of shrinkage cracking has been studied extensively for cemented materials. It has been found to
be related to the shrinkage of the fine aggregate fraction, the amount of water used during placement, and a
lack of adequate curing.

In relation to HBC (commonly ≥ 3% binder), Austroads (2017) states the following measures may be
considered to reduce shrinkage cracking:
• Minimise the total cementitious binder content – the lower the binder content, the lower the moisture
required, and the less the shrinkage. However, this renders the material susceptible to erosion when
subjected to moisture ingress under loading.
• Use slow-setting binders, which promote slightly less shrinkage than type GP cement. These binders are
also likely to require less moisture for compaction which also reduces shrinkage.
• Minimise the clay content of the material to be cemented by controlling the amount and plasticity of fines
in the aggregate. This can be achieved by limiting the fines content to less than 20% passing the 75 µm
sieve and the plasticity index to values not greater than 20.
• Treat the existing pavement materials which have an excess of plastic fines by:
– pre-treating with lime or lime and cement, followed by stabilisation with fly ash blend cement
– mixing in gravel or crushed rock with little or no fines, the amount of material varying with the plasticity
and fines content of the existing pavement compared to the desired levels and the proposed depth of
stabilisation
– applying both of the above treatments
– using the existing material as a subbase only or, alternatively, programming for an early overlay.

Austroads 2020 | page 18


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

• Place a bituminous curing coat as soon as possible after construction to inhibit rapid drying out of the
cemented layer and delay surfacing as long as possible so that cracking occurs before surface
placement.
• In addition, whilst the following two measures do not serve to minimise shrinkage cracking they do
ameliorate the influence of shrinkage cracking on overlying layers:
– in situations where the final seal is to be placed immediately following curing, apply a SAM (or a SAMI)
or geotextile seal to inhibit potential shrinkage cracking of the surfacing
– use an appropriate polymer modified binder asphalt surfacing in preference to conventional asphalt
(refer to Austroads 2009).

The benefits of these treatments are not reflected by the Austroads thickness design process because the
design model is not capable of predicting the onset and development of reflective cracking. Therefore, similar
pavement compositions and structures result regardless of the presence of these treatments. The benefits,
however, can be shown in terms of improved reliability of the design by providing a surfacing less prone to
the onset and development of reflective cracking. The use of SBS or crumb rubber modifiers has been
shown to provide a more elastic response and hence, provide a surface with a greater capacity to resist
reflection cracking. Further discussion on the selection of appropriate PMBs for this type of application can
be found in the Part 4F: Bituminous Binders of the Guide (Austroads 2017b).

Freeman and Little (1998) concluded that laboratory research showed that the following practical factors
affect the amount of base shrinkage:
• initial shrinkage is caused by loss of water due to drying of the base
• the soil type is an important variable: low clay content granular materials shrink less
than fine-grained soils
• a mixture compacted above optimum moisture will shrink more than the same mixture
compacted at optimum moisture content
• changes in stabiliser content, density and temperature have only a minor effect on the
amount of shrinkage compared to the effect of initial compaction moisture content
• the spacing and width of the cracks depend on the tensile strength of the stabilised
material, shrinkage properties (soil type), and the friction between the base and
subgrade/subbase.

Smith and Caltabiano (1987) reported the results of field trials of low-shrinkage cement-treated base (CTB
category 1 with a minimum 7-day UCS of 3 MPa). Based on early-life performance it was concluded that
low-shrinkage cement-treated base reduced but not eliminated shrinkage cracking. Interestingly, it was
observed that the shrinkage cracking of low-shrinkage cement-treated bases was finer for pavements on
high strength subgrades and those opening to traffic soon after construction. The authors concluded that
there is ‘no known process which will eliminate shrinkage cracking in cement-treated bases so the only
course is to limit this inherent weakness as much as possible’. Section 9.5describes minimum support to
LBC bases to inhibit the development of macro-cracking.

4.3.4 Inducing Early-life Micro-cracking to Reduce Shrinkage Cracking

Over the last 10 to 15 years, there has been considerable research in the USA on the effect of inducing
micro-cracking in cement-treated bases on the extent and severity of shrinkage cracking (Sebesta 2005; Wu
& Gaspard 2013).

Micro-cracking is commonly induced by passing a vibratory roller over a newly constructed cement-treated
base (Figure 4.6) to relieve the internal shrinkage stresses that develop during the initial cement hydration.
The desired result of this micro-cracking process is a network of fine, closely spaced cracks that are
uniformly distributed throughout the cement-treated base.

Austroads 2020 | page 19


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 4.6: Micro-cracking of soil-cement

Source: Scullion (2002).

Sebesta (2005) reported the findings of field trials in Texas from which it was concluded that roller-induced
micro-cracking of a 4% cement-treated granular material reduced the extent of shrinkage cracking by 50%.

No information has been reported on the influence of roller-induced micro-cracking on the fatigue cracking
characteristics of cemented materials.

4.4 Fatigue Cracking

4.4.1 Introduction

While Austroads (2017) provides a method to predict the fatigue life of HBC layers, there is currently no
method to design for the fatigue cracking of LBC layers.

Concerning the need to develop a method to predict the fatigue life of LBC, summarised below are the
findings of previous research related to:
• laboratory fatigue testing
• fatigue in the roadbed, including accelerated loading and in-service performance.

4.4.2 Laboratory Fatigue Characterisation

There is a very limited amount of research that has been undertaken on the fatigue of granular materials
treated with around 2% cementitious binder.

Arnold, Morkel and van der Westhuizen (2012) reported the findings of flexural beam fatigue testing of a size
40 alluvial gravel with type GP cement at contents of 1%, 2% and 4%. This testing was undertaken to
complement the results of New Zealand accelerated loading described in Section 4.4.3. Test beams were
subjected to accelerated-curing prior to fatigue testing as follows:
• compacted beams were wet cured in the compaction mould for about 20 hours
• the beams were then removed from the mould, placed in sealed plastic bags and then cured in an oven at
a temperature of 40 °C for 72 hours
• lastly, the beams were removed from the bags and dried for 24 hours in the laboratory, presumably at a
temperature of about 21 °C.

Austroads 2020 | page 20


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

This is a different curing process from the ≥ 90 days moist curing at a temperature of 21 °C used in
Austroads research (Austroads 2014). The use of accelerated curing at 40 °C may have led to more
micro-cracking of the test beams than the Austroads long-term moist curing process. This micro-cracking
may be the reason the test beam moduli were a factor of 10 lower than the Austroads (2014) measured
values. The tolerable strains for a given fatigue life were also significantly higher. In support of their
accelerated-curing process, Arnold, Morkel and van der Westhuizen (2012) state:
a case can therefore be presented for the carefully controlled temperature-induced
cracking in laboratory samples to reflect (field) performance better.

The results of the beam fatigue testing are given in Figure 4.7.

Figure 4.7: Beam flexural fatigue test results

Source: Arnold, Morkel and van der Westhuizen (2012).

Using the laboratory test results, Arnold, Morkel and van der Westhuizen (2012) calculated the tensile strains
in accelerated loading of pavements which included alluvial greywacke gravel stabilised using cement
contents of 1%, 2% and 4%. Figure 4.8 illustrates the pavement structure used in the cemented material
strain predictions.

Figure 4.9 shows the predicted fatigue lives using a design modulus of 2000 MPa for all three cemented
materials rather than using the measured modulus. Note the very low predicted fatigue lives of the 1% and
2% mixes, consistent with the findings of this research project (Section 9.2).

Austroads 2020 | page 21


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 4.8: Typical pavement cross-section of pavements subject to accelerated loading

Note: that the thickness of cemented-stabilised material should read 200 mm rather than 300 mm.
In addition the pavement included a 40 mm dense-graded asphalt surfacing.

Source: Arnold, Morkel and van der Westhuizen (2012).

Figure 4.9: Pavement fatigue lives predicted for the CAPTIF pavements using 2000 MPa modulus

Source: Arnold, Morkel and van der Westhuizen (2012).

In relation to apply the research findings, Arnold et al. (2012) concluded:

Austroads 2020 | page 22


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Materials in this research may have not been fully bound and future research should
investigate higher cement contents. The 1% and 2% cement-stabilised aggregates did
result in solid and ‘bound’ looking beams. However, when tested they were low in tensile
strength and the numbers generated for the design tensile fatigue criteria resulted in short
fatigue lives for typical pavement designs. Therefore, 1% and 2% cement-stabilised
aggregates in this study would always be considered as unbound as their phase 1 bound
behaviour is very short. Thus a boundary in terms of tensile strength or cement content is
not needed to define when a material is unbound, modified and bound because the tensile
strength measured in the laboratory, combined with use in design, will quickly determine if
any phase 1 fatigue life is possible.

4.4.3 Fatigue under Accelerated Loading

Alabaster et al. (2013) describe an extensive New Zealand research project which included the following
objectives:
• Objective1: Determine the benefits of using cement and/or lime-modified aggregates in terms of
increased performance (rut resistance) and incorporate this in a design methodology, filling a gap
identified by Austroads.
• Objective 3: Understand the continuum from unbound (no binder), modified (small amounts of binder) to
bound (high amounts of binder) behaviour.

The methodology to complete the objectives involved a combination of accelerated pavement tests at the
Canterbury Accelerate Pavement Testing Indoor Facility (CAPTIF) in 2007–08 and 2008–09, and a limited
review of the field performance of bound stabilised pavements.

Accelerated loading was undertaken on six material types listed in Table 4.2. All stabilised layers were
200 mm thick with a nominal 40 mm thickness of dense graded asphalt surfacing. Unfortunately, the testing
of the 2% cement mix was affected by accidental flooding when a water pipe burst.

Table 4.2: Pavement materials subject to accelerated loading

Section Material type


A Size 40 greywacke gravel blended with clay, PI = 6%
B Greywacke gravel with 1% type GP cement
C Greywacke gravel with 2% type GP cement
D Greywacke gravel with 2% lime
E Greywacke gravel with 4% type GP cement
F Greywacke gravel with 4% type GP cement – pre-cracked

Source: Alabaster et al. (2013).

About 1.5 million cycles of 60 kN dual-tyre loading were applied which equates to about 7.6 x 106 ESA of
loading. One interesting finding was that at the end of loading no surface cracking, either fatigue or
shrinkage, was observed. This absence of fatigue cracking in the thin asphalt surfacing is consistent with the
very limited surface cracking observed in an Australian accelerated loading trial of crushed rocks stabilised
with 3% and 4% cement (Austroads 2008). However, in this Australian research cracking was observed
when the pavement was excavated. The absence of surface cracking in both these accelerated loading trials
may be due to bituminous surfacings being newly placed and hence more ductile than for in-service
pavements.

Austroads 2020 | page 23


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

As mentioned in Section 4.4.2, laboratory beam testing indicated that the 1, 2 and 4% mixes were all
susceptible to fatigue cracking. However, surface cracking was not observed under accelerated loading.
Nevertheless, the measured vertical subgrade strains (Figure 4.10) and the moduli back-calculated
(Figure 4.11) from Falling Weight Deflectometer (FWD) surface deflections both indicated the cemented
materials decreased in modulus under accelerated loading. (Note the performance of the 2% mixes was
affected by accidental flooding.) Of interest to this research project were the relative moduli of the 1% mix
and 4% mix at the end of loading. The 1% mix was about only 20% lower in modulus than the 4% mix. The
moduli of both these mixes after trafficking were more than twice that for the unbound granular base. This
information was used in developing the proposed Austroads design method (Section 9).

Note that the FWD testing before accelerated loading was taken a month after construction. Consequently,
the cemented materials may not have cured to a long-term modulus value at the commencement of loading.
This may have been a factor in the 4% mix mean back-calculated modulus (4 550 MPa) being low. Also, the
test pavement was compacted to a relatively low field density.

Figure 4.10: Change in measured vertical subgrade strains with loading cycles

Source: Alabaster et al. (2013).

Austroads 2020 | page 24


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 4.11: Change in back-calculated moduli with loading cycles

Source: Alabaster et al. (2013).

According to Alabaster et al. (2013) the key findings were:


• Objective 1: ‘From the results of accelerated loading it was shown that modifying the tested aggregates
with 1% cement could reduce rutting and improve the rutting life of the pavement by 200–300% compared
with the unbound pavement. However, stiffness loss occurred during this testing. This report presents a
design methodology that can be used with initial laboratory data to estimate the initial improvement in
rutting performance. Further research is needed into the implications of this stiffness loss in cement-only
materials. This report presents a wider approach that includes other recent New Zealand Transport
Agency (NZTA) research to address the stiffness loss’.
• Objective 3: ‘The project did not result in a better understanding of the continuum from modified to bound
behaviour; however, it provided some useful findings. The research suggests that bound behaviour
clearly occurs at 3–4% cement contents. The accelerated pavement loading (CAPTIF) testing showed at
4% cement contents there was very little rutting, but significant stiffness loss. At 3% cement contents in
the field, fatigue failures were observed. At CAPTIF, materials with 4% cement showed significant losses
of stiffness and the stiffness tended to the value of 1% cement’.

Furthermore Alabaster et al. (2013) concluded:


It would be a prudent limit for design to start considering bound behaviour at 2% cement
content, which from the CAPTIF test would be a vibratory-hammer-prepared soaked ITS
over a limit of 600 kPa limited (when mixed and tested in the lab). This would form an
upper limit for using the procedures. Above this limit there is a risk of cracking, leading to
water entering the pavement, and potentially rapid failure and difficult repairs. Below this
limit there is a risk of the stiffness reducing and the performance not being as good as
estimated (this is considered in the procedures). These values are considerably higher
than those proposed by Austroads but are, in part, a function of the way New Zealand
engineers ask for the sample to be prepared.

Austroads 2020 | page 25


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

4.5 Investigation of New Zealand Stabilised Pavements


Gray et al. (2011) collated in-service performance data for 14 cement-stabilised bases in New Zealand. The
cement content was generally 3%, but 2% cement was used at Site 12 and Site 16. In this respect, it is
unlikely that most of these sites would meet the proposed Austroads definition of LBC bases.

Whilst Site 16 was in good condition after four years of trafficking, Site 12 cracked and rutted after 1 year of
trafficking (Figure 4.12).

Figure 4.12: Site 12 – Hirini Street Gisborne

Source: Gray et al. (2011).

The investigation included periodic deflection testing at two sites stabilised with 3% cement. From these
deflections, the isotropic moduli of the cement-treated base were back-calculated (Figure 4.13). The solid
blue line is the 50 percentile modulus value which, after the first two years, gradually reduced to an isotropic
modulus of about 1000 MPa, presumably due to fatigue damage. The dashed blue line is the 10 percentile
modulus values corresponding to the weaker areas of pavement that most likely cracked and rutted first. The
isotropic modulus of these weakly cemented materials was about 500 MPa. According to Austroads (2019a),
an isotropic modulus of 500 MPa is roughly equivalent to the following cross-anisotropic characterisation: a
vertical modulus of 550 MPa and a horizontal modulus of 275 MPa. This information was considered in the
development of the proposed elastic characterisation (Section 9.4).

Austroads 2020 | page 26


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 4.13: Variation in back-calculated moduli of 3% cemented materials

Source: Gray et al. (2011).

The authors concluded:


When using cement additive contents of between 1 and 3% by dry mass in stabilised
basecourse materials (with or without seal inclusion) the resulting cement-stabilised
pavement layer should be modelled as a lightly bound material.

4.6 Summary
Road agencies have identified the potential to increase the use of granular bases treated with
1–2% cementitious binders to improve their rut resistance when used in thin bituminous surfaced pavements
and to provide higher modulus layers under thin asphalt surfacings.

This review of Australian and New Zealand research has concluded that:
• LBC materials have the potential to crack due to shrinkage, albeit to a lesser extent than HBC materials.
Much of the extensive research that has been undertaken on measures to minimise the effects of
shrinkage of HBC materials with binder contents of 3% or more is considered applicable to LBC
materials.
• LBC materials have very low fatigue lives.
• HBC bases are susceptible to load-induced block-cracking that forms following transverse shrinkage
cracking. Based on the extensive use of LBC bases in Queensland (Section 5), it is unlikely that
appropriately designed and constructed LBC bases develop block (ladder) cracking and crocodile
cracking. This is the key difference in cracking characteristics that has led to the use of LBC with thin
bituminous surfacings.

Austroads 2020 | page 27


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

5. Performance Review of Selected Queensland


Pavements

5.1 Introduction
An important element of developing a mechanistic-empirical structural design method was an understanding
of in-service performance of pavements which include LBC. The propensity of LBC bases to crack was of
particular interest.

As described in Section 3.2, TMR has developed design procedures and specifications for the use of LBC in
new flexible pavements (TMR 2018) and also for use in pavement rehabilitation (TMR 2020). TMR have
constructed more LBC bases than the other Australian state road agencies combined. Accordingly, it was
decided to review the performance of Queensland pavements as follows:
• As described in Section 5.2, 16 sites were identified on the Queensland road network where granular
base materials had been treated with around 2% cementitious binder by both in situ stabilisation and
plant-mixed materials. Key findings resulted from site inspections, data collation and analysis.
• Queensland network condition data was analysed to assess the extent and type of cracking of LBC
pavements and to identify factors related to cracking.(Section 5.3).
• In addition, coring of the LBC layer at two of the 16 sites described in Section 5.2 provided insight as to
whether the LBC is a bound material with the potential for macro-cracking or if it is in a micro-cracked
state with reduced potential for surface cracking (Section 5.4).

5.2 Queensland Field Site Investigation

5.2.1 Methodology

Investigation of data records of over 8500 km of the Queensland network identified 16 projects that met the
criteria aimed at selecting structures where any evident surface cracking could reasonably be expected to be
attributed to load. The criteria included:
• cement-stabilised granular
• base thicknesses in the range 150–250 mm
• bituminous surfacing thickness 50 mm or less
• pavements without bound subbases or stabilised subgrades
• unlikely to have been constructed on highly expansive subgrades
• more than 200 heavy vehicles per day
• bases constructed at least three years ago.

Using these criteria, 16 projects were chosen, representing 90 km, a little over 1%, of the current lightly
cemented road network. The locations of the sites are shown in Figure 5.1.

Austroads 2020 | page 28


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Field inspection sites were:


• Site 1: Darling Downs District Cunningham Highway (Ipswich – Warwick): Job 110/17B/302
• Site 2 and 2A: Darling Downs District Cunningham Highway (Inglewood – Goondiwindi): Job 122/17D/807
• Site 4: Far North Queensland District Palmerston Highway (Innisfail – Ravenshoe): Job 264/21A/56Z
• Site 5: Far North Queensland District Malanda – Atherton Road: Job 119/645/802 & 803
• Site 6: North West District Flinders Highway (Hughenden – Richmond): Job 55/14C/304
• Site 7 & 7A: North West District Flinders Highway (Hughenden – Richmond): Job 257/14C/67H
• Site 8: Mackay/Whitsunday District Bruce Highway (Bowen – Ayr): Job 125/10K/59
• Site 9: Mackay/Whitsunday District Bruce Highway (Bowen – Ayr (10K)): Job 61/10N/508
• Site 10: Northern Queensland District North Townsville Road (Woolcock St: Townsville Port Road):
Job 150/10M/17
• Site 11: Northern Queensland District Douglas – Garbutt Road (Duckworth St): Job 150/10L/38
• Site 12: Northern Queensland District Bruce Highway (Townsville – Ingham): Job 117/10M/819
• Site 13: Northern Queensland District Bruce Highway (Townsville – Ingham): Job 61/10M/32
• Site 14 & 14A: Fitzroy District Bundaberg – Miriam Vale Road: Job 83/179/23
• Site 15: Fitzroy District Western Yeppoon – Emu Park Road: Job 258/197/201
• Site 16: Fitzroy District Dawson Highway (Gladstone – Biloela): Job 27/46A/304
• Site 17 & 17A: Fitzroy District Dawson Highway (Biloela – Banana): Job 8/46B/303.

(Site 3: Far North Queensland District Bruce Highway (Ingham – Innisfail): Job 216/10N/66Z. This site is not
reported as the treatment was found on inspection to be foamed bitumen stabilisation.)

Figure 5.1: Locations of field sites

Source: http://ontheworldmap.com/australia/state/queensland/.

Austroads 2020 | page 29


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Before field visitation, asset management data, construction drawings and UCS records (if available) were
collated for each project prior to site visits.

A visual inspection focussing on load-induced cracking defects but also the recording of other major defects
and patch repairs were undertaken along with local TMR district staff.

Appendix A contains notes and photographs taken during the inspections which were conducted between
June and September 2016.

5.2.2 Field Site Investigation: Preliminary Findings

From the field inspections conducted it can be concluded that the plant mixed lightly cemented bases have
performed well. As such materials are placed as part of new construction, it is reasonable to assume that the
underlying support to the lightly-cemented base to be expected in new pavement design is a significant
factor in the good performance of these projects. However, in some cases the thin LBC bases were
constructed on cemented subbases and transverse cracking was observed on the surface; this was most
likely due to shrinkage of the cemented subbases.

The in situ stabilised bases showed a much wider range of performance, but generally the extent of block
cracking and crocodile cracking was minimal. All LBC bases that were well supported demonstrated good
performance, as did some pavements where the underlying support could be considered as inadequate from
a structural design point of view.

However, some in situ stabilised pavements exhibited very poor performance with the formation of major
block cracking evident. Additionally, some sections were found to have been extensively patched, indicating
that significant distress had been evident and the pavement had needed treatment to restore serviceability.

Despite extensive attempts to source additional information and records, including the recollections of
regional staff, it is difficult to prove a definitive cause for some of the major distress.

It has been postulated that the application of excessive binder has resulted in a continuous length of one
project exhibiting extreme block cracking in all wheelpaths. Adjacent lengths of road, which were part of the
same rehabilitation project did not demonstrate this cracking and are performing well. Documentation could
not be found to demonstrate that an increase in binder application rate, or the use of binder from a different
source, had been the cause of the poor performance; however, it was the most likely cause of the
block-cracking.

Nevertheless, long lengths of LBC showed little or no signs of block cracking, despite the considerable
passage of time and loading since construction. It is apparent that, with improved design and construction
processes, LBC bases can more consistently perform well.

5.3 Review of Network Pavement Condition Data

5.3.1 Introduction

Following the field inspections, the network pavement condition data was analysed to clarify the extent and
type of cracking at the sites inspected and the factors associated with this cracking.

Austroads 2020 | page 30


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

5.3.2 Cracking Type

From Figure 5.2 it can be seen that the predominant cracking type was longitudinal followed by transverse.
The transverse cracking is likely to be due to shrinkage of the LBC bases, whilst the cause of longitudinal
cracking could be:
• shrinkage cracking (Section 4.2.2)
• expansive subgrade soils
• the beginnings of load-induced fatigue cracking
• construction joints (e.g. pavement widenings).

Based on the inspections (Appendix A), the most likely cause of longitudinal cracking at the sites inspected
was the presence of expansive subgrade soils.

The extent of crocodile cracking was minimal, suggesting the pavements have generally been designed and
constructed as intended.

Figure 5.2: Crack behaviour in the different wheelpath locations

5.3.3 Factors Associated with Cracking

The percentage of the site length cracked is compared with the maximum deflection (D0) and the curvature
(D0-D200) measured using the traffic speed deflectometer (TSD) in Figure 5.3 and Figure 5.4. It was
concluded that the extent of longitudinal cracking was significantly related to surface deflection. This may be
because pavements on highly expansive clay subgrades generally have higher deflections than
higher-strength low-expansive subgrades.

Austroads 2020 | page 31


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

The extent of transverse cracking also tended to increase with deflection, despite the cause of this cracking
type being most likely shrinkage rather than load-induced. A possible reason is that the higher the deflection,
the greater the load-induced movement at a transverse crack with associated gradual loss of load transfer
across the crack and hence the development of wider, more detectable cracking.

The extent of crocodile cracking was so low that no conclusions could be drawn about its dependence on
deflection.

Figure 5.3: Crack percentage as a function of maximum deflection D0

Figure 5.4: Crack percentage as a function of curvature (D0-D200)

Austroads 2020 | page 32


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

A critical issue to investigate was whether the extent of cracking was related to the LBC base thickness.
Accordingly:
• Figure 5.5 shows the variation in the percentage of total length with any cracking with the LBC base
thickness at each site sorted according to subbase thickness.
• Figure 5.6 shows the variation in the percentage of length with longitudinal cracking with the LBC base
thickness at each site sorted according to subbase thickness.
• Figure 5.7 shows the variation in the percentage of length with transverse cracking with the LBC base
thickness at each site sorted according to subbase thickness.
• Figure 5.8 shows the variation in the percentage of length with crocodile cracking with the LBC base
thickness at each site sorted according to subbase thickness.

In each plot the larger the data point plotted the greater the cumulative traffic loading at the time the
pavement was inspected.

Again the extent of crocodile cracking was low which hindered conclusions being drawn relating the
pavement structure to this type of load-induced damage.

LBC thickness does not seem to influence cracking extents. This may be because the procedures used by
TMR to vary the LBC base thickness with traffic loading, based on engineering judgment, are reasonable.

In terms of transverse and longitudinal cracking, high percentages of cracking were observed for pavements
with lower subbase thicknesses (Site 2 and Site14). The subbase seems to be the predominant factor
influencing cracking extent. This can be seen at Site 14, which had the least amount of traffic loading but
much higher cracking extent. Where the subbase depth was above 225 mm, all sites showed a total average
cracking extent of less than 3% (Site 2A, 14A, 16, 17 and 17A).

Since the sites inspected were constructed, the TMR (2018) design procedures have been improved to
provide sufficient granular subbase thickness such that the top sublayer of granular has a minimum design
modulus of 150 MPa.

Figure 5.5: Total crack length versus LBC base thickness

Austroads 2020 | page 33


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 5.6: Longitudinal crack length versus LBC base thickness

Figure 5.7: Transverse crack length versus LBC base thickness

Austroads 2020 | page 34


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 5.8: Crocodile crack length versus LBC base thickness

5.4 Detailed Pavement Investigations

5.4.1 Introduction

As described in Section 5.3, there was only a minor load-induced crocodile and block cracking of the LBC
bases on the Queensland road network. This is likely to differ from the performance of cemented material
bases which have higher strengths than LBC; the cracking of these bases was not addressed in this report.

To investigate this difference in behaviour of the LBC bases, coring was conducted at two of the 16 sites
inspected (Section 5.2.1) to provide insight about the cracking characteristics of LBC bases.

5.4.2 Site 2/2A Cunningham Highway (Inglewood – Goondiwindi)

Pavement Composition

This 1.6 km length of the Cunningham Highway was in situ stabilised in 2007 to a depth of 250 mm using
2% type GB cement. The host material was river gravel.

Note that between chainages 85.4–86.6 km (Site 2) the granular subbase under the stabilised layer was only
50 mm thick, whereas between 86.6–87.0 km (Site 2A) the thickness of granular was 325 mm (Table 5.1).
The subgrade CBR was similar in both sections.

The surfacing was a size 10 mm primerseal followed by a size 14 mm polymer modified binder seal.

Austroads 2020 | page 35


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Table 5.1: Composition of Sites 2 and 2A

Chainages 85.4 to Chainage 86.6 to 87.0 km


86.6 km 2A
Site 2

UCS testing after 7 days curing was undertaken by TMR before construction to determine the binder content
(Figure 5.9). From this testing it was decided to adopt 2% cement as this value produced UCS values within
the specified 7-day UCS range of 1.0 to 1.5 MPa.

Figure 5.9: 7-day UCS testing used to select the binder content
3.0

2.5

2.0

UCS
1.5
(MPa)

1.0

0.5

0.0
1.5 2.0 2.5 3.0 3.5 4.0 4.5
Content of type GB cement (%)

Visual Inspection

This section of Cunningham Highway was a visual inspection in 2016, which was about nine years after
construction and opening to traffic.

Overall 4–5% of the total 1.6 km project was distressed exhibiting either crocodile cracking (Figure 5.10),
transverse cracking, or patching. The distress was more apparent in Site 2 which had only a 50 mm
thickness of granular subbase, whereas Site 2A was generally in good condition.

There was some longitudinal cracking on the shoulder edges, indicating the likelihood of expansive
subgrades.

Austroads 2020 | page 36


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 5.10: Examples of cracking at Site 2

Coring

The LBC base was cored (Figure 5.11, Figure 5.12) from 12 locations along the site (see Table 5.2 and
Table 5.3) and the rut depth measured.

Figure 5.11: Coring rig set-up Figure 5.12: Core extracted

Dry coring rather than wet coring so that the UCS results were not influenced by water used during coring. A
total of 12 cores were attempted, however only four cores were successfully extracted (Figure 5.13,
Figure 5.14). At the other sites, cores could not be retrieved (Figure 5.15, Figure 5.16) due to their low
strength (see notes in Table 5.2, Table 5.3). It could not be established whether or not the LBC base was
broken down due to trafficking or as a result of coring. All the core holes appeared solid and intact, which
seems to indicate an absence of macro-cracking.

Austroads 2020 | page 37


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Table 5.2: Summary of Coring Investigation at Site 2

Location Condition Cores Rutting Notes


Chainage Fair 1 core, 2 mm of rutting in Disintegration of the LBC full depth to subgrade.
86.550 0 successful IWP and OWP Depth cored 270 mm. Observed unravelling of
1.0 m RHS LHS material layer within core hole walls and very
FL dry in condition. Significant amount of small
fines produced by coring.
Chainage Fair 1 core, 2 mm of rutting in Noted complete disintegration of this core
85.550 0 successful IWP and OWP sample during coring procedures full depth to
0.5 m LHS LHS subgrade. Depth cored 270 mm. Observed
FL unravelling of material layer within core hole
walls and very dry in condition. Significant
amount of small fines seen after coring.
Chainage Fair 1 core 20 mm of rutting Extracted core cracked total depth of layer
85.590 successful, but in OWP LHS, no 220 mm. Noted moisture within layer. Core
1.0 m RHS was cracked other rutting broke off from full depth of layer of 270 mm from
FL an unable to existing road surface.
be used for
UCS
Chainage Fair 1 core for UCS 20 mm of rutting Observed old crack and top layer free floating
85.590 in OWP LHS, no (25 mm). Observed moisture within break. Small
1.5 m RHS other rutting to no voids. Good condition. Noted break of
FL layer at 190 mm. Full depth cored 270 mm.
Subgrade observed as dry.
Chainage Fair Core 20 mm of rutting Core sample disintegrated during coring
85.590 disintegrated in OWP LHS, no procedures.
0.5 m LHS other rutting
FL
Chainage Fair 1 core, 4 mm of rutting in Disintegration of this core full depth to
86.310 0 successful OWP RHS and 8 subgrade. Depth cored 290 mm. River gravel
1.0 m RHS mm of rutting in blend with a significant amount > 19 mm in size.
FL IWP RHS Very dry in condition. Moderate amount of small
fines was observed after coring.
Chainage Fair 1 core, 4 mm of rutting in Disintegration of this core full depth to
86.310 0 successful OWP RHS and 8 subgrade. Depth cored 200 mm. River gravel
2.0 m RHS mm of rutting in blend with significant amount > 19 mm in size.
FL IWP RHS Very dry in condition. Moderate amount of small
fines seen after coring.
Chainage Fair 1 core, 4 mm of rutting in Disintegration of this core full depth to
86.310 0 successful OWP RHS and 8 subgrade. Depth cored 170 mm. River gravel
3.0 m RHS mm of rutting in blend with a significant amount > 19 mm in size.
FL IWP RHS Very dry in condition. Moderate amount of small
fines seen after coring.

Note: OWP = outer wheelpath, IWP = inner wheelpath, RHS= right hand side, LHS= left hand side.

Austroads 2020 | page 38


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Table 5.3: Summary of Coring Investigation at Site 2A

Location Condition Cores Rutting Notes


Chainage Fair 1 core 2 mm of rutting in Seal layer and a thin delaminated layer of LBC
85.610 successful, 1 IWP RHS. No (25 mm) separated during coring procedures.
1.5 m RHS for UCS other rutting. Smooth surface of floating layer between
FL observed interfaces. No moisture within the
layer. Full depth of core hole to subgrade layer
255 mm. Subgrade observed to be dry.
Chainage Fair 1 core 2 mm of rutting in Seal layer and a thin delaminated layer of LBC
85.610 successful, 1 IWP RHS. No (25 mm) separated during coring. Noted 65 mm
1.0 m RHS for UCS other rutting. possible floating layer between interfaces
FL disintegrated during the coring procedure. No
moisture within layer. Full depth of core hole to
subgrade, thickness 250 mm.
Chainage Good 1 core, 5 mm of rutting in Disintegration of this core full depth to
86.800 0 successful OWP RHS and subgrade. Delamination of LBC layer with free
1.0 m RHS 2 mm of rutting in floating top surface break. Break observed with
FL IWP RHS smooth interfaces at joins. Depth cored
290 mm. LBC 250 mm in thickness. River
gravel blend. Very dry in condition. Moderate
amount of small fines seen after coring.
Chainage Good 1 core, 5 mm of rutting in Disintegration of this core full depth to
86.800 0 successful OWP RHS and subgrade. Delamination of LBC layer with free
0.5 m LHS FL 2 mm of rutting in floating top surface break. Break observed with
IWP RHS smooth interfaces at joins. Noted significant
amount of river gravel stone above +19 mm.
Depth cored 270 mm. LBC 250 mm in
thickness. Very dry in condition. Moderate
amount of small fines seen after coring.

Note: OWP = outer wheelpath, IWP = inner wheelpath, RHS= right hand side, LHS= left hand side.

Figure 5.13: Successfully extracted core

Austroads 2020 | page 39


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 5.14: Example core hole

Figure 5.15: Example of disintegrated of the LBC base observed after coring

Austroads 2020 | page 40


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 5.16: Example of LBC base layer not able to be extracted in tact

Unconfined Compressive Strength

Coring was carried out at a total of 12 locations in Sites 2 and 2A, however there were only three sound
cores were suitable for UCS testing (Table 5.4). The UCS of these sound field cores was around 2 MPa. It is
important to note here that Site 2 was constructed in 2007 and so was 10 years old when the investigation
was carried out. It would be expected that the in situ UCS would have increased in situ since constructed
due to curing offset possibly by traffic-induced fatigue damage.

Table 5.4: Summary of UCS data of cores

Sample ID Chainage Offset Depth below Sample Sample UCS


surface diameter height (MPa)
(mm) (mm) (mm)
P069/C1A 85590 1.5 m R FL 50–190 94.2 150 1.8
P069/C2A 85610 1.5 m R FL 85–235 144.2 164 2.0
P069/C2B 85610 1.0 m R FL 90–250 144.8 160 2.1

An important finding was that the sites at which cores were able to be extracted were located in areas with
greater cracking distress, whereas for two locations in Site 2A in good condition LBC core could not be
retrieved. For the sites in good condition, the LBC base may have micro-cracked to such an extent that cores
cannot be retrieved. Perhaps due to the fineness of the micro-cracking it was not detectable when the walls
of the core holes were inspected (Figure 5.14).

Austroads 2020 | page 41


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

5.4.3 Site 16 Dawson Highway (Gladstone – Biloela)

Pavement Composition

This 5.3 km length of the Dawson Highway (chainages 50.8 km and 56.1 km) was in situ stabilised in 2007 to
a depth of 150 mm using 2.8% type GB cement. Prior to stabilisation 100 mm thick type 2.5 granular bases
was placed on the existing pavement. After stabilisation, the underlying granular subbase was about 280 mm
thick (Table 5.5).

The surfacing was a sprayed bituminous seal.

Table 5.5: Composition of Sites 16

Site 16

No details of UCS testing before construction could be obtained. However, the ‘as constructed’ drawings
recorded the average cement content used was 2.8% and the UCS values measured during construction
were between 0.9 MPa and 2.3 MPa. The mean UCS was 1.5 MPa, with four of 12 results between
2.2–2.3 MPa. Although not stated in the drawings, it is likely these results relate to the UCS values after
7 days curing.

Visual Inspection

When inspected in 2016, the pavement had been in-service for about nine years old at which time the
accumulated traffic loading was about 106 ESA.

Three forms of distress were observed in isolated areas (Figure 5.17), as follows:
• 97 lineal metres (113 m2) of crocodile cracking
• 42 m2 of rut/shove failure
• 381 m2 of patches.

In total, 1.3% of the total wheelpath length displayed one of these three forms of distress.

Austroads 2020 | page 42


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 5.17: Dawson Highway inspection in 2016

Coring

Cores were taken at eight locations (Table 5.6). None of the core could be extracted intact, in all cases the
materials disintegrated due to traffic loading or due to the coring process.

Figure 5.18 shows the coring hole walls and so some very large aggregate sizes. In most cases the cored
LBC broke down into a granular material (Figure 5.19 and Figure 5.20). There was one instance documented
where some layers of intact material could be extracted (Figure 5.21).

Consistent with the Cunningham Highway findings (Section 5.4), when the pavement was cored at the sites
in sound condition, intact cores of LBC could not be retrieved. However, for the Dawson Highway site the
large aggregate sizes in the LBC may also have led to the poor coring success rate.

In addition, like Cunningham Highway micro-cracking was not apparent on the wall of the core holes, but
again this maybe because the cracking is too fine to be observed.

The LBC base may have micro-cracked consistent with the desired characteristics of LBC. Alternatively, the
intact cores could not be retrieved.

Austroads 2020 | page 43


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Table 5.6: Summary of Coring Investigation at Site 16

Location Condition Cores Rutting Notes


52.20 Eastbound Good 1 core, core 15 mm rutting in OWP, Two fatigued sections sampled
0.9 m LHS FL disintegrated 7 mm in IWP, LHS with 30 mm and 70 mm thickness.
Observed as dry. 37.5 mm stone
size. Subgrade observed to be
dry.
52.20 Westbound Good 1 core, core 16 mm rutting in OWP, Disintegrated on coring. Observed
0.8 m RHS FL disintegrated. 7 mm in IWP, RHS as dry. 37.5 mm stone size.
53.77 Eastbound Good 1 core, core 2 mm rutting in OWP, Disintegrated on coring. Observed
0.9 m LHS FL disintegrated. 2 mm in IWP, LHS as dry. 37.5 mm stone size.
53.77 Westbound Good 1 core, core 5 mm rutting in OWP, Disintegrated on coring. Ceased
0.8 m RHS FL disintegrated. 6 mm in IWP, RHS coring due to crumbling. Observed
as dry. 37.5 mm stone size.
54.60 Eastbound Good 1 core, core 7 mm rutting in OWP, Disintegrated on coring. Ceased
0.9 m LHS FL disintegrated. 3 mm in IWP, LHS coring due to crumbling. Observed
as dry. 37.5 mm stone size.
54.60 Westbound Good 1 core, core 5 mm rutting in OWP, Disintegrated on coring. Ceased
0.8 m RHS FL disintegrated. 7 mm in IWP, RHS coring due to crumbling. Observed
as dry. 37.5 mm stone size.
55.20 Eastbound Good 1 core, core 5 mm rutting in OWP, Disintegrated on coring. Ceased
0.9 m LHS FL disintegrated. 3 mm in IWP, LHS coring due to crumbling. Observed
as dry. 37.5 mm stone size.
55.20 Westbound Good 1 core, core 6 mm rutting in OWP, Disintegrated on coring. Observed
0.8 m RHS FL disintegrated. 6 mm in IWP, RHS as dry. 37.5 mm stone size.
Subgrade observed to be dry.

Figure 5.18: Core hole showing large aggregate size Figure 5.19: Broken down material extracted
from core hole

Austroads 2020 | page 44


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 5.20: Material extracted from core hole Figure 5.21: Partial core extracted with layers
of intact material

5.5 Conclusions
The performance of LBC bases in Queensland was investigated and the following conclusions were drawn:
• LBC bases have been used in Queensland for over 10 years generally produced using in situ stabilisation
but also including plant-mixed LBC for use as base.
• From the field inspections conducted it can be concluded that, plant mixed LBC bases have performed
well. As such materials are placed as part of new construction, it is reasonable to assume that the
underlying support to the LBC base in a new pavement design was a significant factor in the good
performance of these projects. However, in some cases the thin LBC bases were constructed on
cemented materials subbases and for these project transverse cracking was observed on the surface
most likely be due to shrinkage of the cemented material subbases.
• The in situ stabilised LBC bases showed a much wider range of performance. All bases that were well
supported demonstrated good performance, as did some pavements where the underlying support could
be considered as inadequate from a structural design point of view.
• However, some in situ stabilised LBC bases exhibited very poor performance with the formation of major
block cracking evident. Additionally, some sections found to have been extensively patched, indicating
that significant distress had been evident and the pavement had needed treatment to restore
serviceability. It has been postulated that the application of excessive binder has resulted in a continuous
length of one project exhibiting extreme block cracking in all wheelpaths, but evidence in support of this
could not be found. Inadequate subbase support to the LBC base was suggested as a cause of the poor
performance. Current TMR (2018) design procedures reflect this experience as now the LBC bases
needs to be supported by a granular subbase with a minimum design modulus of 150 MPa.

Austroads 2020 | page 45


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

• Several sites were cored to improve understanding of the required properties of good and poorly
performing LBC bases. For sites with good performance with no crocodile cracking, it has not been
possible to extract intact cores and they disintegrated either due to traffic loading or during coring. Two
sites where cores could be extracted from poor performing areas with visible crocodile cracking. These
findings are consistent with TMR design concepts for LBC bases, namely to design and construct a low
strength material that develops fine macro-cracking with sufficient base thickness and subbase support to
limit the extent that micro-cracking leads to macro-cracking.
• Long lengths of LBC bases showed little or no signs of block cracking or crocodile cracking, despite the
considerable passage of time and loading since construction. Further improvement to design and
construction practices should reduce the risk of premature distress.

Austroads 2020 | page 46


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

6. In situ Modulus of LBC Bases

6.1 Introduction
The mechanistic-empirical thickness design method for flexible pavements (Austroads 2017a) includes two
critical components:
• a linear elastic response-to-load model in which critical strains are calculated
• performance relationships that estimate the allowable traffic loadings to pavement distress.

The response-to-load of a pavement structure requires the elastic characterisation of pavement layers and
the subgrade. Consequently, to extend the current design procedures to pavements containing LBC
materials, methods needed to be developed to determine the design modulus of LBC.

To assist the development, two field trials were constructed and monitored over the first year of life to enable
the change in moduli with curing and early-life trafficking to be investigated.

6.2 Field Trial Site A – Bruce Highway, Collinson’s Lagoon

6.2.1 Location

In March 2017, the first trial section was constructed and opened to traffic and the performance monitored over
12 months. The trial site was a 100 m long section of the Bruce Highway between Ayr and Townsville in
Queensland that was opened to traffic. The trial section was as part of in situ cementitious stabilisation works
from chainage 13 300 to 14 400 m at Collinson’s Lagoon (Figure 6.1). This site had suitable clearance from the
abutment of the bridge at chainage 13 955 m and also from the Lochinvar Intersection at chainage 13 650 m.

Figure 6.1: Aerial view of Bruce Highway at Collinson’s Lagoon

Townsville

Trial location

Bruce Hwy –
Ayr Lochinvar Rd
junction

Source: J805 – Pavement Rehabilitation Design Report.

Austroads 2020 | page 47


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

6.2.2 Historical Data and Design Traffic

The original pavement design was conducted in 1992. It was based on a design traffic of 4.1 x 106 ESA’s
and a subgrade design CBR of 3%. According to the pavement design manual in use at that time
(Queensland Department of Transport and Main Roads 1990), the required thickness of cover over the in
situ subgrade is 625 mm, which is exactly that recorded in the TMR plans (Figure 6.2a). The original
pavement structure comprised 300 mm of base gravel on 325 mm of select fill. Based on the design chart in
the TMR manual, the quality of the selected embankment material should have been a minimum CBR of
20%.

In 2016, the design consultant’s pavement design report recommended in situ cement stabilisation to a
depth of 250 mm to upgrade the existing road pavement. For the design of this rehabilitation treatment, TMR
provided NPC (Northern Pavement Consultants) with 2012 traffic data for the 10L link. Traffic data from site
91699 located at the Greenacres weigh-in-motion site (Ch. 20 640 m), was adopted.

The 10 and 20-year design traffic loadings for the project were calculated based on this data assuming year
1 to be 2015. The adopted 10 and 20-year design traffic loadings were 4.9 x 106 ESA and 1.1 x 107 ESA,
respectively (see Table 6.1).

Table 6.1: Design traffic summary

Project Two way Per cent Heavy Number of Design traffic (ESA)
AADT heavy vehicle heavy vehicle
vehicles growth axle groups 10 years 20 years
(HV) rate (%) per heavy
vehicle
Bruce Highway 10L 2614 16.83 2.5 2.80 4.9 x 106 1.1 x 107

6.2.3 Existing Pavement Before Treatment

Pavement Composition

TMR ARMIS (A Road Management Information System) data provides details of the pavement structure
from 1992. The pavement structure in the trial section length consisted of 300 mm of granular material on top
of a minimum of 325 mm of selected embankment material (Figure 6.2a).

In 1993, a 7 mm sprayed bituminous seal surfacing was placed, followed by a 16 mm sprayed seal. A 10 mm
sprayed seal was used to seal the pavement. In 1999 an additional polymer modified binder (PMB) sprayed
seal was placed.

In 2014, Northern Pavement Consultants undertook a pavement investigation in preparation for the
pavement upgrade work (250 mm in situ cement stabilisation). The pavement at the test pit closest location
to the trial section consisted of a 40 mm thick sprayed bituminous seal surfacing over 330 mm of sandy
gravel base and 240 mm of sandy gravel subbase (Figure 6.2b). The subgrade was identified as a silty clay
with sand.

Austroads 2020 | page 48


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 6.2: Pavement composition before in situ stabilisation

b. Following NPC pavement investigation


a. According to TMR 1992 plans
(test pit P4-B – Ch. 13 789)
Bitumen Seal

300 mm Type 2.1 Base Gravel

Minimum 325 mm selected


Embankment Material

Subgrade (Design CBR 3%)

Deflections before treatment

FWD surface deflection measurements were carried out on this section of road in 2014 before cement
stabilisation. At this time, the recorded maximum deflections (D0) ranged from 0.6 to 1.0 mm and curvatures
(D0 – D200) from 0.23 to 0.43 mm (see Figure 6.3).

Figure 6.3: Maximum deflections and curvatures under 40 kN FWD load in 2014

Austroads 2020 | page 49


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

6.2.4 Site Works

Construction phasing

In December 2016, the contractor milled off the existing bituminous surfacing, followed by pulverisation of the
gravel base on the southbound lane between chainage 13 730 and 13 949 m. The target depth for the
pulverisation was 200 mm. This was followed by light compaction and grading to profile, before spreading
top up gravel base, which was about 40 mm thick. Profiling and grading were then undertaken to bring the
levels close to that of the design, in preparation for the in situ stabilisation.

In situ stabilisation was specified as 2.5% type GB cement. This binder content appears to have been
selected based on UCS test results of field cores rather than laboratory-manufactured specimens (see
discussion below). The stabilisation of the southbound lane between chainage 13 730 and 13 949 m was
carried out on 20 December 2016.

Undesirably on the day of stabilisation, the contractor elected to split the stabilising work within the lot into
two stages. A construction joint was created in the works at chainage 13 839 m, close to the middle of the
selected trial section (chainage 13 800 to 13 900 m). Subsequent sampling and testing in this research
project was then undertaken away from this joint. However, it should be kept in mind, especially when
interpreting the deflection results (Section 6.2.6) at chainage 13 840 m.

The stabilised pavement was surfaced with a sprayed bituminous surfacing, details of which are not
available.

In March 2017, due to excessive roughness, a 50 mm thick size 14 mm dense-graded asphalt (AC14M)
surfacing layer was placed.

The asphalt surfacing was placed after the 28-day deflection measurement, but before the retesting
105 days after in situ stabilisation. Note that it was intended to measure deflections at 90 days but this was
delayed due to Cyclone Debbie and its aftermath. Table 6.2 summaries the construction phases.

Table 6.2: Summary of the construction phases

Description of the work Date


Bituminous surfacing milling Early December 2016
Southbound lane between chainages 13 730 and 13 949 m
Gravel top up (thickness 35 and 40 mm) 19/12/2016
Profiling and grading before in situ stabilisation
2.5% type GB cement stabilisation (depth = 250 mm) 20/12/2016
Sprayed sealing To be updated
Asphalt surfacing (50 mm AC size 14 mm) 02/03/2017

The pavement composition after sealing and the subsequent asphalt surfacing of the trial pavement section
are shown in Figure 6.4.

Austroads 2020 | page 50


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 6.4: Pavement composition after in situ stabilisation and surfacing

a. After sealing (21/12/2016) b. After asphalt surfacing (02/03/2017)

Bitumen Seal 50 mm AC14M Asphalt Surfacing layer

250 mm 2.5% insitu cement stabilisation


250 mm 2.5% insitu cement stabilisation

80 mm Sandy Gravel Base 80 mm Sandy Gravel Base

240 mm Sandy Gravel Subbase 240 mm Sandy Gravel Subbase

Subgrade (Silty Clay with Sand) Subgrade (Silty Clay with Sand)

Sampling the untreated material and binder

During construction, a bulk sample of the gravel base was sampled from the pavement approximately 1.0 m
from the centreline at chainage 13 860 m. This material was excavated following light compaction and grader
profiling after top-up gravel was placed. Approximately two tonnes of material was obtained using a small
tracked excavator. This material was delivered to TMR’s Townsville Materials Laboratory and subsequently
sent to ARRB in early May 2017.

The cement used for the stabilisation was also sampled during construction for laboratory testing. Due to the
difficulties in safely sampling raw cement from the on-site spreader, arrangements were made directly with
the supplier (Wagners Townsville) to sample the cement during the loading of the spreader truck.
Approximately 200 kg of cement was taken and placed in sealed containers. This was delivered to TMR’s
Townsville Materials Laboratory and subsequently sent to ARRB’s laboratory in Melbourne in early
May 2017.

Contractor’s quality control data

During the construction, quality control information was collected and compiled by the contractor for the test
lot that encompasses the test section (Lot SPRHS001).

Base gravel for the ‘top-up’ was sourced from BQC Quarries, which provided material quality test results.
The contractor’s surveyor supplied data on probe measurements of the depth of the stabilisation achieved.
The stabilisation depths from a range of offsets ranged from 235 to 289 mm with an average of 259 mm.

A number of density measurements were taken on the materials at various locations along with the
construction. The achieved dry density ratios were between 101.2% and 104.1% of standard Proctor MDD,
with an average density ratio of 102.7%.

Other data about the project included the cement spread rate, which was calculated as 13.1 kg/m2 and the
reported achieved rate was 13.0 kg/m2. However, the 30 minutes maximum time limit between cement
spreading and mixing was exceeded. This was conditionally approved by TMR subject to road roughness
measurements. The final surface levels exceeded the specified tolerance being up to 16 mm above and
14 mm below design levels. TMR delayed the decision on acceptance until the road roughness was
evaluated.

Austroads 2020 | page 51


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

The measured road roughness figures exceeded the permissible 60 NAASRA roughness counts. As a result,
a 50 mm thick dense grade asphalt surfacing was placed in early March 2017. This asphalt surfacing
covered both the southbound and northbound lanes from the project commencement at chainage 13 300 m
to the first bridge (chainage 13 950 m) which covers the research test section.

UCS testing

Table 6.3 summarises the UCS testing undertaken by TMR of field-mixed material sampled behind the
stabiliser and later compacted in the laboratory. Note that the test specimens were compacted below
100% standard Proctor MDD, perhaps due to the time delay between field mixing and compaction in the
laboratory or lower field moisture compared to standard Proctor OMC.

Table 6.3: TMR UCS results of laboratory cylinders at Bruce Highway Collinson’s Lagoon

Chainage Offset Achieved dry density Relative compaction(1) 28 day UCS (MPa)
(t/m3) (%)
Individual Reported
13 860 1.0 m RHS of 2.06 94.5 1.3 1.4
centreline behind
2.06 94.3 1.4
stabiliser
2.04 93.6 0.8(2)
13 890 1.0 m RHS of 2.05 94.3 2.0 2.0
centreline
2.05 94.4 1.7
behind stabiliser
2.06 94.8 2.3

1 Relative to standard Proctor maximum dry density of 2.17 t/m3.


2 This cylinder was damaged, and hence excluded from the calculation of the average figure.

At chainage 13 860 m, the average 28-day UCS was 1.4 MPa at a relative compaction of about 94% of
standard Proctor MDD. At chainage 13 890 m it was slightly higher at 2.0 MPa.

The contractor also measured the UCS testing at 7 days of field mixed and laboratory compacted cylinders
(Table 6.4). In this case the material was compacted close to 100% standard Proctor MDD. Allowing for the
increase in UCS with time, the estimated value at 28 days would be close to 2 MPa.

The contractor also tested cylinders 28 days after construction. This data is reported in Table 6.4, again the
results are close to the TMR upper limit of 2 MPa.

Table 6.4: Contractor’s UCS results at Bruce Highway Collinson’s Lagoon

Chainage (m) Specimen Curing Achieved dry Relative UCS (MPa)(2)


/Offset (days) density compaction(1)
(t/m3) (%) Individual Reported

13 880/3.4 m Field mixed 7 2.18 100.5 1.5 1.5


RHS of cylinders
2.20 101.4 1.5
centreline B89677
2.22 102.3 1.5
Field mixed 28 2.15 99.1 1.9 1.9
B89678
2.19 100.9 1.9
2.18 100.5 2.0
Field mixed 28 2.13 98.2 1.7 1.8
cores
2.17 100.0 1.7
B89679
2.18 100.5 1.9

1 Relative to standard Proctor maximum dry density of 2.17 t/m3.


2 On the report sheets, the additive content is mistakenly reported as 2%, whereas it was 2.5%.

Austroads 2020 | page 52


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

6.2.5 Surface Deflection Measurements

During the project, FWD deflection testing was undertaken over the 100 m trial section at 28 days, 90 days
and 1 year. The FWD testing was carried out at 10 m intervals both between the wheelpaths (BWP) and in
the outer wheelpath (OWP). All deflection data below and in Appendix B have been normalised to values
under a 40 kN FWD load.

An important element of the monitoring program was to investigate how the modulus of the cement-stabilised
layer changed with time and traffic loading. Accordingly, in this section the deflections measured at various
times (28 days, 105 days and 1 year) are compared.

The analysis of the change in FWD deflection was complicated by the placement of the 50 mm thick asphalt
surfacing which was not present when the 28 days testing was conducted but present when the 105 days
and 1 year measurements were conducted.

In addition to this discussion of deflection trends, Section 6.2.6 describes the trend in the modulus
back-calculated from the deflection bowls.

Maximum deflections

Figure 6.5 shows the variation in maximum deflections with time. The highest D0 values were measured at
chainage 30 and 40 m (13 830 and 13 840 m), which corresponds to a construction joint. The deflections in
the OWP followed a similar trend, although they were slightly lower than the values between the wheelpaths.

As expected the 28-day deflections are higher than the 105 day and 1 year. The 28-day measurements were
taken before placement of the 50 mm asphalt surfacing. Had the asphalt surfacing existed it would be
expected the maximum deflections would have been about 15% lower, as shown in the adjusted 28-day
values plotted.

It was concluded that the deflections in the untrafficked areas decreases up to 105 days and thereafter
remained similar. This is consistent with the expected curing of the cementitious binder with time.

Figure 6.5: Comparison of maximum deflections between the wheelpaths

Austroads 2020 | page 53


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 6.6 shows the trends in outer wheelpath deflections, including the 28-day deflections adjusted to the
estimated values with a 50 mm asphalt surfacing.

Unlike the untrafficked area, there was no clear trend in the deflections with time and loading given the
scatter of results. It maybe that the effect of curing in the initial lowing of the deflections has been offset by
traffic-induced damage to the stabilised layer.

Figure 6.6: Comparison of maximum deflections in the outer wheelpath

Curvature

Figure 6.7 shows the variation in curvature (D0-D200) with time. As expected, the 28-day curvatures were
higher than the 105-day and 1-year values. The 28-day measurements were taken before the placement of
the 50 mm asphalt surfacing. Had the asphalt surfacing existed it would be expected the curvatures would
have been about 25% lower, as shown in the adjusted 28-day values plotted.

Consistent with the untrafficked maximum deflection results, it was concluded that the curvatures in the
untrafficked areas decreased up to 105 days and thereafter remained similar. This is consistent with the
expected curing of the cementitious binder with time.

Austroads 2020 | page 54


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 6.7: Comparison of curvatures between the wheelpaths

Figure 6.8 shows the trends in OWP curvatures, including the 28-day curvatures adjusted to the estimated
values with a 50 mm asphalt surfacing.

Unlike the untrafficked area, there is no clear trend in the deflections with time and loading. It may be that the
effect of curing in the initially lowing of the deflections has been offset by traffic-induced damage to the
stabilised layer.

Figure 6.8: Comparison of curvatures in the outer wheelpath

Austroads 2020 | page 55


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

6.2.6 FWD Back-calculation of Modulus Values

Introduction

An important element of the monitoring program was to investigate how the modulus of the cement-stabilised
layer changed with time and traffic loading. Accordingly, in this section the layer moduli back-calculated from
the deflections measured at various times (28 days, 105 days and 1 year) were compared.

To enable the modulus of the various pavement layers to be estimated, the ARRB modulus back-calculation
software. This back-calculation process requires several inputs. Firstly, the pavement structure is required.
The information about the previously existing structure and the cement stabilisation process was used to
determine the expected structure. As noted above a 50 mm thick asphalt surfacing course was applied
between the 28-day and 90-day deflection testing. This has resulted in a structure immediately after
construction and a second structure, which was used in the analysis of the 90-day and 1-year deflection
data. These structures are shown in Figure 6.4a. It can be seen that the initial sprayed seal surface
pavement had 250 mm of cement-stabilised base, over 80 mm of a sandy-gravel base, 240 mm of a sandy
gravel subbase and a silty clay subgrade. After the addition of the asphalt surfacing on 2 March 2017, the
structure remained the same, but with an additional 50 mm thick surface of dense-graded asphalt
(Type AC14M) as shown in Figure 6.4b.

It is general practice during modulus back-calculation to split the subgrade into three separate layers, an
upper layer of 300 mm thick, a middle layer of 500 mm thick and a lower semi-infinite layer
(Austroads 2019a).

Table 6.5 and Table 6.6 summarise the pavement compositions used in the modulus back-calculation and
the seed moduli used to commence the layer moduli iteration. Note that, due to the very low deflections and
curvatures, a high seed modulus of 10 000 MPa was used for cement-stabilised layer.

Table 6.5: Pavement structure and seed moduli with sprayed seal surface

Material Thickness Seed modulus


(mm) (MPa)
Sprayed seal surfacing – –
Cement-treated base (2.5% cement) 250 10 000
Upper granular material 160 350
Lower granular material 240 180
Subgrade (0–300 mm) 300 120
Subgrade (300–800 mm) 500 180
Lower subgrade Semi-infinite 280

Table 6.6: Pavement structure and seed moduli with an asphalt surfacing

Material Thickness Seed modulus


(mm) (MPa)
Asphalt surfacing 0 (28 days) 0 (28 days)
50 (105 days and 1 year) 4000 (fixed)
Cement-treated base (2.5% cement) 250 10 000
Upper granular material 160 350
Lower granular material 240 180
Subgrade (0–300 mm) 300 120
Subgrade (300–800 mm) 500 180
Lower subgrade Semi-infinite 280

Austroads 2020 | page 56


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Moduli between wheelpaths

The back-calculation generated moduli values for between the wheelpaths are reported in Table 6.7,
Table 6.8, and Table 6.9. It is noted that the cement-stabilised layer has a very high modulus. These values
are significantly higher than would normally be expected for a LBC material. It was concluded that the
Collison’s Creek cemented material has a modulus more closely aligned with a HBC material rather than
LBC.

Table 6.7: 28-day back-calculated moduli between wheelpaths

Modulus

Chainage Subgrade Subgrade


Upper Lower Subgrade
(m) Base (upper (middle
granular granular (lower)
(MPa) 300 mm) 500 mm)
(MPa) (MPa) (MPa)
(MPa) (MPa)
0 8 294 500 273 160 244 401
10 8 301 476 274 145 241 420
20 5 684 500 188 326 301 255
30 2 428 200 300 249 238 217
40 2 929 361 83 269 245 202
50 3 804 200 300 189 197 212
60 4 929 285 164 93 140 228
70 4 103 480 300 166 197 255
80 6 517 200 80 349 311 242
90 4 194 336 300 194 211 243
100 5601 485 300 176 190 217
Mean 5 162 366 233 211 229 263

Table 6.8: 90-day back-calculated moduli between wheelpaths

Modulus

Chainage Subgrade Subgrade


Upper Lower Subgrade
(m) Base (upper (middle
granular granular (lower)
(MPa) 300 mm) 500 mm)
(MPa) (MPa) (MPa)
(MPa) (MPa)
0 14 953 420 200 126 181 283
10 14 953 462 200 119 181 299
20 14 939 232 168 248 250 257
30 10 000 350 180 120 180 260
40 4 205 257 142 115 162 250
50 4 862 203 162 219 226 240
60 20 000 200 200 112 156 239
70 17 670 343 126 181 221 296
80 19 433 500 200 104 161 267
90 20 000 486 200 122 183 298
100 10 000 350 180 120 180 260
Mean 13 729 346 178 144 189 268

Austroads 2020 | page 57


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Table 6.9: 1 year back-calculated moduli between wheelpaths

Modulus

Chainage Subgrade Subgrade


Upper Lower Subgrade
(m) Base (upper (middle
granular granular (lower)
(MPa) 300 mm) 500 mm)
(MPa) (MPa) (MPa)
(MPa) (MPa)
0 17 004 500 220 121 190 320
10 12 653 428 192 119 179 291
20 11 828 365 300 131 186 289
30 10 357 233 79 124 168 251
40 7 326 235 128 103 149 234
50 8 187 277 144 105 153 245
60 6 328 276 300 157 175 209
70 9 332 200 67 350 326 236
80 16 137 344 241 123 169 256
90 16 977 287 57 260 263 268
100 24 126 373 56 133 182 274
Mean 12 750 320 162 157 195 261

Outer wheelpath moduli

Similar results were found for the back-calculation of moduli in the outer wheelpath (see Table 6.10,
Table 6.11, and Table 6.12). Again, it is noted that the cement-stabilised layer has a very high modulus.
These values are significantly higher than would normally be expected for a LBC base. It is concluded that
the Collison’s Creek cemented material has a modulus more closely aligned with a HBC material rather than
LBC.

Table 6.10: 28-day back-calculated moduli outer wheelpath

Modulus

Subgrade Subgrade
Chainage (m) Upper Lower Subgrade
Base (upper (middle
granular granular (lower)
(MPa) 300 mm) 500 mm)
(MPa) (MPa) (MPa)
(MPa) (MPa)
0 7 179 500 293 206 257 351
10 7 286 500 300 181 266 425
20 6 526 500 300 323 300 257
30 3 693 249 239 259 250 231
40 5 578 262 300 270 255 226
50 5 557 500 300 209 221 243
60 20 000 496 280 123 168 254
70 8 501 500 214 314 301 276
80 7 398 500 300 241 245 254
90 6 457 333 300 316 296 259
100 5 931 500 170 320 280 205
Mean 7 646 440 272 251 258 271

Austroads 2020 | page 58


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Table 6.11 90-day back-calculated moduli outer wheelpath

Modulus

Subgrade Subgrade
Chainage (m) Upper Lower Subgrade
Base (upper (middle
granular granular (lower)
(MPa) 300 mm) 500 mm)
(MPa) (MPa) (MPa)
(MPa) (MPa)
0 29 945 200 61 30 203 600
10 10 415 359 191 100 162 279
20 11 640 311 178 141 188 276
30 16 88 347 115 178 192 219
40 3 011 200 157 157 176 211
50 8 175 216 166 123 164 240
60 5 110 200 257 133 156 198
70 9 580 282 206 114 165 259
80 15 042 548 257 93 145 242
90 18 296 281 207 97 161 281
100 10 000 350 180 110 180 260
Mean 11 173 299 180 116 172 279

Table 6.12: 1 year back-calculated moduli outer wheelpath

Modulus

Subgrade Subgrade
Chainage (m) Upper Lower Subgrade
Base (upper (middle
granular granular (lower)
(MPa) 300 mm) 500 mm)
(MPa) (MPa) (MPa)
(MPa) (MPa)
0 19 167 369 184 58 174 392
10 14 953 462 227 99 178 326
20 15 013 276 180 181 219 291
30 10 000 350 180 120 180 260
40 10 000 350 180 120 180 260
50 6 044 208 215 235 230 222
60 20 891 500 112 178 202 246
70 9 957 500 96 222 238 269
80 20 070 500 300 133 184 281
90 20 152 325 300 131 184 284
100 22 076 200 185 84 148 267
Mean 15 302 367 196 142 192 282

6.2.7 Trial Site A Conclusions

The Collinson’s Lagoon trial site consists of a 100 m section between chainages 13 800 and 13 900 m along
the southbound carriageway of the Bruce Highway between Ayr and Townsville. This section included a
250 mm deep in situ cement stabilisation with 2.5% type GB cement content. This binder content was
selected without mix design testing of laboratory-mixed and compacted cylinders.

Austroads 2020 | page 59


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

The construction work was carried out in December 2016. Field-mixed material was sampled from behind the
stabiliser and compacted for UCS testing after 28 days of curing. Due either to the time delay between field
sampling and laboratory compaction or the lower field moisture compared to standard Proctor OMC, the
UCS test specimens were not compacted to 100% standard MDD. Consequently, there is uncertainty about
the 28-day UCS of laboratory-manufactured cylinders.

FWD monitoring was carried out after 28 and planned for 90 days (noted that the second monitoring was
actually carried out at 105 days due to adverse weather conditions) and after 1 year. This was carried out
between the wheelpaths and in the OWP.

It was concluded that cement-stabilised layer had very high in situ modulus well over that expected for a
LBC. It was doubtful the moduli at this site was representative of LBC base and hence the results were not
used in developing the proposed design method (Section 9).

6.3 Field Trial Site B – Bruce Highway Barratta Creeks

6.3.1 Location

The second trial section was on the Bruce Highway between Ayr and Townsville near Barratta Creeks. The
project was on a section between chainage 19 400 and 23 200 m. Initially, the chosen 100 m long test
section was from chainage 23 000 to 23 100 m in the northbound lane. However, the contractor chose to
split construction within this section at 23 020 m and so the test section was adjusted to chainage 23 020 m
to chainage 23 120 m, which still gave suitable clearance from the abutment of the West Barratta Bridge at
chainage 23 185 m. An aerial view of the trial section is shown in Figure 6.9.

Figure 6.9: Aerial view of the trial section

Barratta Creek Bridge

East Barratta Creek bridge

Trial location

Source: J765 Pavement Rehabilitation Design Report – Historical Data and Design Traffic.

The 1997 plans provided by TMR indicate that the original pavement design was based on a design traffic
loading of 4.0 x 106 ESA and a subgrade design CBR of 3%. The pavement design manual then in use in
1997 (Queensland Department of Transport and Main Roads 1990), indicates that the total cover thickness
over the subgrade of 620 mm would have been required.

Austroads 2020 | page 60


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Traffic census data from 2011 provided by TMR was used to calculate design traffic values for 10 and
20-year periods. Data was collected from the WIM station at Greenacres (i.e. Ch. 20 640 m). A summary of
the data, assumptions and adopted design loadings are included in Table 6.13. The 10-year and 20-year
design traffic loadings for this section of the project were calculated based on this data assuming year 1 to
be 2015. The adopted design traffic loadings for 10 and 20-year periods were 4.7 x 106 ESA and
1.2 x 107 ESA respectively.

Table 6.13: Field Trial Site B design traffic summary

Two-way Heavy Annual HV Number axle ESA per heavy Design traffic (ESA)
AADT vehicles growth rate groups per vehicle axle
(%) (%) heavy vehicle group 10 20
years years

2567 15.2 4 2.84 2 4.8 x 106 1.2 x 107

6.3.2 Existing Pavement Before Treatment

Pavement composition

Figure 6.10a shows the pavement structure before in situ stabilisation based on TMR ARMIS data defining
the pavement structure and TMR plans. Based on this information, the original pavement consisted of a
sprayed seal surfacing over 250 mm of type 2.1 base gravel, and 370 mm of type 2.5 subbase gravel. There
were variable thicknesses of select fill under the subbase.

In 2014, Northern Pavement Consultants (NPC) investigated the existing pavement. This involved digging
test pits at various locations along the road, including chainage 23 057 m, which is within the test section. At
this chainage the pavement comprised a 70 mm thickness of bituminous seals over 200 to 220 mm thickness
of sandy gravel base and 350 mm of sandy gravel subbase(Figure 6.10b). The subgrade was a sandy clay.

The ARMIS database does not reveal any records of asphalt being placed at this location. This database
reveals a surfacing depth of 41 mm, comprising of a 10 mm and 16 mm spayed bituminous seal in 1996 and
a 15 mm thick slurry seal in 2010.

Figure 6.10: Pavement composition before in situ stabilisation

b. Following NPC pavement investigation in 2014


a. According to TMR plans in 1997
(chainage 23 057)

Austroads 2020 | page 61


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Deflections measured before stabilisation

In 2013 before treatment the surface deflections were measured using the FWD and reported by NPC. The
40 kN maximum deflections and curvatures are shown in Figure 6.11.

From the 2013 measured deflections and the pavement structure, NPC back-calculated the layer moduli
before rehabilitation. NPC concluded that the sandy gravel base had a vertical modulus of 300 to 400 MPa,
whereas the modulus of the sandy gravel subbase was only 50 to 100 MPa. The subgrade modulus was
estimated to be 100 to 130 MPa, but this may have been over-estimated as the subgrade was not
sub-layered.

Figure 6.11: Measured maximum deflections and curvatures at Bruce Highway – Barratta Creeks prior to
stabilisation

6.3.3 Site Works

Preliminary mix design

A series of UCS tests were carried out on gravel base recovered from a test pit at chainage 23 820 m, which
is close to the test section. For this testing the existing in situ gravel base was mixed with an appropriate
proportion of the expected ‘top-up’ gravel (type 2.1) from ‘Pats Pit’, just west of Townsville. The proportions
of this mix were 1 part in situ gravel to 0.36 parts imported gravel.

The 28-day UCS test results of laboratory-manufactured test cylinders were:


• 0.7 MPa with 1% type GB cement
• 1.0 MPa with 2% type GB cement
• 1.3 MPa with 3% type GB cement.

Based on this preliminary testing it was agreed cement stabilisation was appropriate, with a binder content in
the range 2–3%. Later a binder content of 2% was selected.

The test material had a plasticity index of about 7 with about 5–10% passing the 75 µm sieve.

Austroads 2020 | page 62


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Description and phasing of the work

The 100 m long trial section fell within the contractor’s lot SC-6A-1, which covered chainage 22 860 to
23 185 m in the northbound lane. Within this lot, there were no asphalt patches requiring removal and
replacement with top-up gravel.

On 6 March, the existing bituminous surfacing was removed and the gravel base pulverised to a target depth
of 200 mm. This was followed by light compaction and grading to profile, before selected spreading of top-up
gravel only between chainage 23 100 and 23 120 m at an average thickness of 25 mm, which was a lower
proportion of top gravel than used in the mix design.

Profiling and grading proceeded to bring the levels close to design, in preparation for the subsequent in situ
stabilisation process.

In situ cement stabilisation was originally specified as 2.5% type GB cement (by weight) based on the mix
design results. It was agreed to monitor the UCS results during construction to consider whether any
modification was required to this binder contents as it was considered close to that required to produce a
HBC material rather than the intended LBC material.

Based on the progressive UCS testing from the field-mixed material on earlier sections of the project, a
decision was taken before construction of the trial section to reduce the binder content to 2% type GB
cement.

The trial section was stabilised on 7 March 2017 to a target depth of 250 mm.

After construction and before the 28-day deflection measurements, a primerseal was placed. A double seal
was then applied after the 28-day deflections measurements and before the 90-day measurements.
Figure 6.12 details the pavement structure after construction.

Figure 6.12: Pavement composition after stabilisation

Sampling of the gravel base

During the construction, a bulk gravel sample was extracted from the inside run of the stabiliser at chainage
23 070 m. Approximately 2 tonnes of material was obtained and loaded into three bulk bags. This material
was excavated from chainage 23 065 to 23 072 m over 2.2 m wide to a depth of 250 mm. No top-up gravel
was placed in this area instead the area was grader profiled and lightly compacted.

These samples were transported to TMR’s Townsville materials laboratory and subsequently sent to ARRB’s
laboratory in Melbourne in May 2017 for testing (Section 7, Section 8).

Austroads 2020 | page 63


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Sampling the cement

The type GB cement was supplied by Wagners in Townsville. Due to the difficulty of safely sampling raw
cement from the on-site spreader, the cement was sampled at the plant during the loading of the spreader
truck.

A cement sample of approximately 200 kg was taken and placed in sealed containers. This was delivered to
TMR’s Townsville materials laboratory and subsequently sent to ARRB’s laboratory in Melbourne in
May 2017.

Contractor’s quality control data

Relevant construction quality control information was collected and compiled by the contractor for the lot that
encompasses the project test section (Lot SC-6A-1).

The base gravel material which was sourced for the ‘top-up’ material came from the Pats Pit west of
Townsville and quality test results were provided. These results were historic test records from 2015 but
were issued to TMR for acceptance of the top-up gravel. The top-up gravel test results show compliance with
the Type 2.1 granular material specification requirements.

In situ density tests within the project test section were carried out at chainage 23 040 m and another two
tests at chainage 23 018 and 23 123 m, which are just outside the ends of the project test section. Dry
density ratios (based on standard compaction) of 103.2%, 101.8% and 101.0% respectively were recorded.
These were accepted as the specified minimum dry density ratio (standard Proctor compaction) was 100%.

The measured spread rate for cement was 10.4 kg/m2 which slightly exceeded the target rate of 10.3 kg/m2
associated with 2% cement content. The measured depth of stabilisation was recorded as 252 mm, 255 mm
and 256 mm, which corresponded well with the specified depth of 250 mm.

6.3.4 UCS Testing

TMR sampled field mixed material behind the stabiliser at chainage 23 040 and then compacted test
cylinders in the laboratory for UCS testing after 7 and 28 days laboratory curing. The results are summarised
in Table 6.14. The cylinders were compacted to a density considerably less than the target 100% standard
MDD, perhaps due to the time delay between field mixing and compaction in the laboratory or lower field
moisture compared to standard Proctor OMC.

Table 6.14: UCS results obtained by TMR at Bruce Highway – Barratta Creeks

Achieved dry Relative UCS (MPa)


Curing
Chainage Offset density Compaction(1)
period
(t/m3) (%) Individual Reported

23 040 1.6 m LHS of 7 days 2.01 96.6 1.3 1.3


centreline
2.01 96.6 1.3
1.98 95.2 1.2
28 days 1.94 93.3 1.2 1.2
1.95 93.8 1.1
1.94 93.3 0.9(2)

1 Relative to standard Proctor maximum dry density of 2.08 t/m3.


2 The test report mentions that this cylinder was damaged, and hence excluded from calculation of average value.

The mean 7-day was 1.3 MPa at about 96% standard MDD. The mean 28-day UCS value was slightly lower
at only 1.2 MPa in part due to the lower density of the test cylinders.

Austroads 2020 | page 64


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Also, the contractor sampled field-mixed material behind the stabiliser at chainage 23 040 and then
compacted test cylinders in the laboratory for UCS testing. The results are summarised in Table 6.15. The
7-day UCS values were lower than TMR values but that the mean 28-day UCS values was 1.5 MPa at about
95–96% standard Proctor MDD.

Note that the test specimens were not compacted to 100% standard Proctor MDD as provided in the current
Austroads definition of modified materials. Assuming, a 9% increase in UCS for each 1% increase in density
(White 2006), the contractor’s 28-day UCS would be about 2 MPa which is significantly different from the mix
design results.

Table 6.15: Summary of UCS test results measured by the contractor at Barratta Creeks

Achieved dry Relative UCS (MPa)


Chainage
Offset Specimen density Compaction(1)
(m) Individual Reported
(t/m3) (%)
23 040 1.0 m LHS B91259, 2.01 96.6 0.5 0.6
of 7 days
1.97 94.7 0.7
Centreline
2.04 98.1 0.6
B91260, 2.00 96.2 1.4 1.5
28 days
2.00 96.2 1.5
2.01 96.6 1.5
B91261, 1.96 94.2 1.4 1.5
28 day
2.01 96.6 1.5
1.96 94.2 1.5

1 Relative to Standard Proctor maximum dry density of 2.08 t/m3.

6.3.5 Surface Deflection Measurements

An important element of the monitoring program was to investigate how the modulus of the cement-stabilised
layer changed with time and traffic loading. Accordingly, FWD deflections were measured 28 days, 90 days
and 1 year after construction. All maximum deflections and curvatures reported below and Appendix C has
been normalised to the values under a 40 kN FWD load.

It should be noted that during the 28-day testing, the pavement surfacing was a primerseal. A double seal
was placed before the 90-day testing.

Maximum deflections

Figure 6.13 shows the variation with time of between wheelpath maximum deflections. Although the results
are scattered the data tends to show a decrease in deflections between 28 days and 90 days, as expected
for untrafficked areas subject to curing. However, between 90 days and 1 year the deflections increased. A
possible explanation is that the increases in modulus due to curing had largely been completed at 90 days
and thereafter the traffic-induced damage in the wheelpaths was affecting the deflections between the
wheelpaths.

Austroads 2020 | page 65


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 6.13: Maximum deflections between wheelpaths on Bruce Highway – Barratta Creeks section

Figure 6.14 shows the variation of the maximum deflections measured in the outer wheelpath. There was no
significant change in deflections between 28 days and 90 days, possibly due to the increase in modulus due
to curing being offset by the traffic-induced damage. Similar to the between wheelpaths results, the
maximum deflections increased considerably between 90 days and 1 year. Again, a possible explanation is
that the increases in modulus due to curing had largely been completed at 90 days and thereafter the
traffic-induced damage results in an increase in deflections.

Figure 6.14: Maximum deflections in outer wheelpath: Bruce Highway – Barratta Creeks section

Curvatures

Figure 6.15 and Figure 6.16 shows the curvature variations with time and loading. The trends are similar to
those discussed above for maximum deflections, albeit that the curvature results are more variable than the
maximum deflection values.

Austroads 2020 | page 66


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 6.15: Curvature between wheelpaths: Bruce Highway – Barratta Creeks section

Figure 6.16: Curvatures in the outer wheelpath: Bruce Highway – Barratta Creeks section

Austroads 2020 | page 67


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

6.3.6 Back-calculation of Modulus Values

Introduction

The pavement layer and subgrade moduli were back-calculated from the measured deflection bowls as
described in Austroads (2019a).

This modulus back-calculation process requires several inputs. Figure 6.12 shows the pavement structure
assumed at the time the deflections were measured. Consistent with Austroads (2019a), for the
back-calculation the subgrade was split into three separate layers:
• a top layer, 300 mm thick
• a middle layer, 500 mm thick
• a lower semi-infinite layer.

Initial moduli are required to commence the modulus iteration. These so-called seed moduli are shown in
Table 6.16.

Table 6.16: Pavement structure and seed moduli

Thickness Seed vertical modulus


Material
(mm) (MPa)
Sprayed seal surfacing – –
LBC base (2.0% cement) 250 1 000
Granular material 300 150
Subgrade (0–300 mm) 300 70
Subgrade (300–800 mm) 500 85
Lower subgrade Semi-infinite 100

Moduli between wheelpaths

The back-calculation moduli between wheelpaths are listed Table 6.17, Table 6.18and Table 6.19, whilst the
LBC base moduli are plotted in Figure 6.17.

The mean moduli of the lightly cemented base were 786 MPa, 875 MPa and 880 MPa after 28 days, 90 days
and 1-year curing respectively. Due to the scatter of results, there was no significant trend in modulus with
time that could be deduced.

Table 6.17: Back-calculated moduli between wheelpaths after 28 days

Granular Subgrade Subgrade Subgrade


Chainage LBC base
subbase (upper 300 mm) (middle 500 mm) (lower)
(m) (MPa)
(MPa) (MPa) (MPa) (MPa)
10 821 154 50 70 107
20 1 037 128 38 61 104
30 706 164 60 79 114
40 774 157 56 78 119
50 1 465 103 58 80 122
60 1 244 260 43 72 126
70 1 309 296 57 84 135
80 845 178 65 86 125
90 500 138 110 120 140
100 615 144 91 106 134
Mean 932 172 63 84 123

Austroads 2020 | page 68


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Table 6.18: Back-calculated moduli between wheelpaths after 90 days

Granular Subgrade Subgrade Subgrade


Chainage LBC base
subbase (upper 300 mm) (middle 500 mm) (lower)
(m) (MPa)
(MPa) (MPa) (MPa) (MPa)
10 780 225 58 78 115
20 1 074 136 50 72 113
30 866 204 78 95 127
40 1 087 183 68 89 128
50 1 437 139 40 70 127
60 1 231 492 41 71 128
70 895 281 37 62 110
80 893 301 80 100 136
90 662 187 118 130 151
100 866 187 95 112 144
Mean 979 234 67 88 128

Table 6.19: Back-calculated moduli between wheelpaths after 1 year

Granular Subgrade Subgrade Subgrade


Chainage LBC base
subbase (upper 300 mm) (middle 500 mm) (lower)
(m) (MPa)
(MPa) (MPa) (MPa) (MPa)
10 904 154 65 82 112
20 1 094 126 51 72 110
30 675 183 79 93 121
40 754 116 35 57 100
50 1 050 133 48 67 104
60 1 240 187 37 62 107
70 960 125 41 64 107
80 944 142 55 76 115
90 622 141 120 128 145
100 751 163 96 110 136
Mean 899 147 63 81 116

Austroads 2020 | page 69


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 6.17: LBC base modulus between wheelpaths

1500

1400
BWP 28 days
1300 BWP 90 days

1200 BWP 1 year

1100

LCM 1000
modulus
(MPa) 900

800

700

600

500

400
10 20 30 40 50 60 70 80 90 100
Chainage (m)

Moduli in outer wheelpath

The back-calculation moduli in the outer wheelpath are listed in Table 6.20, Table 6.21 and Table 6.22, and
LBC base moduli are plotted in Figure 6.18.

The mean moduli of the LBC base in the outer wheelpaths were 987 MPa, 818 MPa and 695 MPa after
28 days, 90 days and 1 year in-service. The results indicate a progressive reduction in modulus with traffic
loading notwithstanding the effect of curing.

Table 6.20: Back-calculated moduli in outer wheelpath after 28 days

Granular Subgrade Subgrade Subgrade


Chainage LBC base
subbase (upper 300 mm) (middle 500 mm) (lower)
(m) (MPa)
(MPa) (MPa) (MPa) (MPa)
10 694 98 41 57 87
20 824 81 44 60 92
30 810 130 52 69 100
40 1 244 153 59 79 116
50 1 534 55 80 86 98
60 1 523 204 38 62 106
70 1 481 179 48 71 113
80 998 145 40 64 108
90 1 216 109 142 140 137
100 1 357 96 105 115 134
Mean 1 168 125 65 80 109

Austroads 2020 | page 70


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Table 6.21: Back-calculated moduli in outer wheelpath after 90 days

Granular Subgrade Subgrade Subgrade


Chainage LBC base
subbase (upper 300 mm) (middle 500 mm) (lower)
(m) (MPa)
(MPa) (MPa) (MPa) (MPa)
10 740 70 65 73 89
20 724 70 58 71 94
30 768 70 88 95 106
40 1 160 70 107 110 117
50 1 142 70 58 75 106
60 1 735 70 70 82 105
70 1 235 70 55 74 109
80 1 197 70 84 94 113
90 1 577 70 150 171 142
100 1 211 70 150 147 131
Mean 1 149 70 89 99 111

Table 6.22: Back-calculated moduli in outer wheelpath after 1 year

Granular Subgrade Subgrade Subgrade


Chainage LBC base
subbase (upper 300 mm) (middle 500 mm) (lower)
(m) (MPa)
(MPa) (MPa) (MPa) (MPa)
10 457 97 47 63 93
20 557 71 50 65 94
30 532 104 75 86 105
40 782 70 75 84 100
50 774 77 41 61 97
60 993 84 36 57 95
70 948 72 39 58 94
80 761 73 50 67 99
90 849 86 189 168 129
100 1 020 70 149 142 127
Mean 767 80 75 85 103

Austroads 2020 | page 71


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 6.18: LBC base modulus in outer wheelpath

1800

1700
OWP 28 days
1600
OWP 90 days
1500
OWP 1 year
1400

1300

1200
LCM
modulus 1100
(MPa)
1000

900

800

700

600

500

400
10 20 30 40 50 60 70 80 90 100
Chainage (m)

6.3.7 Trial Site B Conclusions

The Barratta Creeks trial site was a 100 m section of the northbound carriageway of the Bruce Highway
between Ayr and Townsville, between chainages 23 020 and 23 120 m. This pavement consisted 250 mm of
in situ cement stabilisation with 2% type GB cement content. The pavement was constructed in March 2017
and surface deflections were monitored over the first year of opening to traffic.

Laboratory UCS testing showed that the stabilised material had a 28-day UCS in the range of a LBC,
i.e. 1–2 MPa.

FWD monitoring was carried out both between the wheelpaths and in the OWP after 28 days, 90 days and
1 year. From this data, the moduli of this LBC base were estimated.

The mean vertical moduli of the LBC base in the outer wheelpath were 1170 MPa, 1150 MPa and 770 MPa
after 28 days, 90 days and 1 year in-service respectively. In contrast, the moduli in the untrafficked area
between the wheelpaths were relatively constant: 930 MPa, 980 MPa and 900 MPa. The reduction in LBC
moduli in the OWP between 90 days and 1 year is consistent with the LBC being fatigue damaged due to
loading. This information was used to develop the proposed modulus characterisation used in the proposed
structural design of LBC bases (Section 9).

In the untrafficked area between the wheelpaths, there was no significant change in modulus with curing
time, which was unexpected. This may have been related to load-induced damage to the LBC base in the
wheelpaths influencing the measured deflections between the wheelpaths.

Austroads 2020 | page 72


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

7. Conventional Laboratory Characterisation


Testing

7.1 Introduction
As described in Section 6.1, the mechanistic-empirical thicknesses design method for flexible pavements
(Austroads 2017a) includes two critical components:
• a linear elastic response-to-load model in which critical strains are calculated
• performance relationships that estimate the allowable traffic loadings to pavement distress.

To supplement the early-life LBC in situ modulus investigation (Section 6), the following laboratory testing
was undertaken of laboratory-manufactured and compacted mixtures:
• UCS (Section 7.2)
• indirect tensile modulus (Section 7.3)
• indirect tensile strength (Section 7.4)
• flexural beam modulus (Section 7.5)
• flexural beam strength (Section 7.6).

In the testing program, the characteristics of the LBC (1%-2% cement content) were compared with the HBC
material containing 3–4% cement content.

Two host materials were used:


• a hornfels crushed rock from Lysterfield Victoria, meeting the Class 2 requirements of VicRoads
specification RC 500.02 (VicRoads 2017)
• the Barratta Creeks sandy gravel, which was stabilised during the field trial (Site 2) (Section 6.3).

7.2 Unconfined Compressive Strength (UCS) Testing


The UCS test is a compressive test where a cylindrical sample is loaded along its axis until failure. UCS
testing was conducted according to AS 5101.4-2008.

After mixing the cement and water into the host material the sample was conditioned for two hours. The
mixture was then poured into a split cylindrical mould and the compacted using modified Proctor compactive
effect. (Note that some road agencies use standard Proctor compactive effect, which results in lower
densities and lower UCS values.)

The UCS samples were capped with a fly ash and sulphur compound to provide uniform load to the sample.
The samples were soaked in water for 4 hours, and then left standing for 15 minutes before being tested.
The load rate was 60 kN/minute as specified in AS 5101.4-2008.

The testing included hornfels mixed with 1.5% and 3% cement contents with both type GB and type GP
cement, and Barratta Creeks gravel mixed with 2% and 4% type GB cement contents. These samples were
cured for either 1, 7, 14, 28 or 90 days before testing. Three samples were tested at each condition. The
UCS data are plotted as a function of curing time, in days, in Figure 7.1 and Figure 7.2.

Austroads 2020 | page 73


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Note that the method used to compact and test UCS specimens differed from that used by TMR. This
hindered an assessment of whether hornfels with 1.5% type GB cement and the Barratta Creeks gravel with
2% type GB cement complied with the proposed definition of a 28-day UCS in the range 1.0–2.0 MPa
(Section 9.3).

As seen from Table 7.1 and Figure 7.2, the 28-day UCS of the 1.5% type GB cement hornfels mixture was
3.3 MPa at 96.7% modified Proctor MDD. Assuming the UCS reduces by 9% for each 1% decrease in
density (White 2006) and that the standard Proctor MDD is about 5% lower than the modified Proctor MDD,
then the 28-day UCS at 100% standard MDD would reduce to about 2.8 MPa, which is above the proposed
maximum UCS limit for LBC (Section 9.3).

Table 7.1: UCS results

Host material Binder Binder Curing Achieved MDD Relative UCS


type content period dry density (Modified Compaction (MPa)
(t/m3) Proctor (%)
Hammer)
Hornfels 1.5 2.179 2.26 96.4 2.19
GB
3 2.189 2.29 95.6 4.21
1.5 2.290 2.27 100.9 2.47
GP 7 days
3 2.258 2.315 97.5 4.54
Barratta Creeks 2 1.960 2.097 93.5 3.31
gravel GB
4 1.958 2.13 91.9 4.35
Hornfels 1.5 2.184 2.26 96.7 3.32
GB
3 2.170 2.29 94.7 4.94
1.5 2.167 2.27 95.5 3.49
GP 28 days
3 2.304 2.315 99.5 6.49
Barratta Creeks 2 1.954 2.097 93.2 3.74
gravel GB
4 1.939 2.13 91.0 6.56

From Table 7.1 and Figure 7.2, the 28-day UCS of the 2% Barratta Creeks mixture was 3.7 MPa at
93% modified Proctor MDD. Assuming the UCS reduces by 9% for each 1% decrease in density
(White 2006) and that the standard MDD is about 5% lower than the modified MDD, then the 28-day UCS
would be about 4.3 MPa. By comparison with the field trial B testing (Section 6.3), the TMR and contractor’s
28-day UCS values of field-mixed laboratory compacted material at 100% standard MDD were about 2 MPa.

There is a need to improved national consistency in the preparation and testing of UCS cylinders.

Austroads 2020 | page 74


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 7.1: UCS of stabilised crushed hornfels

Figure 7.2: UCS of stabilised Barratta Creeks gravel

7.0

6.0

5.0
UCS (MPa)

4.0

3.0 GB_2%
GB_4%
2.0

1.0

0.0
0 20 40 60 80 100
Age (Days)

As part of the UCS testing of the stabilised crushed hornfels, the variation of displacement with load was
measured from which the stress-strain plots shown in Figure 7.3 and Figure 7.4 were derived. The drop in
stress was more pronounced in the higher cement content materials, indicative of a higher strength, more
brittle material.

Austroads 2020 | page 75


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 7.3: Stress-strain relationship during 28-day UCS testing

Figure 7.4: Stress-strain relationship during 90 day UCS testing

Austroads 2020 | page 76


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

7.3 Indirect Tensile Modulus Testing


Indirect tensile testing (ITT) of modulus was conducted on the crushed hornfels stabilised with 1.5% and
3% of both type GP and type GB cements.

The sample size of the test specimens was nominally 150 mm in diameter and 85 mm thick. Samples were
prepared in the laboratory and compacted to 100% modified Proctor MDD using gyratory compaction. The
ITT samples were cured for up to 90 days before testing. The ITT test method used was based on
AS 2891.13.1-2013, modified for cemented materials. To avoid damaging the specimens a target strain of
10 µm was used.

Figure 7.5 shows the results of the ITT modulus testing. The key findings were:
• The 90-day mean moduli of the 1.5% type GB cement mixture was about 9 800 MPa, whereas the mean
90-day modulus of the 3% type GB cement mixture was 19 200 MPa. For the type GB cement, the
modulus of the 1.5% mixture was roughly half the 3% mixture value.
• The 90-day mean modulus of the 1.5% type GP mixture was about 14 600 MPa whereas the mean
90-day modulus of the 3% type GP cement mixture was 20 800 MPa. The modulus of 1.5% cement
content mixture was about 70% of the 3% mixture value.
• The rate of strength gain in the type GP mixture was faster than the type GB mixture.

Figure 7.5: Stabilised crushed hornfels indirect tensile modulus values

25000

20000
Modulus (MPa)

15000
GB 1.5%
GP 1.5%
GB 3%
10000
GP 3%

5000

0
0 10 20 30 40 50 60 70 80 90 100
Curing time (days)

Austroads 2020 | page 77


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

7.4 Indirect Tensile Strength Testing


After modulus testing was completed the stabilised crushed hornfels samples were then subjected to tensile
strength testing. A constant loading rate of 20 ± 2 kN/min was applied until the samples could no longer
sustain any load. The test results are shown in Figure 7.6.

Findings from this testing were as follows:


• The 90-day mean strength of the 1.5% type GB cement mixture was about 0.22 MPa whereas the mean
90-day modulus of the 3% type GB cement mixture was about 0.61 MPa. For the type GB cement, the
strength of the 1.5% mixture was about a third of the 3% mixture value.
• The 90-day mean strength of the 1.5% type GP mixture was about 0.32 MPa whereas the mean 28-day
modulus of the 3% type GP cement mixture was 0.55 MPa. For type GP cement, the strength of the
1.5% cement-content mixture was about 60% of the 3% mixture value.

Figure 7.6: Stabilised crushed hornfels indirect tensile strength values

0.7

0.6

0.5
Strength (MPa)

0.4
GB 1.5%
GP 1.5%
0.3
GB 3.0%
GP 3.0%
0.2

0.1

0
0 10 20 30 40 50 60 70 80 90 100
Curing time (days)

7.5 Flexural Modulus Testing


Test beams were prepared by compacting loose mixtures using a segmental roller as described in Austroads
Test Method AGPT T600-18. The mixtures were prepared at modified Proctor optimum moisture content and
compacted to 100% modified Proctor maximum dry density.

The flexural modulus of each test beam was measured in line with Austroads Test Method AGPT-T600-18.

The crushed hornfels mixtures with 1.5% and 3% type GB cement were cured for 14 days before testing. As
seen from Figure 7.7, the results for the 1.5% mixture were very scattered with a mean modulus of about
5 800 MPa. The mean modulus of the 3% cement mixture was higher, about 14 300 MPa. The modulus of
the 1.5% mixture was about 40% of the 3% mixture modulus reasonably consistent with the indirect tensile
modulus findings (Section 7.3).

Austroads 2020 | page 78


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 7.7: Stabilised crushed hornfels flexural modulus values

Note: The ‘field beams’ and ‘laboratory beams’ points are those from Austroads (2008).

The crushed hornfels from the same quarry with 3% type GP cement was previously tested
(Austroads 2008). The modulus results of laboratory manufactured beams were similar, however, the flexural
moduli of field beams (Austroads 2008) were lower, consistent with the different curing that occurs in the
roadbed.

Similarly, beams of the Barratta Creeks material stabilised with 2% or 4% type GB cement and cured for 1, 7,
28 or 90 days were prepared. As seen Figure 7.8, the flexural moduli were very high, even for the
2% cement material, with the 28-day flexural modulus being about 15 000 MPa. The influence of cement
content on modulus was less significant than that observed for the crushed hornfels (Section 7.3).

Figure 7.8: Stabilised Barratta Creeks gravel flexural modulus values


20000

15000
Flexural modulus (MPa)

10000

5000

0
0 10 20 30 40 50 60 70 80 90 100
Age (days)
2.0% GB Barratta Creeks 4.0% GB Barratta Creeks

Austroads 2020 | page 79


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

7.6 Flexural Strength Testing


After flexural modulus testing (Section 7.5), the flexural strength of the stabilised crushed hornfels beams
were measured in accordance with AS 1012.11-2000: Methods of Testing Concrete: Determination of the
Modulus of Rupture. The testing included firstly applying a seating load of 50 N for the six seconds, after
which the load was increased at a rate of 3.3 kN per minute until the sample failed as described in AS
1012.11-2000. Figure 7.9 shows the test set-up.

The crushed hornfels stabilised with 1.5% and 3% type GB cement were prepared and cured for 14 days. As
seen from Figure 7.10, the flexural strength the 1.5% mixture was very low (about 0.2 MPa) which was about
one-third of the mean strength of the 3% mixture. This was reasonably consistent with the indirect tensile
strength findings (Section 7.4).

Also plotted in Figure 7.10 are results for the same hornfels stabilised with 3% type GP reported previously in
Austroads (2008). This shows that the 14-day strength of 3% type GB material was similar to the 7-day
strength to the 3% type GP material reported previously.

Figure 7.9: Flexural beam testing in the laboratory

Austroads 2020 | page 80


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 7.10: Stabilised crushed hornfels flexural strengths

Similarly, beams of the Barratta Creeks gravel stabilised were with 2% or 4% type GB cement and cured for
1, 7, 28 or 90 days. As seen Figure 7.11, the flexural strength of the 2% cement mixture was unexpectedly
high but consistent with the flexural modulus results. The influence of cement content on modulus was less
significant than that observed for the crushed hornfels and also the indirect tensile modulus testing
(Section 7.3).

Figure 7.11: Stabilised Barratta Creeks gravel flexural strengths

1.6
Flexural strength (MPa)

1.2

0.8

0.4

0
0 10 20 30 40 50 60 70 80 90 100

Age (days)
2.0% GB Barratta Creeks 4.0% GB Barratta Creeks

Austroads 2020 | page 81


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

7.7 Summary
In summary:
• The estimated 28-day UCS of the Barratta Creeks gravel stabilised with 2% type GB cement was greater
than 2.5 MPa, which may not have complied with the proposed Austroads upper limit of 2.0 MPa for LBC.
However, this UCS result was higher than that measured by TMR and the contractor before and during
construction (Section 6.3.4). This difference in UCS cannot be explained by the difference in the densities
of the test specimens. A national test method for the preparation of UCS test specimens is required to
improve the consistency of UCS testing.
• The estimated 28-day UCS of the crushed hornfels stabilised with 1.5% type GB cement was 2.5 MPa
when compacted to 100% standard Proctor compaction. This value was above the proposed upper limit
of 2 MPa for LBC. However, given the above findings for the Barratta Creeks material, had this material
been tested by TMR the UCS may have been lower.
• Concerning the relative moduli of a 1.5% type GB mixture compared to a 3% mixture, for the crushed
hornfels both the indirect tensile and flexural modulus testing indicated the modulus of the 1.5% mixture
was roughly half of the 3% mixture value. However, the flexural moduli of the Barratta Creeks gravel
2% mixture were unexpectedly high (about 14 000 MPa), which was similar to the 4% mixture value.
• Concerning the relative strength of a 1.5% type GB mixture compared to the 3% mixture, both the indirect
tensile and flexural strength testing of the crushed hornfels indicated that the modulus of the 1.5% mixture
was approximately one-third the 3% mixture. However, the flexural strength of the Barratta Creeks gravel
2% mixture was unexpectedly high (about 1 MPa at 90 days) which was only slightly lower than the
4% mixture value.

Austroads 2020 | page 82


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

8. Wheel-Tracking Fatigue Testing of Cemented


Materials

8.1 Introduction
The aim of the mix design of LBC materials is to inhibit shrinkage cracking and crocodile and ladder cracking
which can lead to durability issues of the sprayed sealing or thin asphalt surfacing. It is important to
understand the relationship between the material binder content and the type of cracking mechanism to
manage the cost-effectiveness of the technology.

At the commencement of the project, laboratory test methods to investigate the cracking mechanisms of
stabilised materials, and to differentiate between materials susceptible to block cracking as observed for
HBC bases (Section 4) as opposed to the absence of block cracking for LBC bases (Section 5), did not exist.

As described below, an extra-large wheel-tracking (XL-WT) test was developed to investigate cracking
mechanisms of cemented materials in the laboratory. A key advantage of the wheel-tracking test is that it
applies a rolling wheel load and provides support and confinement to the specimen which more closely
mimics field conditions than traditional tests.

This section describes the development of the testing equipment, the test method and the results of testing
stabilised materials with different binder contents.

8.2 Development of Test Method

8.2.1 Introduction

As part of the research project a test method for evaluating the fatigue characteristics of LBC was developed.
This draft methods are provided in Appendix D, Appendix E and Appendix F, the development of which is
described below.

8.2.2 Equipment

The extra-large wheel tracking (XL-WT) device (Figure 8.1) was designed and manufactured to test unbound
granular materials as part of a previous Austroads project (Austroads 2013a). This work provided the starting
point for the development of a wheel-tracking test using the same XL-WT device to apply heavy
wheel-loading to LBC to induce cracking.

To assess the cracking characteristics, the test was modified as described below to enable flexure of the
stabilised material by the provision of flexible support instead of the steel base plate used in rut-resistance
testing.

The dimensions of the stabilised material test slabs (Figure 8.2) are as follows:
• slab length of 700 mm (direction of compaction and wheel-tracking)
• slab width of 500 mm mould width (transverse to the direction of traffic)
• slab thickness typically 100 mm; the thickness can be adjusted considering the desired amount of slab
flexure and the maximum aggregate size.

Austroads 2020 | page 83


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 8.1: Extra-large wheel tracking device

Source: Photo courtesy of IPC Global.

Figure 8.2: Schematic view of test mould

Note: Where LDPE sheets are labelled, this support setup, for some specimens, was replaced with neoprene sheets.

To allow flexure of the test slab specimens under the rolling wheel load, two variations for supporting the
slabs were used.
• Setup 1: slabs were supported by layers of 9 mm thick, low-density polyethylene (LDPE) sheets (nominal
flexural modulus of 200 to 400 MPa (Design Flow 2019) depending on the grade).The total thickness of
LDPE is 200 mm when testing a 100 mm thick LBC slab. While slab support is slightly stiffer than LBC
bases in service, the LDPE still allows flexure of the test slabs deflections under rolling wheel loading.
• Setup 2: slabs were supported by an alternate sandwich support structure of 9 mm thick LDPE sheets
and 10 mm thick neoprene rubber sheets with a flexural modulus of 1.65–2.1 MPa (Course Hero 2020).

The mould assembly used for testing is shown in Figure 8.2. The steel base plate is designed to attach
displacement gauges as in Section 8.2.3. Around two-thirds of the mould is filled with LDPE (or neoprene)
sheets providing support for specimen compaction and testing.

In the XL-WT device, a rolling wheel load (Figure 8.3) is applied using a solid rubber tyre which is capable of
carrying loads up to 20 kN This load was applied along a length of 500 mm centred around the middle of the
700 mm slab length.

Austroads 2020 | page 84


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 8.3: Rolling wheel loading Figure 8.4: Planetary concrete mixer used to mix stabilised
materials

Source: Photo courtesy of IPC Global (2012). Source: Austroads (2014).

8.2.3 Method to Prepare Test Slabs

In a previous Austroads project (Austroads 2013a and b), a test method was developed to mix and compact
unbound granular materials for rut-resistance testing. This method provided the starting point for the
development of a method to mix and compact stabilised materials for fatigue testing.

Following a method previously used for cemented materials (Austroads 2014), the untreated material is firstly
mixed dry using a planetary mixer (Figure 8.4). The desired amount of water was then added, and mixing
was continued for two minutes. The material was then placed in large plastic containers and stored overnight
to allow homogenisation of the moisture content. The following day the required cementitious binder was
added followed by mixing using a planetary mixer for 10 seconds. Additional water was added to bring the
mixture up to optimum moisture content along with additional mixing for a further two minutes. The required
mass of material was then placed into the steel mould for compaction within 30 minutes of mixing being
completed.

The compaction of the test slabs was undertaken within the XL-WT device. The loose mixture was firstly
spread and tampered in the steel mould on top of either the LDPE or neoprene sheets Figure 8.2. A steel
compaction foot then folds down around the moulded rubber tyre and this was used to compact the material
into the mould (Figure 8.5). The speed of the compaction foot is slower than when testing the slab for fatigue
characteristics using the rubber tyre. The compaction was carried out in position control mode, where the
machine controls by height as opposed to load applied. This was found to help maintain a level surface
during the process. The target wet density was commonly determined as a percentage of standard or
modified Proctor maximum dry density to reflect in-service densities.

Austroads 2020 | page 85


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 8.5: Austrack Extra Large Wheel-tracking device a) with compaction foot down, b) in wheel-tracking
mode

a. Compaction foot b. Loading wheel

Specifically, the compaction is undertaken using the following steps:


• Step 1: lower compaction foot down until just touching the base of the empty mould. This gives a set-point
to determine the target height which will achieve the nominal slab thickness of 100 mm.
• Step 2: fill mould with material, spread the loose material uniformly by hand and tamping.
• Step 3 (compaction phase 1): lower compaction foot until it is just touching the material (applying a load
of approximately 2 kN), this will set the level to give the initial height of the material. From this it is known
how much compaction is required to reach the desired slab thickness of 100 mm. Then perform four
passes (two complete cycles) from the middle of the slab, finishing in the middle again. This will level the
material.
• Step 4 (compaction phase 2): reduce in 2 mm steps, performing four passes at this level. Repeat this step
until the load raises to 10 kN.
• Step 5 (compaction phase 3): from this point on reducing the slab thickness in 1 mm steps, performing
four passes at each level, until a slab thickness of 100 mm is reached.

The compacted specimens were cured in humidity-controlled conditions using a fog room (100% humidity)
for their relevant curing times before testing.

8.2.4 Fatigue Test Method

For fatigue testing, a load of 20 kN was applied to a solid rubber tyre. This loading regime differs from
Austroads Test Method AGPT-T054-15 used for the rut-resistance of granular materials (Austroads 2013a
and b). For the latter, an 8 kN load is applied to a pneumatic tyre, however this load was considered
insufficient to induce fatigue damage in stabilised materials. As such, it was decided to use a 20 kN tyre load
which is closer to the damaging tyre loads applied to in-service pavements. To apply this load level, it was
necessary to use a solid rubber tyre.

The wheel-tracking load is bi-directional, with a full wheel cycle taking 2.5 s for a travel frequency of 0.4 Hz.
The load is applied over a 500 mm length rather than the entire slab length of 700 mm.

The wheel-tracking was performed in two phases:


• For at least 100 000 cycles the specimen was tested at the moisture content after curing. Any visible
cracking at the surface was recorded daily (around each 30 000 cycles).

Austroads 2020 | page 86


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

• The second phase of the testing involves spreading water over the surface of the specimen. The test was
stopped regularly, and water was applied. The objective of this phase was to accelerate the damage
which has initiated during the first phase and facilitate the identification of the major cracks in the material.
The number and length of the cracks observed at the surface of the specimen increased significantly in
this phase.

The addition of water in the second phase did not influence the rate at which the damage accumulated. The
water was added after the slabs were defined to have reached their fatigue lives and hence the performance
in second phase was not used in the analysis.

8.2.5 Deflection Instrumentation and Monitoring

Under the repeated application of rolling wheel loads, stabilised materials may be damaged due to fatigue. In
the field, fatigue damage to cemented materials has been assessed using variation in the measured surface
deflections with time and/or cumulated loading (Jameson, Sharp & Yeo 1992; Austroads 2008). The same
principle was applied in the laboratory wheel tracking test where a means of measurement slab deflection
was developed.

A set of displacement sensors was installed underneath the test slab to monitor the dynamic deflection of the
bottom of the slab under the rolling wheel load. The sensors were attached to the steel base plate as
illustrated in Figure 8.6.

Figure 8.6: Schematic view and photographs of the deflection monitoring system developed for wheel tracking
slab

b. Top view of the gauge locations

a. Schematic view of the deflection measurements system

c. Mould fitted with deflection sensors, before placement of LDPE


sheets
d. Deflection sensor with protection

Austroads 2020 | page 87


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

8.2.6 Fatigue Damage Characterisation

After wheel tracking testing, the fatigue damage was evaluated using several methods as follows:

Visual inspection

After trafficking the slabs were investigated to determine the cracking characteristics. Initially the slab surface
was visually inspected and the surface cracking pattern recorded, together with the condition of the bottom of
the slab. Figure 8.7 shows examples of surface cracking of an LBC slab and a HBC slab and also
differences in loose material at the bottom of the test slabs.

Figure 8.7: Examples of visual condition at the end of testing

LBC (1.5% cement) HBC (3% cement)

Material break down

After visual inspection the amount of loose material generated by the wheel loading was weighed as a
measure of fatigue distress. A vibrating table was used to shake the loose material from the test slab,
including a steel cradle to safely hold the slabs during this process (Figure 8.8). The material generating after
vibrating the slab for 2, 4, 6, 8, 10 and 15 minutes was collected in trays housed in the cradle and weighed.

Austroads 2020 | page 88


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 8.8: Vibrating table with support cradle

Cradle Slab on cradle Trays to collect lose material

Typical analysis of the deflection magnitude vs cycles

Under the repeated effect of the rolling wheel load the test slabs were gradually damaged. As the damage
was not visible until cracks propagate to the surface, as described in Section 8.2.5 slab deflections were
measured to quantify the changes of the test material modulus. The modulus changes provide an indication
of the fatigue damage in addition to the visual observations.

During the wheel trafficking, the sensors under the slab measured the deflections as a function of
time/loading. The deflections of centre sensor were recorded by an external data logger. Figure 8.9 is an
example of the measured elastic and permanent displacement of the bottom of the slab under repeated
loading passes. The peak deflections occurred when the tyre load was directly above the displacement
sensor and the troughs when the tyre was furthest away from the sensor (close to the edge of the mould).

Figure 8.9: Example of measured slab elastic and permanent deflections

Austroads 2020 | page 89


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

The deflection of the bottom of the slab in each cycle was determined from the difference in displacement
between the peak and trough.

The deflection per loading cycle were then plotted as a function of the number of cycles. Figure 8.10 shows
an example where the deflection in each cycle of loading initially increased significantly and then reached a
point after which the deflections tended to plateau.

Figure 8.10: Example of variation in deflections per loading cycles with number of loading cycles

It was postulated that the higher the deflection per cycle, the lower the slab modulus and that modulus
reduction is indicative of the fatigue damage under cyclic loading. Furthermore, it was hypothesised that
when a deflection plateau is reached the material has degraded to a fully micro-cracked state.

To compare the fatigue characteristics of different materials, further analysis of the deflection data was
performed to determine parameters that may be relevant to fatigue damage. The measured deflections were
divided into several segments (Figure 8.11) that may be related to the damage state.

Figure 8.11: Analysing peak to peak deflections from the centre sensor under the wheelpath

Austroads 2020 | page 90


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

The key parameters identified to evaluate were:


• the initial deflection (D0) measured at the start of the test
• the characteristics of the inflection point (N1, D1) when the deflection stabilises.

These parameters are represented schematically in Figure 8.11 and are described in Table 8.1.

Table 8.1: Parameters determined from the deflection magnitude signals

Parameter Notation Description


Initial deflection D0 The initial deflection is expected to be related to the initial stiffness of the
material. The stiffer the material is, the lower the deflection is expected to
be. The initial deflection is also a function of the modulus and thickness of
the support (LDPE sheets).
Number of cycles to N1 The number of loading cycles applied to the specimen from the start of the
inflection point test up until the inflection point.
Deflection reached at the D1 The point at which the rate of change in the deflection magnitude changes
inflection point from positive to zero or negative. This is typically displayed as a peak in the
deflection vs loading cycles plot. This inflection point shows a change in the
material’s behaviour, where the cracking mechanism has potentially
changed. In some cases, the variation nature of the slab deflections
necessitated some judgement in the selection of the inflection point.

8.3 Application of Test Method

8.3.1 Introduction

In this project, a total of 22 test slabs (varying from lightly bound to heavily bound materials) were prepared
and tested. Although the number of specimens for each cement content, type, and curing time is limited,
these test results allowed an appreciation of the usefulness of the test.

8.3.2 Test Materials

Two unbound granular materials were used in the testing:


• Material 1: Size 20 mm base quality (Class 2) crushed hornfels (particle distribution is shown in
Figure 8.12) (VicRoads 2017). This material was selected as its beam flexural fatigue characteristics
when stabilised with 3% cement, was measured in a previous Austroads project (Austroads 2012) and
before investigated in laboratory testing and accelerated loading (Austroads 2008).
• Material 2: Gravel from Barratta Creeks (particle distribution is shown in Figure 8.13). This material was
selected as it was the material used in the field trial B (Section 6.3).

Austroads 2020 | page 91


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 8.12: Particle size distribution of size 20 mm crushed hornfels

Figure 8.13: Particle size distribution of Barratta Creeks gravel

Table 8.2 shows the liquid limit, plastic limit, plasticity index (PI), maximum dry density (MDD) and optimum
moisture content (OMC) as determined using modified Proctor compaction test, for both the aforementioned
materials.

Table 8.2: Properties of selected materials

Liquid Plastic Plastic Maximum dry Optimum


Material limit limit index density (MDD)1 moisture content
(%) (%) (PI) (t/m3) (OMC) (%)
Untreated crushed hornfels 25 18 7 2.31 5.7
Untreated Barratta Creeks gravel 25 22 3 2.11 7.4

Austroads 2020 | page 92


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

When LBC materials are constructed in the field, it is common to use a slower-setting cementitious binder
such as type GB cement. However, it was determined from laboratory testing that the crushed hornfels
material stabilised with type GP cement after 14 days curing had a similar UCS to type GB cement after
28 days curing (Section 7.2). Based on this finding it was decided to use type GP cement to initially prepare
the stabilised crushed hornfels slabs as it expedited testing. Later type GB cement was used. The stabilised
Barratta Creek gravel was prepared using type GB cement, consistent with field trial B (Section 6.3).

Slabs have been prepared at various cement content between 1% and 3% and tested after a range of moist
curing times (Table 8.3).

Table 8.3: Slabs tested for fatigue damage characteristics

Cement content Curing time


Material Slab ID Cement type
(%) (days)
Stabilised 5555 GP 1.0 14
crushed hornfels
5568 GP 1.0 14
5688 GB 1.0 28
5694* GP 1.5 14
5541* GP 3.0 14
Stabilised 5922 GB 2.0 1
Barratta Creeks
6012 GB 2.0 1
gravel
5998 GB 2.0 28
6028 GB 2.0 28
5945 GB 2.0 90
5951 GB 2.0 90
5928 GB 4.0 1
6095 GB 4.0 28

*These specimens had a slurry layer applied to their surface before trafficking in the XL-WT device.

8.3.3 Compaction of Mixtures

The target compaction levels were 95% of the modified Proctor maximum dry densities (MDD). The MDD
values and optimum moisture contents (OMC) of the materials tested are reported in Table 8.4 and
Table 8.5.

Table 8.4: Maximum dry densities and optimum moisture contents of the stabilised crushed hornfels

1.0% GP cement* 1.5% GP cement 3.0% GP cement


MDD (t/m3) 2.26 2.27 2.32
OMC (%) 6.6 6.5 6.0

* 1% GB cement slabs were compacted to the same MDD and OMC values.

Table 8.5: Maximum dry densities and optimum moisture contents for the stabilised Barratta Creeks gravel

2% GB cement 4% GB cement
MDD (t/m3) 2.10 2.13
OMC (%) 8.3 8.0

Austroads 2020 | page 93


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Using the MDD, the slab target dry densities were calculated (Table 8.6) for use in compacting the test slabs.
The slab dry densities were measured post-trafficking by oven-drying a portion of each slab, coating it with
silicone and wax, and then tested as per a combination of the two methods: TMR Q306C 2016 and
AS/NZS 2891.9.1-2014. These measured dry density results are listed in Table 8.6. Generally, the measured
dry densities were within 2% of the target value.

Table 8.6: Density results for test slabs

Measured dry Target dry


Cement content
Host material Slab ID density density
and type
(t/m3) (t/m3)
Crushed hornfels 5555 1.0 GP 2.14 2.15
5568 1.0 GP 2.16 2.15
5688 1.0 GB 2.19 2.15
4833 1.5 GP 2.18 2.16
4866 1.5 GP N/A** 2.16
4932 1.5 GP 2.16 2.16
5029 1.5 GP 2.18 2.16
5383 1.5 GP 2.19 2.16
5492* 1.5 GP 2.17 2.16
5694* 1.5 GP 2.17 2.16
4811 3.0 GP 2.21 2.20
4876 3.0 GP 2.21 2.20
5469 3.0 GP 2.18 2.20
5541* 3.0 GP 2.17 2.20
Barratta Creeks 5922 2.0 GB 1.98 1.99
gravel
6012 2.0 GB 2.04 1.99
5998 2.0 GB 2.01 1.99
6028 2.0 GB 2.07 1.99
5945 2.0 GB 2.04 1.99
5951 2.0 GB 1.98 1.99
5928 4.0 GB 2.03 2.02
6095 4.0 GB 2.05 2.02

*These specimens had a slurry layer applied to their surface before trafficking in the XL-WT device.

**The density for slab 4866 could not be assessed.

8.3.4 Curing

The specimens were cured in ideal humidity-controlled conditions using a fog room (100% humidity) for their
relevant curing times before testing.

For type GP cement-treated specimens, 14 days curing was considered ideal for laboratory testing to enable
more specimen conditions to be tested during the laboratory program. This was the main reason for the
selection of type GP cement as it cures faster than GB cement that would be used in the field. Laboratory
investigation showed that the UCS of a type GP cement specimen after 14 days curing was the equivalent of
28 days curing of a specimen containing type GB cement.

As such, the average curing time for specimens treated with type GB cement were longer (often 28 or
90 days instead of 14 days) to allow sufficient strength gain to take place before fatigue testing.

Austroads 2020 | page 94


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

8.3.5 Results

Surface cracking

It was expected that a significant difference would be observed in the pattern of surface cracking with the
amount of cementitious binder. Figure 8.7 is an example of observed surface cracking variation with binder
content. It was concluded that due to the similarity of surface cracking observed in the XL-WT test
(e.g. Figure 8.7) irrespective of the binder content, surface cracking in the XL-WT was not indicative of the
different cracking characteristics of these materials in-service.

However, the visual inspection of slabs did tend to indicate a difference in the slab break down due to
loading. The bottom of the slabs with lower binder contents tended to break down more than those with
higher values, the latter tending to break into larger block cracks after trafficking.

Load-induced breakdown
For each of the test specimens, the mass of loose material collected in each tray (that is; left of the
wheelpath, right of the wheelpath, and underneath the wheelpath) was recorded after 2, 4, 6, 8, 10 and
15 minutes of slab vibration. The total mass collected from each slab after 15 minutes is given in Table 8.7,
with the mass recorded at each time interval provided in Appendix H.

For the Barratta Creeks gravel the results were too scattered to reach conclusions about the variation in the
mass of loosened with the strength of the stabilised material at the commencement of testing. In terms of the
stabilised hornfels, the 28-day result was unexpectedly high. In terms of the 14-day crushed hornfels results,
the mass for the loosened material tended to be greater for the slabs with lower cement content, i.e. the
degree of breakdown varied with the strength of the cemented material at the time of traffic loading.

Table 8.7: Vibrating table results

Cement content Curing time Total loose material


Host material Slab ID
and type (days) (kg)
Crushed hornfels 5555 1.0 GP 14 5.62
5568 1.0 GP 14 4.73
5688 1.0 GB 28 28.15
5694* 1.5 GP 14 1.90
5541* 3.0 GP 14 0.13
Barratta Creeks 5922 2.0 GB 1 0.12
gravel
6012 2.0 GB 1 0.15
5998 2.0 GB 28 0.07
6028 2.0 GB 28 5.92
5945 2.0 GB 90 0.17
5951 2.0 GB 90 2.65
5928 4.0 GB 1 0.06
6095 4.0 GB 28 0.27

* These specimens had a slurry layer applied to their surface before trafficking in the XL-WT device.

Note: Slabs 4833, 4866, 4932, 5029, 5383, 5492, 4811, 4876 and 5469 are not included here as they were unable to be
assessed on the vibrating table.

Slab deflections and estimated fatigue lives


As described in Section 8.2.6, the slab deflection data were analysed to provide insight into the fatigue
characteristics using the parameters listed in Table 8.1.
The results are reported in Appendix D and summarised in Table 8.8, Figure 8.14 and Figure 8.15. Table 8.9
also lists the mean results for each cement content and curing time.

Austroads 2020 | page 95


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Table 8.8: Summary of deflection data from trafficked slabs

Host Cement Cement Curing time D0 D1 Fatigue life


Slab ID
material type dosage (%) (days) (mm) (mm) N1 (x 103)
Crushed 5555 GP 1.0 14 0.33 0.48 17.6
hornfels
5568 GP 1.0 14 0.30 0.46 18.9
5688 GB 1.0 28 0.33 0.47 9.9
4833 GP 1.5 14 0.34 0.46 15.1
4866 GP 1.5 14 0.34 0.44 9.6
4932 GP 1.5 14 0.32 0.44 16.9
5029 GP 1.5 14 0.23 0.33 18.7
5383 GP 1.5 14 0.29 0.46 6.3
5492* GP 1.5 14 0.23 0.40 26.6
5694* GP 1.5 14 0.27 0.38 21.9
4811 GP 3.0 14 0.21 0.39 50.8
4876 GP 3.0 14 0.12 0.21 38.1
5469 GP 3.0 14 0.14 0.20 35.7
5541* GP 3.0 14 0.14 0.19 32.5
Barratta 5922 GB 2.0 1 0.24 0.27 89.2
Creeks
5998 GB 2.0 28 0.15 0.24 41.7
gravel
6028 GB 2.0 28 1.00 1.19 20.2
5945 GB 2.0 90 0.51 0.71 74.5
5951 GB 2.0 90 0.69 1.04 30.3
5928 GB 4.0 1 0.14 0.12 28.7
6095 GB 4.0 28 0.81 0.98 50.9

* These specimens had a slurry layer applied to their surface before trafficking in the XL-WT device

Table 8.9: Summary of average deflection data trafficked slabs of comparable cement contents and curing times

Cement Cement dosage Curing time D0 D1 Fatigue life


Host material
type (%) (days) (mm) (mm) N1 (x103)
Crushed GP 1.0 14 0.31 0.47 18.3
hornfels
GB 1.0 28 0.33 0.47 9.89
GP(1) 1.5 14 0.30 0.43 13.3
GP(2) 1.5 14 0.25 0.39 24.3
GP(3) 1.5 14 0.29 0.42 16.4
GP(1) 3.0 14 0.16 0.27 41.6
GP(2) 3.0 14 0.14 0.19 32.5
GP(3) 3.0 14 0.15 0.25 39.3
Barratta Creeks GB 2.0 1 0.24 0.27 89.2
gravel
GB 2.0 28 0.58 0.72 30.9
GB 2.0 90 0.60 0.88 52.4
GB 4.0 1 0.14 0.12 28.7
GB 4.0 28 0.81 0.98 50.9

1 Average of specimens of in this category without slurry.


2 Average of specimens of in this category with slurry applied.
3 Average of specimens in this category, including specimens with and without slurry applied.

Austroads 2020 | page 96


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 8.14: Overview of XL-WT testing for stabilised crushed hornfels

Figure 8.15: Overview of XL-WT testing for stabilised Barratta Creeks gravel

The initial deflection (D0) reflects the initial modulus of the slab. For the stabilised crushed hornfels, the
measured initial deflections generally decreased with cement content consistent with the expected increase
in modulus. As expected, slabs stabilised with 1.0% and 1.5% of cement had lower initial deflections than the
3% cement slabs. No similar trends were observed for the stabilised Barratta Creeks gravel.

The final deflection (D1) reflects at the inflection point. To some extent it may reflect the slab modulus in the
cracked state. For the stabilised crushed hornfels, three of the four 3% cement slabs had significantly lower
D1 values than the 1.0% and 1.5% cement slabs. This data suggests that the higher the pre-cracking
modulus, the higher the post-cracking modulus. No similar trends were observed for the stabilised Barratta
Creeks gravel.

The most interesting finding related to the relative fatigue lives. Concerning the stabilised hornfels,
Figure 8.16 shows the variation in the number of cycles to the inflection point with initial slab deflections. As
the higher the deflection the higher the tensile strain at the bottom of the stabilised layer, it is not surprising
that the number of cycles increases with reducing deflection.

Austroads 2020 | page 97


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 8.16: Stabilised crushed hornfels slab deflections

60000

50000 1.5% cement

3 % cement

40000

Number of
load cycles to 30000
inflection point

20000

10000

0
0.1 0.12 0.14 0.16 0.18 0.2 0.22 0.24 0.26 0.28 0.3 0.32 0.34 0.36

Initial slab vertical deflection (mm)

For the hornfels stabilised with 1.5% cement, the number of load cycles to the inflection point was
approximately half that for the 3% cement mixture, suggesting the fatigue life of 1.5% cement mixture was
half the 3% mixture. Such a difference in fatigue life for a 100 mm thick slab is in agreement with the
predictions of the Austroads (2017) cemented materials fatigue relationship assuming the material stabilised
with 1.5% cement has half the design modulus and half the design flexural strength of a material stabilised
with 3% cement. This provided assurance that the fatigue life of LBC could be approximated by extrapolation
of the Austroads method of predicting the fatigue life of cemented materials as used in Section 9.2.

The Barratta Creeks gravel results were limited and due to the variability of results the significance of binder
content could not be determined.

8.4 Summary of Findings


An extra-large wheel-tracking (XL-WT) test was developed to investigate cracking mechanisms of lightly
cemented material in the laboratory. A total of 22 test slabs (varying from LBC to cemented materials) were
prepared and tested under repeated loading.

The key findings were as follows:


• In terms of surface cracking that developed with loading cycles, there was not a significant difference in
the nature of the cracking with binder content. The block cracking that may occur in-service for
pavements with HBC bases could not be replicated in the wheel-tracking test. Most likely this relates to
the limited length, width and depth of the test slab compared to the dimensions in the roadbed.
• For the crushed hornfels tested after 14 days curing the mass of material loosened from the slab due to
loading tended to be greater for the slabs with lower cement contents i.e. the degree of breakdown varied
with cemented material strength at the time of traffic loading.
• The initial deflection (D0) reflected the initial modulus of the slab. For the stabilised crushed hornfels, the
measured initial deflections generally decreased with cement content, which is consistent with the
expected increase in modulus. As expected, slabs stabilised with 1.0% and 1.5% of cement had higher
initial deflections than the 3% cement slabs. A similar trend was not observed for the stabilised Barratta
Creeks gravel.

Austroads 2020 | page 98


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

• The final deflection (D1) is the deflection at the inflection point. To some extent it may reflect the slab
modulus in the cracked state. For the stabilised crushed hornfels, three of the four 3% cement slabs had
significantly lower D1 values than the 1.0% and 1.5% cement slabs. This data suggests that the higher
the pre-cracking modulus, the higher the post-cracking modulus. A similar trend was not observed for the
stabilised Barratta Creeks gravel.
• The most interesting finding related to the relative fatigue lives of the LBC and HBC materials. For the
hornfels stabilised with 1.5% cement, the number of load cycles to the deflection inflection point was
approximately half that for the 3% cement mixture. This finding was used in the development of the
design method (Section 9). However, the Barratta Creeks gravel results were limited; due to the variability
of the results the significance of binder content on fatigue could not be determined.

Austroads 2020 | page 99


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

9. Structural Design of Pavements Containing


Cracked Cemented Materials

9.1 Introduction
In this section, a framework is proposed for the structural design of new flexible pavements containing LBC for
incorporation into updates of the Austroads Guide to Pavement Technology Part 2: Pavement Structural Design
(Austroads 2017a) and to the design of rehabilitation treatments provided in the Austroads Guide to Pavement
Technology Part 5: Pavement Evaluation and Treatment Design (Austroads 2019a). This design framework has
been developed based on the use and performance of LBC for moderate-to-heavily trafficked roads. Its
applicability to the design of lightly trafficked roads needs consideration in the revision of Austroads guides.

As discussed below, the development of design procedures for LBC materials required a reconsideration of the
design procedures for HBC materials post-fatigue cracking. Hence, this section proposed changes to cemented
materials in the post-cracking phase of life irrespective of the type of cemented material pre-cracking.

Based on the project research, the following are the guiding principles used in developing this method:
• Like other cemented materials, LBC materials are susceptible to fatigue cracking, but the severity of
fatigue cracking is different from cement materials with binder contents of 3% or more.
• Pavements containing LBC bases are designed and constructed with limited strength such that the
allowable traffic loading before fatigue cracking is insignificant compared to the allowable traffic loading
post-cracking. As such, there is no need to consider the fatigue life of LBC in structural design. The
structural design only needs to consider pavement performance in the post-cracking phase of LBC life.
• In the design of new pavements containing HBC materials (Austroads 2017a), a post-cracking phase of
the design life can only be considered in the mechanistic-empirical thickness design calculations if
cracking from the fatigued HBC material does not reflect through to the surface. To reduce the risk of
reflective cracking the pavement should provide a minimum cover equivalent to 175 mm of asphalt over
the HBC material. To enable relaxation of this minimum cover requirement when LBC materials are used
as bases, it is essential that bases are designed and constructed such that in the post-cracking phase, to
limit the develop of block cracking and crocodile cracking from a state of fatigue-induced micro-cracking.
• The field performance of Queensland pavement containing LBC bases has indicated that when LBC
bases are constructed on low modulus supporting layers there is an increased risk that the fatigue of the
LBC may result in macro-cracking. Accordingly, when LBC layers are designed to inhibit macro-cracking,
minimum support conditions need to be provided.
• LBC materials are susceptible to fine shrinkage cracking and load-induced micro-cracking. To reduce the
likelihood that macro-cracking will develop, the loss load transfer across such fine cracking with traffic
loading needs to be minimised. This is similar to the need to maintain load transfer across undowelled
contraction joints in plain concrete pavements. In the micro-cracked state the extent of load transfer
depends on the properties of the granular host material, the support of the underlying materials and
importantly the thickness of the LBC layer thickness. As a consequence, the proposed design method
provides a minimum LBC layer thickness when LBC layers are designed to inhibit macro-cracking.
• Like cracked HBC materials, LBC materials are more rut-resistant than unbound granular materials
however no change to the design procedures is proposed to allow for this enhanced performance.
• South African experience suggests that weakly cemented materials are susceptible to crushing. This was
confirmed in Australian during accelerated pavement testing of cement-stabilised fly ash (Jameson et
al. 1995). To avoid this type of distress and to maintain load transfer across micro-cracks, limits are
placed on the quality granular materials used in LBC materials. Austroads (2019a) provides guidance on
materials suitable for cementitious stabilisation.

Austroads 2020 | page 100


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

• LBC materials may be used as subbase, in which case it is usually not be necessary to inhibit
macro-cracking. In such cases, the above-mentioned minimum thicknesses and support conditions are
not necessary.

9.2 Fatigue Life of LBC Layers


It was of interest to evaluate a design method which inhibits fatigue of LBC bases by adapting the
Austroads (2017) method used to determine the fatigue life of HBC materials. Based on the testing the
following characteristics were assumed for the LBC base:
• LBC pre-cracking design modulus of 2000 MPa, this being half the presumptive design modulus of
crushed rock subbase with 3–4% cement (Table 6.9 of Austroads 2017a). Modulus testing of the
stabilised hornfels supported the use of this modulus reduction (Section 7).
• LBC design flexural strength of 0.6 MPa, this being half the presumptive design modulus of crushed rock
subbase with 3–4% cement (Table 6.9 of Austroads 2017a). Strength testing of the stabilised hornfels
supported the use of one-third the strength (Section 7); however, as discussed below the wheel tracking
results suggest the use of a strength of one-half the value for a 3% cement mixture. Also, the findings of
Yeo (2011), shown in Figure 9.1, indicate the strength of 2% cement is about half the strength at
3–4% cement.
• Using equation 14 of Austroads (2017), the in-service fatigue constant (K) = 319.

Figure 9.1: Variation in 7-day flexural strength with cement content

Source: Yeo (2011).

The fatigue results in the laboratory wheel tracking testing (Section 8.3) tended to support this method of
predicting the fatigue life of LBC. In the wheel tracking test the fatigue life of LBC was about half the value for
cemented materials with 3% cement (Figure 8.16). The slab subjected to wheel tracking was 100 mm thick.
When the Austroads (2017) design method is applied to a 100 mm thick cemented material layer supported
by a modulus of 150 MPa, the predicted fatigue life of the 1.5% cement mixture is half the value for a mixture
with 3% cement using the following design inputs:
• 1.5% cement mixture: a pre-cracking design modulus of 2000 MPa and a design flexural strength of
0.6 MPa
• 3% cement mixture: a pre-cracking design modulus of 4000 MPa and a design flexural strength of
1.2 MPa.

Austroads 2020 | page 101


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

As this predicted reduction in fatigue life with reduction in cement content agrees with the measured values
(Section 8.3.5), it was concluded that the in-service fatigue life of LBC could be approximated by using the
Austroads (2017) method currently used for HBC but with a design modulus of 2000 MPa and a design
strength of 0.6 MPa.

Using the Austroads (2017) method, including the Appendix F traffic load distribution, the required LBC base
thicknesses for a range of traffic loadings were calculated using:
• a subgrade vertical design modulus of 50 MPa
• a range of granular subbase thicknesses resulting in top granular sublayer moduli of 150 MPa and
250 MPa supporting the LBC base.

As shown in Figure 9.2, for a design traffic loading of 106 ESA and 200 mm subbase thickness
(Etop = 150 MPa), a 400 mm thick LBC base would be required. This required thickness of the LBC base to
inhibit fatigue cracking is about 170 mm more than that required when only the post-cracking phase is
considered (Figure 9.10).

It is apparent that, in order to inhibit fatigue cracking of low strength LBC bases, very thick base layers would
be required and, as such, they would generally not be cost-effective. Accordingly, it is proposed that the
Austroads thickness design method considers only the post-cracking phase of LBC material life.

Figure 9.2: Example of LBC base thickness requirement to inhibit fatigue cracking

400

380

360

340
Etop = 150 MPa
320
Etop = 250 MPa
300
LBC
base
280
thickness
(mm) 260

240

220

200

180

160
10 100 1,000 10,000 100,000 1,000,000
Design traffic in ESA (DESA)

Austroads 2020 | page 102


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

9.3 Definition of LBC

9.3.1 Current Austroads Definitions

In providing a structural design method for LBC material, this material type needs to be defined.

Austroads (2017) defines lightly bound materials as follows, which provides a useful starting point for the
definition of LBC materials:
Lightly bound are granular materials to which moderate amounts of stabilising agent have
been added to improve modulus, and where an increase in tensile capacity may occur.
Lightly bound granular materials may exhibit behaviour between modified granular
materials (Section 6.3) and more heavily bound cemented materials. While the properties
of lightly bound materials and their response to loading is the subject of further research, it
is currently common practice to categorise materials with a 28-day UCS of 1.0 to 2.0 MPa
as lightly bound. At this stage this Part does not provide specific guidance for these lightly
bound materials.

Given that LBC materials are a particular type of cemented material, in defining LBC material consideration
was given to the current definition of cemented materials (Austroads 2017a):
Cemented materials are described as a combination of a cementitious binder, water and
granular material that are mixed and compacted in the early stages of the hydration
process to form a pavement layer that is subsequently cured. The cementitious binder may
consist of Portland cement, blended cement, lime, or other chemical binder and may
include one or more supplementary cementitious materials such as fly ash or ground
granulated blast furnace slag. The binder should be added in sufficient quantity to produce
a bound layer with significant tensile strength.

An important characteristic that is not defined by Austroads for cemented materials is the minimum 28-day
UCS. Australian road agency specifications for cemented materials were reviewed (Austroads 2013c) and
the results are shown in Table 9.1 with several updates. It is beyond the scope of this research project to
investigate appropriate minimum UCS values for HBC materials.

Table 9.1: Summary of road agency the minimum strength requirements of HBC materials

Road agency 7-day UCS 28-day UCS Test density


Roads and Maritime Services 3 MPa 100% standard MDD
New South Wales (Roads and
Maritime)
Roads Corporation, Victoria GP cement: 5 MPa – 100% modified MDD
(VicRoads)(1) GB cement: 3.5 MPa
Supplementary cementitious
blends: 3 MPa
Queensland Department of Category 1: 100% standard MDD
Transport and Main Roads 3.0 to 6.0 MPa unsoaked
(Queensland TMR)(2) Category 2:
2.0 to 4.0 MPa
Department of Planning, – GB cement: 4 MPa 96% modified MDD
Transport and Infrastructure,
South Australia (DPTI)

1 For plant mixed cement treated material used as subbase under thick asphalt use soaked values. For in situ
stabilisation unsoaked values are used for cementitious binder contents 3% or greater.
2 Updated in accordance with TMR (2018).

Source: Updated from Austroads (2013c).

Austroads 2020 | page 103


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Similarly, it was also useful to consider the current Austroads definition of modified granular materials
(Austroads 2017a):
Modified granular materials are granular materials to which small amounts of stabilising
agents have been added to improve modulus or to correct other deficiencies in properties
(e.g. by reducing plasticity) without causing a significant increase in tensile capacity
(i.e. producing a bound material). Modified materials have a maximum 28-day unconfined
compressive strength (UCS) of 1.0 MPa, tested after moist curing but without soaking at
100% standard maximum dry density and optimum moisture content.

9.3.2 TMR Definitions

In the design of new pavements TMR (2018) defines lightly bound materials as follows:
Lightly bound granular materials are typically specified to have a UCS between 1.0 and
2.0 MPa at 28 days when used in lightly bound base layers, and 1.0 to 2.0 MPa at seven
days when used in lightly bound improved layers.

Note that when lightly bound materials are used as improved layer/subbase in new pavements the required
range of strength effectively increases (TMR 2018), being approximately equivalent to a 28-day strength
between 1.5 and 2.5 MPa.

In the design rehabilitation treatments (TMR 2020), UCS limits used by TMR differ, namely a 7-day UCS
range of 1–2 MPa with a target 7-day UCS of 1.5 MPa. However, for slow setting binders a 28-day UCS
range of 1–2 MPa is stated.

9.3.3 Proposed Austroads Definition of Lightly Bound Cemented Materials

The proposed design produces are primarily based on the use of LBC materials by TMR. Consistent with
TMR (2018) and Austroads (2017), it is proposed that Austroads continue to use a 28-day UCS range of
1.0–2.0 MPa for LBC materials.

Accordingly, the proposed Austroads definition of LBC materials is:


Lightly bound cemented materials are described as a combination of a cementitious binder,
water and granular material that are mixed and compacted in the early stages of the
hydration process to form a pavement layer. The cementitious binder may consist of
Portland cement, blended cement, and lime and may include one or more supplementary
cementitious materials such as fly ash or ground granulated blast furnace slag. The binder
should be added in sufficient quantity to produce a weakly-bound layer with low tensile
strength. Where lightly bound cemented bases are designed to inhibit block cracking and
crocodile cracking, a maximum 28-day unconfined compressive strength (UCS) of 2.0 MPa
applies, and a minimum above 1.0 MPa determined after moist curing but without soaking
at 100% standard maximum dry density and optimum moisture content applies.

TMR has recommended that additional requirements be included in the definition, namely that there needs to
be a ‘steady’ increase in UCS between 7 and 28 days along with a minimum 7-day UCS of 1.0 MPa or a
degree of saturation requirement if the 7-day UCS < 1.0 MPa. These additional requirements need
consideration in the revision of the Austroads guides. TMR will also consider the use of modified compaction
in the future, possibly with adjusted UCS requirements in the longer term.

Austroads 2020 | page 104


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

9.3.4 Austroads Test Method for Preparation of UCS Test Specimens

A major issue that needs to be addressed with the inclusion of more detailed LBC design methods in
Austroads (2017 and 2019a) is consistency in the methods of preparing and curing UCS test specimens. The
differences in methods between laboratories include:
• The target compacted densities being related to both standard and modified Proctor maximum dry
densities.
• The target moisture contents during compaction being related to both standard and modified Proctor
optimum moisture content.
• The use of laboratory-mixed and field-mixed materials and consequently variable time delays between
mixing and test specimen compaction. Also, the field-mixed material can be at a lower moisture content
during laboratory compaction than the laboratory-mixed material.
• Soaking or not soaking before testing.

It is recommended that Austroads develop a test method related to the mixing, compacting and curing of
UCS test cylinders.

9.4 Elastic Characterisation for Pavement Design

9.4.1 Introduction

In undertaking mechanistic-empirical designs (Austroads 2017a), pavement materials are characterised by


their elastic parameters (modulus and Poisson’s ratios). A significant challenge in developing the elastic
characterisation of cracked cemented materials is the absence of a means of preparing laboratory test
specimens that reflect the structure of the cracked material in the roadbed. As a consequence, presumptive
moduli have been developed as described below.

As discussed in Section 3.2 TMR uses two design methods:


• TMR (2018) describes the method for designing lightly bound materials in new pavement constructions.
• TMR (2020) describes the method for designing lightly bound materials in rehabilitation treatments.

The elastic characterisation of the two methods differs. However, in both cases the elastic characterisation is
a variation on the Austroads (2017) elastic characterisation of unbound granular and modified materials.

For unbound granular materials, there is strong evidence that the modulus in the vertical direction is different
from that in the horizontal direction (i.e. they are anisotropic). In the mechanistic-empirical design procedure
the vertical modulus of unbound granular materials is taken as being equal to twice the horizontal modulus.
Conversely, Austroads (2017) assumes cemented materials are isotropic. In terms of cracked cemented
materials, no such evidence exists as to whether they are isotropic or anisotropic. It is proposed that all
cracked cemented materials be assumed to anisotropic consistent with the characterisation of unbound
granular and modified materials.

Similarly, there is an absence of data related to Poisson’s ratio of cracked cemented materials. Again, it is
proposed that all cracked cemented materials be assumed to have a value of 0.35 consistent with the
characterisation of unbound granular and modified materials.

The following provides guidance for the determination of the vertical design modulus of cracked cemented
materials, that is LBC and HBC materials.

Austroads 2020 | page 105


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

9.4.2 Factors Affecting Cracked Cemented Material Moduli

To develop a method to determine the design modulus of cracked cemented materials, the principal factors
affecting modulus needed to be considered. It is not unreasonable to assume the modulus of cemented
materials in post-cracking phases may depend on:
• the pre-cracking design modulus of the cemented material, reflecting the quality of the host materials,
cementitious type, and content and in situ density
• applied stress level as affected by the thicknesses and moduli of overlying bound (e.g. asphalt) layers
• the underlying support to cement material layer and the cemented material thickness.

Effect of pre-cracking design modulus

In the pre-cracking phase of life, the design modulus is influenced by host materials, cementitious type and
content, and in situ density. In the absence of a laboratory method to prepare cement material test
specimens in an appropriate cracked state hinders consideration of whether the moduli in the cracked state
should vary or not with these same factors.

From the laboratory wheel-tracking of slabs with varying cement contents (Section 8), the vertical slab
deflections (D1) at the point of inflection may provide insight as to whether or not the cracked moduli are
related to the pre-cracking design modulus. The slab deflections of the stabilised crushed hornfels in the
cracked state were generally lower the higher the pre-cracked modulus. Assuming the measured slab
deflection is inversely related to the slab modulus, this suggests the post-cracking moduli are related to the
pre-cracking moduli.

However, the back-calculated moduli of 1% and 4% cement-stabilised materials at the end of accelerated
loading reported by Alabaster et al. (2013) (Figure 4.11), indicated the opposite; namely the cracked
cemented material modulus was not related to the pre-cracking modulus. Although the initial back-calculated
modulus of 1% cement mixture was about half the 4% mixture value, the moduli were similar at the end of
loading.

Given this limited and conflicting information, it is proposed to adopt the simpler approach that the
post-cracking modulus does not vary with the pre-cracking modulus. That is, the same values apply for
cracked LBC materials and cracked HBC materials.

Modulus stress-dependency

For unbound granular and modified granular materials, the maximum vertical modulus that a material can
develop depends on the applied stress. The applied stress reduces with the thickness and modulus of
overlying bound materials (Austroads 2017a). Consequently, granular design moduli reduce as the thickness
and modulus of overlying materials increase.

Conversely, for bound materials (e.g. asphalt, cemented materials), in Austroads (2017) no allowance is
made for the variation in design modulus with the level of applied stress.

Concerning the stress dependency of cracked cemented materials, Austroads (2008) reported the findings of
accelerated pavement loading of a 3% cement-treated crushed rock subbase supported by a sand subbase.
Figure 9.3 shows the variation in back-calculated moduli with the number of cycles of loading.

Austroads 2020 | page 106


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 9.3: Modulus reduction of 3% cement-treated crushed rock subbase on a sand subbase

Source: Austroads (2008).

Using the data from experiment 3310, Figure 9.4 shows how the back-calculated modulus of the
3% cement-treated crushed rock at the end of loading varied with the applied FWD plate contact stress. This
material would be categorised as HBC material. At the end of trafficking when the HBC material was
micro-cracked (Austroads 2008) the back-calculated modulus tended to increase with load (Figure 9.4), yet
before accelerated loading there was no statistically significant variation in back-calculated modulus with
stress.

Figure 9.4: Experiment 3310 back-calculated HBC materials at the end of the loading

2000

1800

1600

1400 y = 0.982x + 468


R² = 0.47

1200
Cemented
materials
modulus 1000
(MPa)

800

600

400
500 550 600 650 700 750 800 850 900 950 1000

FWD plate stress (kPa)

Austroads 2020 | page 107


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

This data indicates that cracked HBC material modulus increases in modulus by about 50% when the
applied stress is doubled. This modulus increase with stress is about half the increase of unbound granular
base moduli (Figure 6.3 of Austroads 2017a). In the absence of other data, it is proposed that the cracked
HBC material bases be reduced for the effect of overlaying bound material thickness and modulus at half the
rate at which granular moduli reduce (Figure 9.5). Although, there is no supporting data concerning LBC
material, it is proposed this reduction also applies to this material as it is consistent with the assumption that
the characterisation in the cracked state is similar irrespective difference in pre-cracking characterisation.

For unbound granular materials and modified granular materials, the Austroads (2017) sub-layering process
in part addresses the effect of stress reductions with depth below the surface. However, as this sub-layering
process also addresses the effect of the underlying support, the significance of the modulus
stress-dependency in the sub-layering rules is uncertain. It would also be possible to allow for the
stress-dependency within cracked cemented material layers. However, a simple set of design rules suitable
for any layer depth below the surface would be complex. Hence it is not proposed to sublayer cemented
materials in mechanistic-empirical design calculations.

Figure 9.5: Proposed cracked cemented material modulus reduction factors with overlaying thickness of bound
materials

Effect of underlying supporting layers

For unbound granular and modified granular materials, the maximum vertical modulus that a material can
develop is dependent on the modulus of the underlying supporting layer (Austroads 2017a). Conversely, for
bound materials (e.g. asphalt, cemented materials), no allowance is made for the variation in modulus with
the underlying support.

Concerning cracked cemented materials, Figure 9.6 shows the variation of back-calculated moduli of a
150 mm thick 3% cement-treated crushed rock subbase under accelerated loading (Jameson, Sharp &
Yeo 1992). This HBC material was placed directly on a clay subgrade with an estimated in situ CBR of about
6%. The back-calculated HBC vertical modulus reduced to about 250 MPa when cracking was observed on
surface.

Austroads 2020 | page 108


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 9.6: Modulus reduction of 3% cement-treated crushed rock subbase on clay subgrade

Source: Adapted from Jameson, Sharp and Yeo (1992).

However, in a later research project (Austroads 2008) when the 3% cement-treated crushed rock was
supported by a sand subbase (Figure 9.4) the back-calculated HBC vertical modulus reduced to about
800 MPa or more at the end of trafficking. In this case, surface cracking was not observed but cracking was
detected when the pavement was excavated at the completion of loading.

Based on this Australian accelerated loading trial data, the modulus of cracked cemented materials needs to
vary with the support provided by the underlying layers: the lower the modulus of the supporting layers, the
lower the cracked cemented material modulus.

However, the results of New Zealand accelerated pavement testing (Section 4.4.3) on a subgrade with an in
situ CBR of about 4% were not consistent with this conclusion. As seen from Figure 4.11, under a 60 kN dual
wheel load the back-calculated moduli of 1% and 4% cement-treated crushed rock reduced to 1400 MPa
despite the low support to the cemented material layers. Note that the strength of the subgrade was similar
to the Australian data used to derived Figure 9.6, however the cracked cemented materials at the end of
trafficking differed markedly. Again, in this research surface cracking was not observed.

It was not possible to resolve the findings from these three accelerated loading trials other than to note that
surfacing cracking was observed at the completion of the initial Australian research (Figure 9.6) but not at the
end of the New Zealand testing (Figure 4.11) or later Australian research (Figure 9.4).

It is proposed to base the Austroads procedures on the initial Australian modulus reductions (Figure 9.6) as
this resulted in a more conservative design moduli of cemented materials in the cracked state.

Note that the effect of the supporting layers on limiting modulus is different from that currently used for
unbound granular materials and modified granular materials. For instance, the 250 MPa back-calculated
modulus for the cracked cemented material (Figure 9.6) is about four times the modulus of the supporting
subgrade. Using Austroads (2017), the top modulus of granular material on this subgrade would be about
140 MPa. The Austroads (2017) sub-layering method for unbound granular materials would be overly
conservative for cracked cemented materials.

The most conservative design rule from these three accelerated loading trials is that the modulus to which
cemented materials reduce is limited to four times the modulus of the supporting layer.

Austroads 2020 | page 109


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 9.7 shows the variation of vertical design modulus for cracked cemented materials with the modulus
of the supporting layer. This method is based on the assumption that the modulus of cracked cemented
material is limited to a maximum of four times the modulus of the supporting layer.

As there was no available data about high modulus (> 150 MPa) supporting layers, no increase in modulus
above a supporting value of 150 MPa is proposed. This chart includes a maximum vertical design modulus of
600 MPa as discussed below.

Figure 9.7: Limits on design moduli cracked cemented materials


650

600

550

500

450
Cracked cemented
material design
400
modulus
(MPa)
350

300

250

200

150
30 50 70 90 110 130 150 170
Supporting layer vertical design modulus (MPa)

Maximum modulus of cracked cemented materials

The maximum modulus a cracked cemented material base can develop occurs when the base modulus is:
• not limited by the modulus of the supporting layer and
• the applied traffic-induced stresses are not limited by the thickness and modulus of overlying bound
materials (e.g. asphalt).

This is consistent with the characterisation of unbound granular materials and modified granular materials.

Concerning LBC bases, modulus data in the cracked state was limited to the back-calculated modulus of the
250 mm thick LBC base in field trial B (Section 6.3). After one year of trafficking the mean back-calculated
modulus reduced to about 770 MPa (Table 6.22), with the moduli at three of the ten sites being below
600 MPa. At that time the mean back-calculated modulus of the underlying granular subbase was low,
80 MPa. Given the very low fatigue lives LBC bases it is not expected that the back-calculated modulus
would decrease significantly with additional loading.

As mentioned above, in Austroads (2008) research when the 3% cement-treated crushed rock was
supported by a sand subbase (Figure 9.4) the back-calculated vertical moduli of the cracked cement material
at the end of trafficking exceeded 800 MPa.

Austroads 2020 | page 110


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

As discussed in Section 4.5, as part of a review of the in-service performance of New Zealand stabilised
pavements, back-calculated moduli of two pavements stabilised with 3% cement were estimated
(Figure 4.13). The weaker areas of the test pavements had an isotropic modulus of 500 MPa which did not
change over 10 years. Converting this isotropic modulus to the Austroads cross-anisotropic characterisation,
this equates to a vertical modulus of about 550 MPa.

As seen from the above, information about the modulus of cracked cemented materials was limited to
back-calculated moduli rather than laboratory testing as a method of preparing cracked test specimens has
yet to be developed. It should be noted that back-calculated moduli are currently not used in the design of
new pavements (Austroads 2017a) so care is needed in determining design moduli from back-calculated
values.

The New Zealand research (Figure 4.11) indicated that cemented materials after accelerated loading had
about twice the back-calculated modulus of an unbound granular base. Consequently, it seems reasonable
to assume that the maximum design modulus that a cracked cemented material base can develop should
exceed the presumptive vertical design moduli for unbound granular bases produced from the same source.

Considering the above information and TMR design method (Section 3.2.2) it is proposed that the maximum
presumptive vertical design modulus of 600 MPa be used for cracked LBC and HBC materials.

For unbound granular materials and modified granular materials, a vertical design modulus greater than
600 MPa may be derived from repeated load triaxial testing (Section 6.2.3 of Austroads 2017a). The use of
higher design moduli for these materials than cracked cemented materials is counter-intuitive.

Consequently, it is proposed that the maximum modulus of cracked cemented materials be the greater of
600 MPa and design modulus obtained by direct measurement of a cement-modified granular material
produced from the same host material, limited to a maximum of 1000 MPa.

Note that the direct measurement of resilient modulus of modified granular materials is recommended for
review to ensure the condition at the time of testing reflects in-service values including load-induced damage.

As described above, research data existed (Jameson, Sharp & Yeo 1992; Austroads 2008) about the
modulus of cement treated crushed rock subbases manufactured for Victorian host materials which most
likely had a laboratory soaked CBR of 30% or more. The characteristics of such high-quality subbases was
considered in the proposed maximum modulus of 600 MPa. However, there is lack of information about the
moduli of LBC materials manufactured from lower quality subbase materials. Until such information is
obtained, it is proposed that the design modulus of LBC materials manufactured from subbase materials with
a laboratory soaked CBR less 30% be limited to a maximum modulus of 210 MPa. This value is the
maximum modulus for subbase layers currently used by TMR (Section 3.2.1).

9.4.3 Proposed Austroads Elastic Characterisation

Procedures have yet to be developed to prepare laboratory test specimens in a cracked state that reflects
the state in situ. Consequently, a presumptive elastic characterisation is proposed as described above. As
discussed in Section 9.4.2, it is proposed that this elastic characterisation be used for cracked LBC and
cracked HBC materials.

For mechanistic-empirical design, the following elastic characterisation is proposed for LBC and HBC
material in the post-fatigue cracking phase of life:
• For subbases manufactured from granular materials with a laboratory soaked CBR less than 30%, the
vertical design modulus is the minimum of 210 MPa (Section 9.4.2) and the value obtained from
Figure 9.7 reflecting the effect of the supporting layer.

Austroads 2020 | page 111


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

• For bases and subbases manufactured from granular materials with a laboratory soaked CBR of 30% or
more, the vertical design modulus is the minimum of:
– the value obtained from Figure 9.7 reflecting the effect of the supporting layer,
– the greater of 600 MPa and the design modulus obtained by direct measurement (Austroads 2017a) of
an unbound granular material or cement-modified granular material produced from the same host
material, limited to a maximum of 1000 MPa, and
– For LBC layers covered by bound materials, the Table 9.2 maximum vertical design moduli which
reflects the modulus stress-dependency of these materials.
• For LBC and HBC layers in the post-fatigue cracking phase of life, the above design modulus are applied
to the entire layer thickness: the layers are not sublayered.
• Cracked cemented materials are cross-anisotropic (with a degree of anisotropy of 2) and have a
Poisson’s ratio of 0.35.

Table 9.2: Suggested maximum vertical modulus of cracked cemented material bases and subbases

Thickness of Modulus of the overlying bound material (MPa)


overlying
bound material
1000 2000 3000 4000 5000
(mm)
40 600 600 600 600 600
75 600 600 590 580 570
100 600 570 550 530 520
125 570 530 510 490 470
150 540 500 470 440 430
175 510 460 430 430 430
200 480 430 430 430 430
225 460 430 430 430 430
≥250 430 430 430 430 430

Note: These maxima apply to LBC bases and subbases, manufactured using unbound granular material with a
laboratory soaked CBR greater than 30%. For LBC materials manufactured using unbound granular material with a
laboratory soaked CBR less than 30%, a maximum modulus of 210 MPa applies regardless of the thickness and
modulus of the overlaying bound materials.

9.5 Inhibiting Macro-cracking of LBC bases

9.5.1 Introduction

The guidance below is limited to inhibiting macro-cracking when LBC materials are used as base as it is
unlikely LBC subbases will need to be designed to inhibit macro-cracking given likely overlaying surfacing
and base thicknesses.

As discussed in Section 9.2, due to the use of low cement contents LBC materials have low strength and
modulus and as a consequence, very thick LBC layers are required to inhibit fatigue cracking. Hence, an
Austroads design method to inhibit fatigue cracking of LBC materials is not proposed.

The design method is based on the premise that LBC layers are susceptible to fine shrinkage cracking and
load-induced micro-cracking. The thickness design method needs to inhibit the micro-cracking of LBC bases
with thin bituminous surfacings leading to surface macro-cracking, resulting in additional road maintenance
and rehabilitation.

Austroads 2020 | page 112


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

There is empirical evidence (Section 5) that two key characteristics are important to inhibiting block cracking
and crocodile cracking:
• the support provided to the LBC base by the underlying layers
• the LBC base thickness.

Both these characteristics are considered in the TMR design process which requires a minimum LBC base
thickness of 200 mm and a minimum design modulus of 150 MPa for the layer supporting the LBC.

9.5.2 Minimum Support to LBC Base

As discussed in Section 5.3.3, the analysis of selected Queensland pavements tended to suggest the extent
of surface longitudinal and transverse cracking reduced with increasing granular subbase thickness.

The current TMR (2018) design method recognises the importance of the support provided to the LBC base
as follows:
For lightly bound base pavements, the base is typically supported on a subbase with
thickness of at least 150 mm and which achieves a vertical design modulus of at least
150 MPa at the top of the subbase (determined using the procedures detailed in
Sections 8.2.2 and 8.2.3 of AGPT02).

This may be achieved by increasing the thickness of the subbase, and/or including
additional select fill or unbound granular material beneath the subbase.

It is proposed the Austroads method for LBC bases include the TMR minimum support requirements.

9.5.3 Minimum Base Thicknesses

In addition to the minimum support provided to LBC bases, consideration was given as to whether additional
guidance is required to inhibit macro-cracking. In particular whether the required LBC base thickness needs
to increase with design traffic loading. The need for minimum thickness can be appreciated considering the
very limited fatigue life of LBC bases less than 200 mm thick as illustrated in the example designs in
Figure 9.2. There is a risk that low base thicknesses will result in block and crocodile cracking due to one
overloaded truck and this may result in undesirable cracking characteristics.

In the post-cracking phase of LBC life, it was considered the formation of LBC macro-cracking from
micro-cracking is influenced by the maintenance of load transfer across micro-cracking. The provision of
minimum support under the LBC base is important in this respect as it influences the load-induced
movement of cracks. It is postulated that by reducing crack movement the attrition of fines at cracks will be
reduced and hence inhibit micro-cracks widening into macro-cracks.

Of particular importance to the design method is whether the extent of crack movement needs to reduce as
the number of load repetitions increase. It was considered the Austroads thickness design procedures for
concrete pavements provided insight in this respect.

In the design of concrete pavement, erosion refers to the load-induced pumping of the subgrade/subbase
underlying the concrete base arising from repeated deflections at joints and planned cracks. The resulting
voids lead to faulting at transverse contraction joints under the slab. The Austroads thickness design
procedure for erosion was developed from finite element modelling and the performance of AASHO Road
Test concrete pavements (Vorobieff & Hodgkinson 2001). The amount of erosion was correlated with the
predicted deflection of the corner of the slab, the pressure on the subbase/subgrade, shape of the surface
deflection bowl and the load transfer across the joint (Figure 9.8).

Austroads 2020 | page 113


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 9.8: Concept of load transfer at joints in concrete pavements

Source: Vorobieff and Hodgkinson (2001).

Using the Austroads erosion damage method for a jointed plain concrete pavement, Figure 9.9 shows the
influence of concrete base thickness in erosion life for an effective subgrade CBR of 10%. Increasing the
base thickness:
• reduces the stress on the top of the subbase/subgrade
• reduces deflections at the slab corners
• lengthens the surface deflection bowl, thereby reducing the rate of which energy is applied to the
subbase/subgrade
• increases the contact area at the contraction joint, thereby increasing load transfer across the joint.

Figure 9.9: Example variation in allowable traffic loading in terms of erosion with concrete base thickness

260

250

240 SAST 60 kN
SADT 80 kN
230
TADT 170 kN

220 TRDT 210 kN

Concrete
210
base
thickness 200
(mm)
190

180

170

160

150
1.0E+05 1.0E+06 1.0E+07 1.0E+08
Allowable axle group repetitions

In order to maintain load transfer across joints, the extent of joint movement clearly needs to reduce as the
number of load repetitions increases. This concept was considered to apply to LBC bases designed to inhibit
the development of macro-cracking from micro-cracking: namely, the load-induced movement of
micro-cracks needs to reduce as the number of load repetitions increase.

Austroads 2020 | page 114


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

The proposed LBC base design method addresses this issue by providing a minimum base thickness that
varies with the design traffic and the modulus of the underlying supporting layers. Figure 9.10 shows the
proposed design chart for minimum base thicknesses with thin bituminous surfacings. For pavements where
the LBC base is covered with asphalt less than 175 mm thick, the minimum LBC base thickness obtained
from Figure 9.10 may be reduced by the thickness of structural asphalt subject to a minimum LBC thickness
of 200 mm. These Figure 9.10 minimum layer thicknesses do not apply to LBC bases covered by 175 mm or
more thickness of structural asphalt as such asphalt thickness inhibit reflection cracking.

Figure 9.10: Minimum LBC base thicknesses for pavements with thin bituminous surfacings

300

290 Etop 150 MPa

280 Etop 200 MPa

270 Etop 250 MPa

260

Minimum
250
LBC
thickness
240
(mm)
230

220

210

200

190
1.0E+05 1.0E+06 1.0E+07 1.0E+08
Design traffic in ESA (DESA)

To develop this method, a pavement structure needed to be identified with the required ability to inhibit
development of macro-cracking from micro-cracking. The Queensland pavements reviewed in Section 5
were used to guide this reference case together with the TMR (2018) design method. A 20-year design traffic
loading of 5 x 106 ESA is a mid-range value for the Queensland application of LBC bases. With the provision
of a vertical design modulus of 150 MPa at the top of the subbase, a LBC base thickness of 250 mm was
selected which is within the 200–300 mm range commonly used by TMR. This was the assumed reference
case with the desired structural characteristics to inhibit macro-cracking.

The next issue to address was how to vary the LBC base thickness with the number of load repetitions, such
that the daily damage to micro-cracking was the same as the reference case. It was postulated that if the
early-life daily fatigue damage was the same as the reference case, the damage to the micro-cracking would
be the same. This seemed reasonable given that fatigue damage causes micro-cracking. As seen from
Figure 9.2 when the LBC base thickness increases from 250 mm to 285 mm, the fatigue life increases by a
factor of 10. Using this thickness variation, it was assumed that to limit the daily damage to the
micro-cracking the LBC base, the base thickness needs to increase by 35 mm for each 10-fold increase in
design traffic loading. Figure 9.11 shows the resulting variation in LBC base thickness when supported by a
granular subbase with design modulus of 150 MPa at the top of the subbase consistent with the reference
case.

Austroads 2020 | page 115


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 9.11: Minimum LBC base thickness for granular subbase with E = 150 MPa

300

290

280

270

260

Minimum 250
LBC
thickness 240
(mm)
230

220

210

200

190
1.0E+05 1.0E+06 1.0E+07 1.0E+08
Design traffic in ESA (DESA)

Also, it is reasonable to assume the higher the support provided by the granular subbase the lower the
movement of the micro-crack which enables a reduction in LBC base thickness. From Figure 9.2, it can be
seen that for a given fatigue life increasing the top modulus from 150 MPa to 250 MPa reduces the required
LBC base thickness by 25 mm. Using this base thickness adjustment, Figure 9.10 shows the proposed
design chart for minimum LBC thickness. Note the thicknesses were all derived based on the assumption
that if the daily fatigue damage to the LBC base was the same as the reference case, the extent of
macro-cracking would be the same as the reference case.

9.5.4 Design Moduli of Underlying Granular Layers

In relation to the moduli of granular materials under the LBC, typically these granular materials would be of
subbase quality but may extend to a normal standard base material. The modulus of the top granular
sublayer under the LBC base is limited in a similar manner to described in Section 6.2.3 of Austroads (2017).
Assuming the LBC base has a vertical design modulus of 600 MPa, Figure 9.12 shows the maximum vertical
modulus of normal standard base material with LBC base thickness increases. In relation to selecting the
maximum moduli of subbase materials: for high-quality crushed rock subbases, which have a laboratory
soaked CBR greater than 30%, an assigned maximum modulus of the lesser of the value from Figure 9.12
and 210 MPa may be used, otherwise a maximum value of 150 MPa may be assigned.

Austroads 2020 | page 116


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure 9.12: Suggested maximum vertical modulus of top sublayer of normal standard base material

260

250

240

230

220

Maximum vertical 210


modulus of
200
top sublayer
(MPa) 190

180

170

160

150

140
200 210 220 230 240 250 260 270 280 290 300
LCM base thickness (mm)

9.6 Use of Mechanistic – Empirical Design Method


Flexible pavements containing LBC materials are proposed to be structurally designed using the
mechanistic-empirical method (Austroads 2017a).

The proposed methods of elastic characterisation of LBC and HBC materials in the post-cracking phase of
life are described above. In addition, guidance is provided on the proposed minimum LBC layer thicknesses
to inhibit macro-cracking and determined on design moduli of underlying granular materials.

Thereafter, the pavement thickness design steps are the same as currently provided in Austroads (2019a)
for the design of cement materials in the post-cracking phase of life. This includes the use of the
mechanistic-empirical design method to determine the pavement composition required to inhibit permanent
deformation and asphalt fatigue cracking.

9.7 Proposed Changes to the Guide to Pavement Technology


The following parts of the Guide to Pavement Technology will need to be amended if the proposed design
methods are adopted:
• Part 2 – Pavement structural design
• Part 4D – Stabilised materials
• Part 5 – Pavement evaluation and treatment design.

Given the use of LBC materials to date has been largely limited to Queensland, an option for consideration
by Austroads to allow the national use of the guidance provided in this report for an agreed period of time
prior to incorporation in these Austroads guides.

Austroads 2020 | page 117


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

10. Summary and Conclusions


Treating granular materials with cementitious and bituminous binders is becoming commonplace as reserves
of high-quality manufactured granular materials are exhausted or hauling such materials over long distances
is cost-prohibitive. Improved knowledge in this area would directly contribute to the government road reform
agenda concerning greater certainty on access decisions for high-productivity freight vehicles and in relation
to the heavy vehicle charging and investment reforms. In particular, road agencies and industry identified a
need for improved guidance for the design of pavements using materials with low binder contents. The
addition of a small amount of cementitious binder to non-standard granular materials may result in a
fit-for-purpose base or subbase at a significantly lower cost than the use of crushed rock complying with
standard specifications. Such lightly bound cemented (LBC) materials have particularly use in pavement
rehabilitation and heavy patching.

Queensland Department of Transport and Main Roads (TMR) has adopted cement stabilisation of pavement
materials on an intermittent basis for many years. In recent years, the use of LBC bases has been more
frequent in some regions, particularly North Queensland. These LBC bases are generally used to rehabilitate
existing pavements using in situ stabilisation; however, plant-mixed and paver-laid LBC bases have been
used for new pavements. The LBC bases are generally surfaced with a sprayed seal or thin asphalt
surfacing. These bases have been designed and constructed to inhibit the development of block cracking
and crocodile fatigue cracking that can develop with heavily bound cemented (HBC) bases. In the design of
rehabilitation treatments, the binder content is selected such that the UCS after 7 days curing is 1–2 MPa.

The objectives of this research project were to investigate the in-service performance of LBC pavements, to
improve understanding of the characteristics of LBC bases that inhibit the development of block cracking,
and to develop pavement structural design methods.

Consistent with Austroads (2017), the proposed definition of LBC materials is:
Lightly bound cemented materials are described as a combination of a cementitious binder,
water and granular material that are mixed and compacted in the early stages of the
hydration process to form a pavement layer. The cementitious binder may consist of
Portland cement, blended cement, and lime and may include one or more supplementary
cementitious materials such as fly ash or ground granulated blast furnace slag. The binder
should be added in sufficient quantity to produce a weakly-bound layer with low tensile
strength. Where lightly bound cemented bases are designed to inhibit block cracking and
crocodile cracking, a maximum 28-day unconfined compressive strength (UCS) of 2.0 MPa
applies, and a minimum above 1.0 MPa determined after moist curing but without soaking
at 100% standard maximum dry density and optimum moisture content applies.

TMR has recommended that additional requirements be included in the definition, namely that there needs to
be a ‘steady’ increase in UCS between 7 and 28 days along with a minimum 7-day UCS of 1.0 MPa or a
degree of saturation requirement if the 7-day UCS < 1.0 MPa. These additional requirements need
consideration in the revision of the Austroads guides. TMR will also consider the use of modified compaction
in the future, possibly with adjusted UCS requirements in the longer term.

The performance of LBC bases in Queensland was investigated and the following conclusions were drawn:
• From the field inspections conducted it can be concluded that plant mixed lightly cemented bases have
performed well. As such materials are placed as part of new construction, it is reasonable to assume that
the underlying support to the LBC base to be expected in new pavement design is a significant factor in
the good performance of these projects. However, in some cases the thin LBC bases were constructed
on cemented subbases and transverse cracking was observed on the surface, most likely related to
shrinkage of the cemented material subbases.

Austroads 2020 | page 118


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

• The in situ stabilised bases showed a much wider range of performance. All LBC bases that were well
supported demonstrated good performance, as did some pavements where the underlying support could
be considered as inadequate from a structural design point of view.
• However, some in situ stabilised pavements exhibited very poor performance with the formation of major
block cracking evident. Additionally, some sections were found to have been extensively patched,
indicating that significant distress had been evident and the pavement had needed treatment to restore
serviceability. It has been postulated that the application of excessive binder has resulted in a continuous
length of one project exhibiting extreme block cracking in all wheelpaths, but evidence in support of this
could not be found. Inadequate subbase support to the LBC was suggested as the most likely cause of
poor performance. Current TMR (2018) design procedures reflect this experience as now the LBC bases
needs to be supported by a granular subbase with a minimum design modulus of 150 MPa.
• Several sites were cored to improve understanding of the required properties of good- and
poor-performing LBC bases. For sites with good performance with no block cracking, it was not possible
to extract intact cores because they disintegrated, either due to traffic loading or during coring. Two sites
where cores could be extracted were in poor-performing areas with visible block cracking. These findings
are consistent with TMR design concepts for LBC bases, namely to design and construct a low-strength
material that develops fine micro-cracking with sufficient base thickness and subbase support to limit the
extent that micro-cracking leads to macro-cracking.
• Long lengths of LBC bases showed little or no signs of block cracking or crocodile cracking after up to
10 years of trafficking.

To develop a design method for pavements containing LBC, a method first needed to be developed to
determine the design modulus of LBC. To assist the development, two field trials were constructed and
monitored for 12 months to enable the change in moduli with curing and early-life trafficking to be
investigated. Field trial A was located on the Bruce Highway at Collinson’s Lagoon. The 250 mm thick LBC
base was constructed by in situ stabilisation using 2.5% type GB cement. This binder content was selected
without the guidance of UCS testing of laboratory-manufactured cylinders. From deflection testing it was
concluded that the cement-stabilised layer had very high in situ modulus (> 5000 MPa) well above that
expected for a LBC base. It was doubtful that the moduli at this site were representative of a LBC base and
so the results were not used in developing the proposed design method.

Field trial B was located on the Bruce Highway near Barratta Creeks. The 250 mm LBC base was
constructed by in situ stabilisation of sandy gravel using 2% type GB cement; the UCS after 28 days was
determined as being approximately 2 MPa. From measured deflections the mean back-calculated modulus in
the outer wheelpath decreased from 1150 MPa after 90 days of trafficking to 770 MPa after one year. This
information was used to select the design modulus of LBC bases for use in the Austroads design method.

An important component in the development of a design method was gaining an understanding of the fatigue
characteristics of LBC compared to HBC materials. In terms of HBC materials, which commonly have binder
contents of 3% or more, flexural strength and flexural modulus can be used to predict fatigue life.
Accordingly, the strength and modulus of LBC materials and HBC materials were measured and the relative
fatigue lives determined. The key findings were as follows:
• One of the two LBC materials was Barratta Creeks gravel stabilised with 2% type GB cement. The
measured 28-day UCS of this material was 2.5 MPa and hence did not comply with the proposed
Austroads upper limit for LBC material. This UCS result was contrary to the testing carried out by TMR
and the contractor prior to and during construction (Section 6.3.4). This difference in UCS could not be
explained by the difference in the densities of the test specimens.
• The other LBC material was crushed hornfels stabilised with 1.5% type GB cement. It had a 28-day UCS
of 2.5 MPa and hence did not complied with the proposed Austroads upper limit for LBC material.
• In terms of the relative moduli of the hornfels stabilised with 1.5% type GB cement compared to the
3% mixture, both the indirect tensile and flexural modulus testing indicated that the modulus of
1.5% mixture was approximately half of the 3% mixture value. However, for the Barratta Creeks gravel
the flexural modulus of the 2% mixture was unexpectedly high (about 14 000 MPa), a value similar to the
4% mixture.

Austroads 2020 | page 119


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

• In terms of the relative strength of the hornfels stabilised with 1.5% type GB cement compared to a
3% mixture, both the indirect tensile and flexural strength testing indicated that the modulus of the
1.5% mixture was approximately one-third of the 3% mixture value. However, for the Barratta Creeks
gravel the flexural strength of the 2% mixture was unexpectedly high (about 1 MPa after 90 days), a value
only slightly lower than the 4% mixture.

An extra-large wheel-tracking (XL-WT) laboratory test was developed to investigate the cracking
mechanisms of LBC materials compared to HBC materials. A total of 22 test slabs were prepared and tested
under repeated rolling wheel loading. The main findings were as follows:
• In terms of surface cracking that developed with loading cycles, there was not a significant difference in
the nature of the cracking with binder content. The block cracking that may occur in pavements with HBC
bases in-service could not be replicated in the wheel-tracking test. This most likely relates to the limited
length, width and depth of the test slabs compared to the dimensions in the roadbed.
• The mass of material loosened from the bottom of the test slab under loading tended to be greater for the
slabs with lower cement contents and shorter curing times. That is, the degree of breakdown varied with
cemented material strength at the time of traffic loading. This supports the view that early opening of
cemented materials to traffic may reduce the risk of macro-cracking.
• For the stabilised crushed hornfels, three of the four 3% cement slabs had significantly lower slab
deflections at the end of testing than the 1.0% and 1.5% cement slabs. This suggests that the higher the
pre-cracking modulus, the higher the post-cracking modulus. A similar trend was not observed for the
stabilised Barratta Creeks gravel.
• The most interesting finding related to the relative fatigue lives of the LBC and HBC materials. For the
hornfels stabilised with 1.5% cement the number of load cycles to the deflection inflection point was
approximately half that for the 3% cement mixture. This finding was used in the development of the
design method. However, the Barratta Creeks gravel results were limited and, due to the variability of
results, the significance of binder content in relation to fatigue could not be determined.

From this laboratory testing, a procedure was developed to predict the fatigue of LBC materials, this being an
extrapolation of the current method for HBC materials. It was concluded that LBC fatigue lives are so low that
excessive LBC thicknesses would be required to design pavements to inhibit fatigue cracking. Hence, and
consistent with the TMR design method, it is proposed that the Austroads design method only consider the
post-cracking phase of LBC life.

The research highlighted the need for a national test method for mixing, compaction and curing of UCS
specimens to improve consistency of UCS results across jurisdictions. Given that the use of LBC materials
may increase with the improved design method, there is an increased need for this method lest HBC bases
are used as LBC bases.

A structural design method was developed for pavements containing LBC materials and HBC materials in
the post-fatigue cracking phase of life, including:
• A new method for the elastic characterisation applicable to LBC materials and HBC materials in the
fatigue cracked state, including methods to vary the design modulus with the design modulus of the layer
supporting the cracked material and with the thickness and modulus of overlying bound materials.
• Design charts to select LBC base thicknesses to inhibit the development of block cracking and crocodile
cracking, with the minimum thickness varying with design traffic loading, the support provided by the layer
under the LBC base and thickness of asphalt surfacing (if any).

This framework is applicable to materials suitable for cement stabilisation as defined in Austroads (2019b).

This design framework was developed based considering the use and performance of LBC for
moderate-to-heavily trafficked roads. Its applicability to the design of lightly trafficked roads needs
consideration in the revision of Austroads guides.

Austroads 2020 | page 120


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Austroads (2019a) provides guidance about surfacing treatments to inhibiting reflection cracking due to
cracking of cemented materials. TMR (2018) requires the use of SAM/SAMI over LBC bases. This
information can be utilised in the revision of the Austroads guides.

Given the use of LBC materials to date has been largely limited to Queensland, an option for consideration
by Austroads to allow the national use of the guidance provided in this report for an agreed period of time
prior to incorporation in Austroads guides.

Austroads 2020 | page 121


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

References
Alabaster, D, Patrick, J, Arampamoorthy, H & Gonzalez, A 2013, The design of stabilised pavements in New
Zealand, research report 498, NZ Transport Agency, Wellington, New Zealand.

Arnold, G, Morkel, C & van der Westhuizen, G 2012, Development of tensile fatigue criteria for bound
materials, research report 463, NZ Transport Agency, Wellington, New Zealand.

Austroads 2008, Fatigue performance of cemented materials under accelerated loading: influence of vertical
loading on the performance of unbound and cemented materials, AP-T102-08, Austroads, Sydney, NSW.

Austroads 2009, Guide to pavement technology part 3: pavement surfacings, AGPT03-09, Austroads,
Sydney, NSW.

Austroads 2012, Preliminary investigation of the influence of micro-cracking on fatigue life of cemented
materials, AP-T198-12, Austroads, Sydney, NSW.

Austroads 2013a, Improved rut resistance characterisation of granular bases: manufacture and
commissioning of a wheel-tracking device, AP-T239-13, Austroads, Sydney, NSW.

Austroads 2013b, Development of a wheel tracking test for rut resistance characterisation on unbound
granular materials, AP-T240-13, Austroads, Sydney, NSW.

Austroads 2013c, Review of definition of modified granular materials and bound materials, AP-R434-13,
Austroads, Sydney, NSW.

Austroads 2014, Cemented materials characterisation: final report, AP-R462-14, Austroads, Sydney, NSW.

Austroads 2017a, Guide to pavement technology part 2: pavement structural design, AGPT02-17,
Austroads, Sydney, NSW.

Austroads 2017b, Guide to pavement technology part 4F: bituminous binders, AGPT04F-17, Austroads,
Sydney, NSW.

Austroads 2019a, Guide to pavement technology part 5: pavement evaluation and treatment design,
AGPT05-19, Austroads, Sydney, NSW.

Austroads 2019b Guide to pavement technology part 4d: stabilised materials , AGPT05-19, Austroads,
Sydney, NSW.

Chakrabarti, S, Kodikara, J & Pardo, L 2001, ‘An overview of stabilisation methods and performance of local
government roads in Australia’, International symposium on subgrade stabilisation and in situ pavement
recycling using cement, 1st, 2001, Salamanca, Spain, Conference Organising Committee, Salamanca,
Spain, 16 pp.

Design Flow 2019, Polyethylene sheets, web page, Design Flow, Carrum Downs, Vic, viewed 24 June 2020,
<https://www.matrixpanels.com.au/polyethylene_sheets.html>.

Dunlop, RJ, Moss, PJ & Dodd, TAH 1975, ‘Prediction of cracking in soil-cement’, Australia-New Zealand
conference on geomechanics, 2nd, 1975, Brisbane, Qld, Institution of Engineers, Barton, ACT, pp. 120–4.

Freeman, TJ & Little, DN 1998, Develop maintenance strategy selection procedures for pavements
incorporating semi-rigid or chemically stabilized layers, FHWA/TX-99/17222-2, Texas Transportation
Institute, Austin, TX, USA.

Austroads 2020 | page 122


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

George, KP 2002, Minimising cracking in cement-treated materials for improved performance, research and
development bulletin RD123, Portland Cement Association, Skokie, IL, USA.

Gray, W, Frobel, T, Browne, A, Salt, G & Stevens, D 2011, Characterisation and use of stabilised
basecourse materials in transportation projects in New Zealand, research report 461, NZ Transport
Agency, Wellington, New Zealand.

Jameson, GW, Sharp, KG & Yeo, R 1992, Cement treated crushed rock pavement fatigue under accelerated
loading: the Mulgrave (Victoria) ALF trial, 1989/1991, ARR 229, Australian Road Research Board,
Vermont South, Vic.

Jameson, GW, Dash, DM, Tharan, Y & Vertessy, NJ 1995, Performance of deep-lift in situ pavement
recycling under accelerated loading: the Cooma ALF trial, 1994. APRG report 11, Australian Road
Research Board, Vermont South, Vic.

Course Hero 2020, Katholieke Universiteit Leuven n.d., CR (chloroprene rubber _ neoprene), webpage,
Katholieke Universiteit Leuven, Leuven, Belgium, viewed 24 June 2020,
<https://www.coursehero.com/file/18781343/CR-Chloroprene-Rubber-Neoprene/>.

Lam, T & Bryant, P 2014, ‘CMB pavement trial: Frank Mallan Drive Lytton (constructed August 2014)’,
Queensland Department of Transport and Main Roads, Brisbane, Qld.

Queensland Department of Transport and Main Roads 1990, Pavement design manual, 2nd edn, TMR,
Brisbane, Qld.

Queensland Department of Transport and Main Roads 2018, Pavement design supplement: Supplement to
Part 2 Pavement Structural Design of the Austroads Guide to Pavement Technology, TMR, Brisbane,
Qld.

Queensland Department of Transport and Main Roads 2019, Transport and Main Roads specifications
MRTS10: plant-mixed lightly bound pavements, TMR, Brisbane, Qld.

Queensland Department of Transport and Main Roads 2020, Pavement rehabilitation manual, TMR,
Brisbane, Qld.

Roads and Maritime Services 2018, Construction of unbound and modified pavement course, QA-R71, RMS,
Sydney, NSW.

Scullion, T 2002, ‘Pre-cracking of soil-cement bases to reduce reflection cracking: field investigation’,
Transportation Research Record, no. 1787, pp. 22-30.

Sebesta, S 2005, Continued evaluation of microcracking in Texas, FHWA/TX-06/0-4502-2, Texas


Transportation Institute, College Station, TX, USA.

Smith, G & Caltabiano, M 1987, ‘The use of low shrinkage CTB in district 14 (metropolitan south)’, Main
Roads Department of Metropolitan Divisional technical symposium, Brisbane, Qld, Queensland
Department of Transport and Main Roads, Brisbane, Qld.

VicRoads 2008, In situ stabilisation of pavements with cementitious binders, section 307, VicRoads, Kew,
Vic.

VicRoads 2011, ‘Potential effects of a low percentage of GP cement in subbase crushed rock’, TR 211,
VicRoads, Kew, Vic.

VicRoads 2017, Registration of crushed rock mixes, RC 500.02, VicRoads, Kew, Vic.

Austroads 2020 | page 123


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Vorobieff, G & Hodgkinson, J 2001, ‘Technical basis of the 1992 guide procedures for design of rigid
pavements’, in G Jameson (ed) 2003, Technical basis of Austroads Guide to Pavement Technology Part
2: Pavement Structural Design, ARR 384, ARRB Transport Research, Vermont South, Vic.

Vorobieff, G, Walter, P & Anwar, S 2010, ‘Long-term performance of recycled granular materials in southern
NSW’, ARRB conference, 24th, 2010, Melbourne, Victoria, ARRB Group, Vermont South, Vic, 20 pp.

White, G 2006, ‘Laboratory characterisation of cementitiously stabilised pavement materials’, MEng thesis,
University of New South Wales, Canberra, ACT.

Wu, Z & Gaspard, K 2013, Minimizing shrinkage cracking in cement-stabilized bases through the use of
micro-cracking, research project capsule 12-3P, Louisiana Transportation Research Center, Baton
Rouge, LA, USA.

Yeo, YS 2011, Characterisation of cement-treated crushed rock basecourse for Western Australian roads,
PhD thesis, Curtin University, Perth, WA.

Zhang, J, Hou, D & Han, Y 2012, ‘Micromechanical modelling on autogenous and drying shrinkages of
concrete’, Construction and Building Materials, vol. 29, no. 3, pp. 230–40.

Australian/New Zealand Standards

AS 1012.11-2000 (R2014), Methods of testing concrete: determination of the modulus of rupture.

AS 5101.4-2008, Methods for preparation and testing of stabilized materials: unconfined compressive
strength of compacted materials.

AS 2891.13.1-2013, Methods of sampling and testing asphalt: determination of the resilient modulus of
asphalt: indirect tensile method.

AS/NZS 2891.9.1-2014, Methods of sampling and testing asphalt: determination of bulk density of
compacted asphalt: waxing procedure.

Austroads Test Methods

AGPT-T600-18, Flexural beam test methods for cemented materials.

AGPT-T054-15, Determination of permanent deformation characteristics of unbound granular materials by


the wheel-tracking test.

Roads and Maritime Services Test Method

T147 2012, Working time for road construction materials (blended in the laboratory with slow setting
binders).

Queensland Department of Transport and Main Roads Test Method

Q306C 2016, Compacted density of asphalt: silicone sealed.

Austroads 2020 | page 124


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Appendix A Performance Review of Selected


Queensland Pavements: Inspections

A.1 Introduction
As described in Section 5.2, selected sites in Queensland were inspected as part of the project. This
Appendix contained notes and photographs related to the roads inspected. The text is based on an
unpublished project Working Paper dated October 2016 prepared by Phil Hunt of Road Engineering Services
and Dr Michael Moffatt of ARRB as part of Austroads project TT1897.

Field inspection sites were (Figure A 1):


• Site 1: Darling Downs District Cunningham Highway (Ipswich – Warwick): Job 110/17B/302
• Site 2 and 2A: Darling Downs District Cunningham Highway (Inglewood – Goondiwindi): Job 122/17D/807
• Site 3: Far North Queensland District Bruce Highway (Ingham – Innisfail): Job 216/10N/66Z. This site is
not reported below as the treatment was found on inspection to be foamed bitumen stabilisation.
• Site 4: Far North Queensland District Palmerston Highway (Innisfail – Ravenshoe): Job 264/21A/56Z.
This site is not reported below as it was a re-stabilisation of previously prematurely cracked previous
stabilisation treatment.
• Site 5: Far North Queensland District Malanda – Atherton Road: Job 119/645/802 & 803
• Site 6: North West District Flinders Highway (Hughenden – Richmond): Job 55/14C/304
• Site 7: North West District Flinders Highway (Hughenden – Richmond): Job 257/14C/67H
• Site 8: Mackay/Whitsunday District Bruce Highway (Bowen – Ayr): Job 125/10K/59
• Site 9: Mackay/Whitsunday District Bruce Highway (Bowen – Ayr (10K)): Job 61/10N/508
• Site 10: Northern Queensland District North Townsville Road (Woolcock St: Townsville Port Road): Job
150/10M/17
• Site 11: Northern Queensland District Douglas – Garbutt Road (Duckworth St): Job 150/10L/38
• Site 12: Northern Queensland District Bruce Highway (Townsville – Ingham): Job 117/10M/819
• Site 13: Northern Queensland District Bruce Highway (Townsville – Ingham): Job 61/10M/32
• Site 14 & 14A: Fitzroy District Bundaberg – Miriam Vale Road: Job 83/179/23
• Site 15: Fitzroy District Western Yeppoon – Emu Park Road: Job 258/197/201
• Site 16: Fitzroy District Dawson Highway (Gladstone – Biloela): Job 27/46A/304
• Site 17 & 17A: Fitzroy District Dawson Highway (Biloela – Banana): Job 8/46B/303.

Austroads 2020 | page 125


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure A 1: Locations of field monitoring sites

Source: http://ontheworldmap.com/australia/state/queensland/.

Prior to field visitation, asset management data, construction drawings, UCS records (if available), were
collated for each project.

The site inspections were undertaken between June and September 2016.

A.2 Site 1: Cunningham Highway (Ipswich – Warwick)

A.2.1 Overview

The site inspected was Cunningham Highway (National Highway), between Ipswich and Warwick. 2.9 km
between chainages 85.8 and 88.7 km, in Darling Downs District. The road is located in undulating terrain
with good pavement drainage.

The pavement was widening and rehabilitated in 2008 with the following pavement structure:
• 250 mm LBC
• 300 mm of granular pavement
• subgrade design CBR of 15% due to use of selected fill on embankments and sound cuttings.

The existing lanes were overlaid with 250 mm thickness of Type 2 base and then in situ stabilised.

The LBC was produced using 1.5% cementitious binder, the binder being a 60% type GP/40% fly ash blend.

Design testing provided the following results:


• 1.5% binder: 7-day UCS of 0.6 to 1.5 MPa and 28-day UCS of 1.6–1.9 MPa
• 2.0% binder: 7-day UCS of 1.6 to 2.0 MPa and 28-day UCS of 1.9–2.5 MPa
• 2.5% binder: 7-day UCS of 2.2 to 2.6 MPa and 28-day UCS of 2.8–3.5 MPa.

Austroads 2020 | page 126


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

A.2.2 Performance Summary

The 2.9 km study section was eight year old when inspected in 2016 and had experienced a cumulative
traffic loading 4.4 x 106 ESA.

Fifty per cent of the total through-lane wheelpaths (4 x 2.9 km) either have been patched or currently show a
rut/shove or fatigue cracking (Figure A 2).

During construction UCS testing was taken at two locations, both locations subsequently needed to be
patched due premature distress. The UCS results of field-mixed and laboratory compacted specimens were:
• Chainage 86.79: 7-day UCS of 1.3 & 1.6 MPa
• Chainage 88.315: 7-day UCS of 0.6 MPa.

In general, the patch repairs have consisted of 150–200 mm deep, 2% type GB re-modification (adding
additional cement using a stabiliser/reclaimer). The repairs are performing very well, with the exception of
one section that is showing signs of crocodile cracking. As shown in photograph on the immediate right of
Figure A 2.

Due to the extensive distress very early in life, this pavement was clearly a poor performer.

Austroads 2020 | page 127


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure A 2: Site 1 Cunningham Highway

Austroads 2020 | page 128


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

A.3 Site 2 and 2A: Cunningham Highway (Inglewood – Goondiwindi)

A.3.1 Overview

The two sites inspected were on the Cunningham Highway (Inglewood – Goondiwindi) between chainages
85.4 and 87.0 km (Figure A 3) in Darling Downs District. The terrain is very flat grades. The presence of
longitudinal cracking possible indicates expansive clay subgrades in parts of the project.

In 2007 the pavement was rehabilitated with the resulting pavement compositions:
• Site 2 85.4–86.6 km: 250 mm LBC on remaining 50 mm of unbound granular material. Total pavement
depth 300 mm.
• Site 2A 86.6–87.0 km: 250 mm LBC on 325 mm of unbound granular material. Total pavement depth
575 mm.

The surfacing comprised a size 10 mm primerseal followed by size 14 mm final size using PMB Olexobit
S0.3 binder.

The subgrade materials included clays to sandy clays with estimated low to medium CBRs in the ranged of
3–5% in weakest areas to CBR of 7–9% elsewhere.

The existing pavement materials were in situ stabilised to a depth of 250 mm using 2% type GB cement. The
7-day UCS at 2% binder was 1.3 MPa which was within the 7-day target range of 1.0 to 1.5 MPa.

Figure A 3: Cunningham Highway (Inglewood – Goondiwindi)

A.3.2 Performance Summary

The 1.6 km study section was 9 years old when inspected in 2016 and had cumulative traffic loading of
4.4 x 106 ESA.

Site 2 with 300 mm total pavement depth has performed to a fair standard (Figure A 4), whilst Site 2A with
575 mm total pavement depth has performed to a good standard.

Overall 4.5% of the wheelpaths exhibits either crocodile cracking, transverse cracking or patching.

There is some longitudinal cracking on the shoulder edges, indicating the likelihood of expansive subgrades.

Austroads 2020 | page 129


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure A 4: Site 2 inspection 2016

Austroads 2020 | page 130


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

A.4 Sites 5 and 5A: Malanda – Atherton Road

A.4.1 Overview

Sites 5 and 5A were located on the Malanda-Atherton Road between chainages 13.5–14.4 km and
15.5–16.3 km, in far North Queensland District. The terrain is flat to very flat with good table drains
(Figure A 5).

Figure A 5: Site 5 Malanda – Atherton Road

The pavement was rehabilitated in 2009 using LBC, with the resulting pavement compositions:
• Site 5 13.5–14.4 km: 200 mm thick overlay of plant-mixed LBC on 240 mm unbound granular on a red
clay subgrade with a design CBR of 6% (soaked). Total pavement depth 440 mm; however, this could be
greater where the old pavement existed.
• Site 5A 15.5–16.3 km: 200 mm thick overlay of plant-mixed LBC on 160 mm unbound granular pavement
on a red clay subgrade with a design CBR of 6% (soaked). Total pavement depth 360 mm.

The surfacing was a size 16/7 sprayed seal.

The LBC was produced in a pugmill using 1.5% type GB cement and placed with a paver. Design or
construction UCS results were not obtained in the investigation.

At the time the pavements were inspected after 7 years of trafficking, the cumulative traffic loading was
8.4 x 105 ESA and 5.2 x 105 ESA for Sites 5 and 5A respectively.

A.4.2 Performance Summary

After seven years of trafficking both sites are in sound condition with no signs of any pavement distress. It is
possible that a reseal was placed before the inspection was masking cracking and other repairs, however the
pavement maintenance costs were low and there was no surface deformation.

Austroads 2020 | page 131


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

A.5 Site 6: Flinders Highway (Hughenden – Richmond)

A.5.1 Overview

Site 6 was located on the Flinders Highway (Hughenden- Richmond) between chainages 70.5 km and
74.5 km in North West District. The terrain was flat to slightly undulating terrain with good table drains. The
road formation is 0.5–1.0 m above natural subgrade, which is an expansive black soil (Figure A 6).

Figure A 6: Site 6 Flinders Highway

The pavement was rehabilitated in 2005 and the resulting pavement composition was:
• 250 mm thickness LBC achieved by in situ stabilised of existing pavement after it was topped up with
50–100 mm Type 3 granular base.
• 100–250 mm thickness of sandy loam subbase
• black clay subgrade.

The LBC was produced using 5% of a 50% lime/50% fly ash blend.

Before construction UCS testing was undertaken using both type GB cement and 50% lime/50% fly ash
blend. The target 7-day UCS for the type GB cement was 1.0 to 1.5 MPa, whereas for 50% lime/50% fly ash
the target was a 28-day UCS of 1.0 to 1.5 MPa. From the results (Figure A 7) it was decided to 5% of
50% lime/50% fly ash.

Austroads 2020 | page 132


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure A 7: Site 6 UCS testing

A.5.2 Performance Summary

This pavement was about 11 years old when inspected 2016 and had experienced a cumulative traffic
loading was 9.3 x 105 ESA.

The pavement was in good condition (Figure A 8). However, a reseal six months before the inspection may
have masked some defects and/or old maintenance repairs. There is also evidence that a 1 m wide
shoulder/OWP geotextile reinforced seal placed a year after construction (2006) was used to control
environmental longitudinal cracking on a reactive clay subgrade (Figure A 9).

Austroads 2020 | page 133


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure A 8: Site 6 inspection

Austroads 2020 | page 134


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure A 9: Geotextile reinforced seal to inhibit cracking

A.6 Sites 7: Flinders Highway (Hughenden – Richmond)

A.6.1 Overview

Site 7 was located on the Flinders Highway (Hughenden-Richmond) between chainages 91.6 km and
102.3 km in North West District. The terrain is flat to slightly undulating, with the pavement on an elevated
formation with good table drains. The subgrade was an expansive black soil, which had a design CBR of 3%
or less.

Figure A 10: Sites 7 Flinders Highway

In 2012 after a major rain event damaged the existing granular pavement, the pavement was rehabilitated by
in situ stabilisation to a depth of 250 mm. The resulting pavement composition was:
• 250 mm thickness of LBC
• 75 mm thickness of granular subbase
• black clay subgrade with a design CBR of 3%.

Austroads 2020 | page 135


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

The LBC was produced using 2% type GB cement. UCS test results were not available for this site.

A.6.2 Performance Summary

This pavement was four years old when inspected in 2016 and had experienced a cumulative traffic loading
was 4.8 x 105 ESA.

The performance of the four-year-old LBC pavement was considered generally good. However, the 2015
network road condition data indicated average rutting between 7 to 10 mm, with maximum rutting up to
15 mm occurring. This high amount of rutting would be considered poor performance for a 4-year-old
pavement. This depth of rutting was not readily observed onsite.

There was one very poor performing area between 101.6 km to 102.9 km (16% of job length) exhibiting block
cracking in all wheelpaths (Figure A 11). It was suspected that a major cement overdose or double cement
application after some type of initial construction failure was experienced.

Figure A 11: Site 7 block cracking

A.7 Site 8: Bruce Highway (Bowen – Ayr)

A.7.1 Overview

Site 8 was located on the Bruce Highway (Bowen – Ayr) between chainages 43.7 km and 45.6 km in
Mackay/Whitsunday District. The terrain is flat to very flat (Figure A 12) with a bridge crossing of the Elliot
River at chainage 44.6–44.75 km.

The existing pavement was reconstructed in 1999, with the following composition:
• 200 mm thickness of LBC
• 125 mm thickness of Type 2.3 granular subbase
• 175 mm thickness of Type 2.5 granular subbase
• clay subgrade with a design CBR of 4%.

The design traffic loading was 5.2 x 106 ESA.

No details could be found concerning whether the LBC was plant-mixed or produced by in situ stabilisation,
but given the year of construction was 1999 it seemed more likely to be in situ stabilisation. The LBC was
manufactured using 1.5% type GB cement. UCS results were not available for this site.

Austroads 2020 | page 136


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure A 12: Site 8 Bruce Highway Bowen to Ayr

A.7.2 Performance Summary

The 1.9 km study section was 17 years old when inspected and was in very good condition (Figure A 13)
with generally no deflects after 5.2 x106 ESA of traffic loading. In 2010, the pavement was resealed using a
polymer modified binder. It is not known whether there was surface cracking before the reseal.

Figure A 13: Site 8 inspection

A.8 Site 9: Bruce Highway (Bowen – Ayr)

A.8.1 Overview

Site 9 was located on the Bruce Highway (Bowen – Ayr) between chainages 75.2 km and 78.9 km in
Mackay/Whitsunday District. The terrain was flat with good surface drainage (Figure A 14). This site was not
identified before commencing the field inspections but was noticed when travelling to Site 8 because of the
substantial extent of block cracking.

The pavement was rehabilitated in 2008. Limited information was available about the pavement composition
other than it includes a 200 mm thickness of LBC on an unbound granular subbase (depth unconfirmed) on a
clay subgrade with an estimated subgrade CBR in the range of 3–4%.

Austroads 2020 | page 137


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

No details could be found about the cement content and type used to produce the LBC.

Figure A 14: Site 9 Bruce Highway Bowen to Ayr

A.8.2 Performance Summary

At the time of inspection in 2016, the pavement was eight years old and had experienced a cumulative traffic
loading of about 1.8 x106 ESA. This pavement was severely cracked with block cracking on a total 23% of
the length (Figure A 15).

As details of the cement content and type are not known it is uncertain whether this is an LBC. Due to this
doubt its performance was not considered in the final analysis.

Austroads 2020 | page 138


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure A 15: Site 9 inspection

Austroads 2020 | page 139


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

A.9 Site 10: North Townsville Road

A.9.1 Overview

Site 10 was the Townsville North Road (Woolcock St: Townsville Port Road) between chainages 0.8 km and
4.7 km in Northern Queensland District. This road was originally part of the old Bruce Highway through
Townsville. It is now known as the North Townsville Road or colloquially as Woolcock St: Townsville Port Rd.
This asphalt surfaced urban road is a divided road with two lanes in each direction (Figure A 16).

Figure A 16: Site 10 North Townsville Road

The terrain is flat to slightly undulating, with the pavement on elevated select fill (CBR = 10) and good
drainage. The underlying subgrade is low strength clay (possibly originally swamps) with a design CBR of
1%.

The pavement was constructed in 1998 with the following composition:


• 45 mm thickness of dense graded asphalt (DGA)
• 125 mm thickness of LBC base
• 405 mm thickness of Type 2.3 granular subbase
• 250 mm select fill with a design CBR of 10%
• varying depth of earthworks (typically 1–1.5 m)
• natural subgrade, design CBR of 1%.

The LBC was plant-mixed using 1.5% type GB cement and placed with a paver. TMR testing indicated the
resulting 7-day UCS was 0.50–0.75 MPa, a low strength compared to other in situ stabilised projects.

In 2008 the pavement was overlaid with 45 mm of DGA providing a total of 90 mm of DGA.

A.9.2 Performance Summary

The 3.9 km long site was 18 years old when inspected and had experienced a cumulative traffic loading of
5.2 x 106 ESA. The pavement had no visual deflects in 2016 and had performed very well (Figure A 17).

Austroads 2020 | page 140


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure A 17: Site 10 inspection

A.10 Site 11: Douglas – Garbutt Road (Duckworth St)

A.10.1 Overview

Site 11 was the Douglass – Garbutt Road (Duckworth St) between chainages 4.1 km and 5.4 km in Northern
Queensland District. This road is an asphalt surfaced urban arterial in Townsville (Figure A 18). The road
was originally part of the old Bruce Highway.

Austroads 2020 | page 141


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure A 18: Site 11 Douglass – Garbutt Road

The pavement was constructed in 1998 with the following composition:


• 45 mm thickness of dense graded asphalt (DGA)
• 125 mm thickness of LBC base
• 290 mm thickness of cement-treated subbase
• 200 mm thickness of working platform
• subgrade with a design CBR of 3%.

The design traffic loading was 6 x 106 ESA.

The LBC base was plant-mixed using 1.5% type GB cement and placed with a paver. UCS test results were
not available for this project.

In 2011 after 13 years in-service the original 45 mm DGA surfacing layer was removed and replaced with the
same material 45 mm DGA.

A.10.2 Performance Summary

The 1.3 km study section was 18 years old when inspected in 2016 at which time it had a cumulative traffic
loading experience of 7.6 x 106 ESA, exceeding the design traffic loading of 6 x 106 ESA.

No pavement defects were observed, the pavement was in sound condition (Figure A 19).

Austroads 2020 | page 142


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure A 19: Site 11 inspection

A.11 Site 12: Bruce Highway (Townsville – Ingham)

A.11.1 Overview

Site 12 was the Bruce Highway (Townsville to Ingram) between chainages 42.8 km and 50.6 km in Northern
Queensland District. The site had a flat terrain with good table drains (Figure A 20).

Figure A 20: Site 12 Bruce Highway Townsville to Ingram

The pavement was constructed in 2000 with the following pavement composition:
• 200 mm thickness of LBC base
• 250 mm thickness of cement-treated subbase (TMR category 1)
• subgrade with a design CBR of 4%.

The design traffic loading was 5.9 x 106 ESA.

Austroads 2020 | page 143


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

The LBC was plant-mixed using 1.5% type GB cement and placed with a paver. UCS test results were not
available for this project.

Details of the original surfacing are not available. In 2011 the pavement was resurfaced using
microsurfacing.

A.11.2 Performance Summary

The 7.8 km study section was 16 years old when inspected in 2016 and had experienced a cumulative traffic
loading 6.6 x 106 ESA which exceeded the design traffic loading of 5.9 x 106 ESA. The pavement was
generally in sound condition considering its age.

The pavement had performed well except for:


• fine to large width transverse shoulder cracks (cementitious origin) occurring at a varying (5–15 m)
spacing along the entire project
• one small (10 m2) section of block cracking which is considered negligible in the 7.8 km long project
(Figure A 21).

A microsurfacing was placed in 2011 and had performing well when inspected. Whilst the reason for the
microsurfacing is unknown, it is usually selected for its rut filling and surface texture correction abilities. The
more recent microsurfacing treatments utilise a polymer modified bitumen binder and a fibre to resist
cracking. Whilst there was a lot of fine transverse cracking in the shoulder that was beginning to show
through, the microsurfacing did not appear to resist the larger crack widths in the shoulder area. Interestingly,
the transverse cracks were very rarely observed in the traffic lanes.

Figure A 21: Site 12 inspection

Austroads 2020 | page 144


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

A.12 Site 13: Bruce Highway (Townsville – Ingham)

A.12.1 Overview

Site 12 was the Bruce Highway (Townsville to Ingram) between chainages 87.6 km and 92.3 km in Northern
Queensland District. The terrain was flat between 87.6–90 km and slightly undulating 90.0–92.3 km
(Figure A 22).

Figure A 22: Site 13 Bruce Highway Townsville to Ingram

The pavement was constructed in 2001 with the following pavement composition:
• 180 mm thickness of LBC base
• 205 mm thickness of cement-treated subbase (TMR category 1)
• subgrade with a design CBR of 7%.

The design traffic loading was 5.7 x 106 ESA.

The LBC was plant-mixed using 1.5% type GB cement and placed with a paver. UCS test results were not
available for this project.

Details of the original surfacing are not available. In 2009 the pavement was resurfaced using
microsurfacing.

A.12.2 Pavement Summary

The 4.7 km study section was 15 years old when inspected in 2016 and had experienced a cumulative traffic
loading of 5.2 x 106 ESA which was close to the design traffic loading of 5.7 x 106 ESA.

Similar to Site 12, the pavement had performed well except for (Figure A 23):
• fine to large width transverse cracks (cementitious origin) occur intermittently in the shoulder region
• two locations (240 m2) of crocodile cracking in the OWP, equating to 1.1% of the total project wheelpath
area.

Austroads 2020 | page 145


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure A 23: Site 13 inspection

A microsurfacing was placed in 2009 was performing well. Whilst the reason for the microsurfacing is
unknown, it is usually selected for its rut filling and surface texture correction abilities. The more recent
microsurfacing treatments utilise a polymer modified bitumen binder and a fibre to resist cracking. It is
suspected that microsurfacing works in 2009 may not have included the fibres. It was not able to resist the
crocodile cracking occurring at chainages 89.8 km and 90.5 km.

A.13 Sites 14 & 14A: Bundaberg – Miriam Vale Road

A.13.1 Overview

Sites 14 and 14A were located on the Bundaberg – Miriam Vale Road between chainages 59.9 km and
61.6 km in Fitzroy District. The terrain is slightly undulating, dipping down to baffle creek on both sides.
Whilst some table drains are close to the pavement, they drain well (Figure A 24).

The pavement was rehabilitated in 2010 to form the following pavement structure:

Pavement structure:
• Site 14 59.9–61.0 km: 200 mm thickness of LBC directly on subgrade with design CBR of 10–20% as
indicated on the TMR project drawings.
• Site 14A 61.2–61.6 km: 200 mm thickness of LBC on 350 mm of unbound granular subbase on subgrade
with a design CBR of 10–20%.

The design traffic loading was 3 x 106 ESA.

The CBM was produced by in situ stabilisation. To provide sufficient material to stabilise, a 50 mm thickness
of imported granular base was placed on the surface before in situ stabilisation. The LBC was produced
using 2% type GB cement. UCS test results were not available for this project.

Austroads 2020 | page 146


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure A 24: Sites 14 and 14A Bundaberg-Miriam Vale Road

A.13.2 Performance Summary

This 1.7 km long section was about 6 years old when inspected in 2016 and had experienced a cumulative
traffic loading of 4.5 x105 ESA which is 15% of the design loading 3 x 106 ESA.

Two distinct pavement compositions either side of Baffle Creek, labelled site 14 and site 14A as described
above. The inspection found (Figure A 25):
• Site 14A which includes 350 mm thick granular subbase) has no defects.
• Site 14 where the LBC is not supported by a granular subbase had defects (cracking, patches, and
rut/shoves) over about 13% (558 m) of its length after only 6 years.

It was concluded that Site 14A had performed reasonably well (Figure A 25), but that Ste 14 was not
performing well given its condition after only 6 years in-service.

Figure A 25: Site 14A inspection

Austroads 2020 | page 147


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

A.14 Site 15 Western Yeppoon – Emu Park Road

A.14.1 Overview

Site 15 was Western Yeppoon – Emu Park Road between chainages 14.4 km and 16.3 km in Fitzroy District.
The terrain was flat with the majority of this section across swampy wet terrain (Figure A 26).

The pavement was rehabilitated in 2010 by in situ stabilisation to a depth of 150 mm. The 150 mm thick LBC
pavement is on an embankment for the majority of the length, well elevated over swampy ground. The
pavement layers under the LBC are not recorded in the TMR asset management database. It is suspected
the LBC was supported by at least 200 mm thickness of the unbound granular pavement, 500 mm
earthworks (select fill, CBR ≥ 10) and geotextile wrapped rockfill at natural surface level exists. The natural
subgrade had an estimated design CBR of 1%.

Figure A 26: Site 15 Western Yeppoon-Emu Park Road

The CBM was produced by in situ stabilisation to a depth of 150 mm using 2% type GB cement.

With 2% type GB cement, the 7-day UCS was measured to be 0.5–0.8 MPa, which was below the 7-day
target range of 1.0–1.5 MPa.

A.14.2 Performance Summary

The 1.9 km study section was six years old when inspected in 2016 at which time it had a cumulative traffic
loading of 4.2 x 105 ESA.

No cracking was evident, and the pavement was performing reasonably well after six years, except for
flushing/embedment of the sprayed seal surfacing (Figure A 27).

Austroads 2020 | page 148


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure A 27: Site 15 inspection

A.15 Site 16: Dawson Highway (Gladstone – Biloela)

A.15.1 Overview

Site 16 was the Dawson Highway (Gladstone – Biloela) between chainages 50.8 km and 56.1 km in Fitzroy
District (Figure A 28). The terrain was undulating with table drainage mostly good, but some sections only
fair.

Figure A 28: Site 16 Dawson Highway Gladstone to Biloela

Austroads 2020 | page 149


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

In 2007 Dawson Highway was rehabilitated using in situ cementitious stabilisation to a depth of 150 mm. The
material stabilised comprised 100 mm thickness of imported gravel plus 50 mm of existing pavement
material. Whilst there are some anomalies in the TMR asset management database, it was considered that
there was 280 mm of granular material under the LBC. The underlying sandy/gritty clay had an estimated
CBR of 5–7%.

The LBC base was produced by in situ stabilisation to a depth of 150 mm using 2.5% type GB cement.

The ‘as constructed’ drawings reported cement contents and UCS test results measured during construction.
The average cement content cement content was 2.8% resulting an average 7-day UCS of 1.5 MPa at the
top end of the 7-day target range of 1.0–1.5 MPa. Individual UCS results were in the range 0.9 MPa and
2.3 MPa, with 4 out 12 results between 2.2–2.3 MPa.

A.15.2 Performance Summary

The 5.3 km study section was about nine years old when inspected in 2016 and had experienced a
cumulative traffic loading of 1.0 x 106 ESA.

Defects recorded in this section included crocodile cracking, patches and rut/shoving (Figure A 29). The
defects totalled 1.7% (362 m) of the total wheelpath length.

Considering the minor amount of distress after nine years, the pavement was concluded to be performing
satisfactorily.

Figure A 29: Site 16 inspection

Austroads 2020 | page 150


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

A.16 Sites 17 & 17A: Dawson Highway (Biloela – Banana)

A.16.1 Overview

Sites 17 and 17A were located on the Dawson Highway (Biloela – Banana) between chainages 9.7 km and
45.0 km in Fitzroy District. The terrain was flat between 9.7–12 km, slightly undulating to hilly 12–42 km the
flat again between 42–45 km. The surface drainage varied from poor to fair to good (Figure A 30).

Figure A 30: Site 17 and 17A Dawson Highway Biloela to Banana

In 2007 the pavement was rehabilitated by in situ stabilisation using 2% type GB cement. No records of UCS
testing were obtained. The thickness of LBC varied along the project (Figure A 31, Figure A 32) but was
generally 150 mm and 180 mm thick.

According to local TMR staff, the TMR asset management database did not represent accurately the
granular materials under the LBC for this project. It seems likely that there is 250–400 mm thickness of
existing unbound granular pavement.

The review of the performance was complicated by severe rain depressions in 2011–13 which caused
approximately 9.9 km (full-width equivalent) or 28% of total wheelpath length to be reconstructed in 2013.
The reconstruction work consisted of re-in situ modification via stabiliser/reclaimer with some sections also
obtaining a 125 mm plant-mixed (1% type GB) cement-modified gravel overlay. The majority (80%) of the
reconstruction occurred between chainages 30–45 km.

The review was divided as follows:


• Site 17 9.7–26.8 km: limited areas of re-stabilised in 2013
• Site 17A 26.8–45 km: the majority of length was re-stabilised in 2013.

Austroads 2020 | page 151


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure A 31: Project construction records sub-sections 1A to 1G

Source: TMR.

Figure A 32: Project construction records sub-sections 2A to 3G

Source: TMR.

Austroads 2020 | page 152


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

A.16.2 Performance Summary

This large 35.3 km section was about nine years old when inspected in 2016 at which time the pavements
had experienced the following cumulative traffic loadings over nine years:
• Site 17 9.7–26.8 km: 1.17 x 106 ESA, which was 98% of the design traffic loading 1.2 x 106 ESA
• Site 17A 26.8–45 km: 9.4 x 105 ESA, which was 62% of the design traffic loading 1.2 x 106 ESA.

The majority of the 2013 reconstruction occurred in Site 17A.

Concerning the performance of Site 17 (Figure A 33), which included only minor areas of re-stabilisation,
pavement distress was observed in isolated areas:
• two areas totalling 40 lineal metres (48 m2) of crocodile cracking
• 50 lineal metres of longitudinal cracking in the OWP
• 354 m2 of rut/shove failure
• 2730 m2 of patches.

For Site 17 a total of 2% of the wheelpath length displayed one of these three forms of distress. Given the
pavement had already carried its design traffic by 2016, the LBC performed well at Site 17.

Site 17A has fewer deflects in 2016 but this was likely to be due to substantial lengths that needed to be
reconstructed in 2013. Because substantial areas needed to be reconstructed in 2013, the performance of
the LBC was considered to be poor in this section of the project.

Austroads 2020 | page 153


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure A 33: Site 17 inspection

Austroads 2020 | page 154


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Appendix B Surface Deflections Bruce Highway Collinson’s Lagoon


Figure B 1: 28-day measured deflections

Normalised to 40 kN Deflection Results


Client ARRB Start Date of Testing 17-Jan-17 Target Load (kN) 40
Road Bruce Highway (Ayr to Townsville) Road No. 10L Start Reference FWD 0.0 m = TMR 13.90 km (Approx. 280 m north of the Lochinvar Road
Comments2B = Southbound Towards Ayr In Between Wheelpaths; 2L = Southbound Towards Ayr In Outer Wheelpath Centreline)
FWD DTMR Deflections (microns) Surf
Air Temp.
Chainage Chainage Lane Geophone Radius (mm) Temp.
(oC)
(m) (m) D0 D200 D300 D400 D500 D600 D700 D800 D900 D1000 D1200 D1400 D1600 D1800 D2000 D2250 D2500 (oC)
0 13900 2B 116 93 85 78 73 64 58 53 49 45 34 26 22 20 18 15 13 27.0 29.0
10 13890 2B 117 97 88 79 73 64 58 51 46 43 32 25 22 21 19 15 15 27.0 29.0
20 13880 2B 137 111 99 89 84 74 66 60 55 53 41 35 29 27 25 20 19 27.0 29.0
30 13870 2B 223 171 145 124 110 93 82 72 66 62 47 40 35 33 29 25 23 27.0 29.0
40 13860 2B 224 181 157 136 121 105 92 81 73 67 53 44 38 36 31 26 24 27.0 29.0
50 13850 2B 198 158 141 126 112 98 88 78 72 66 51 43 36 33 30 25 23 27.0 29.0
60 13840 2B 200 171 157 136 124 111 101 90 82 74 56 46 39 36 32 26 23 27.0 29.0
70 13830 2B 171 131 119 106 98 86 78 69 63 59 46 38 33 30 27 22 21 27.0 29.0
80 13820 2B 161 140 126 112 101 89 79 71 64 60 47 39 34 32 28 24 22 27.0 29.0
90 13810 2B 171 136 122 108 98 85 77 68 62 58 46 38 33 31 27 23 21 27.0 29.0
100 13800 2B 158 130 119 108 101 89 82 74 69 65 52 43 38 34 30 25 22 27.0 30.0
0 13900 2L 122 100 91 82 75 66 59 53 49 45 34 28 23 21 19 15 14 27.0 30.0
10 13890 2L 115 94 85 75 70 61 55 49 45 42 31 24 22 21 19 16 15 27.0 29.0
20 13880 2L 125 101 92 83 78 69 63 57 53 50 40 34 30 28 25 21 19 27.0 29.0
30 13870 2L 183 148 131 112 102 87 77 68 62 57 45 38 33 31 28 23 21 27.0 29.0
40 13860 2L 153 124 113 102 94 82 75 67 63 59 46 39 34 32 28 24 22 27.0 29.0
50 13850 2L 145 117 106 96 89 80 73 65 60 56 45 38 33 31 28 23 22 27.0 29.0
60 13840 2L 104 94 91 87 85 77 72 67 63 59 47 39 33 31 28 23 21 27.0 29.0
70 13830 2L 116 96 88 81 76 68 61 55 52 49 38 33 29 27 25 21 19 27.0 29.0
80 13820 2L 127 107 97 88 83 73 67 61 56 53 42 36 32 30 27 22 21 27.0 29.0
90 13810 2L 132 111 102 88 81 70 64 58 53 50 39 34 30 28 26 22 20 28.0 29.0
100 13800 2L 147 122 111 100 94 84 76 69 64 61 49 42 37 34 30 25 23 28.0 29.0

Austroads 2020 | page 155


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure B 2: 90-day measured deflections


Normalised to 40 kN Deflection Results
Client ARRB Start Date of Testing 04-Apr-17 ( pp Target Load (kN) 40
Road Bruce Highway (Ayr to Townsville) Road No. 10L Start Reference Centreline)
Comments2B = Southbound Towards Ayr In Between Wheelpaths; 2L = Southbound Towards Ayr In Outer Wheelpath
FWD DTMR Deflections (microns) Surf
Air Temp.
Chainage Chainage Lane Geophone Radius (mm) Temp.
(oC)
(m) (m) D0 D200 D300 D400 D500 D600 D700 D800 D900 D1000 D1200 D1400 D1600 D1800 D2000 D2250 D2500 (oC)
0 13900 2L 112 93 89 84 81 73 69 60 60 56 46 38 32 28 23 18 18 25.3 24.0
10 13890 2L 107 91 87 82 79 72 68 59 58 55 44 36 29 26 23 18 19 25.3 23.9
20 13880 2L 104 90 86 80 77 69 64 57 56 54 44 38 33 30 27 22 22 25.6 24.1
30 13870 2L 124 110 106 102 98 92 83 74 68 64 52 43 38 34 30 24 24 25.6 23.9
40 13860 2L 146 141 122 109 102 91 83 75 70 65 51 43 37 34 31 25 25 25.6 24.0
50 13850 2L 158 123 112 102 95 84 76 70 64 60 48 40 35 32 29 24 24 25.6 23.9
60 13840 2L 115 97 94 89 86 78 74 70 66 63 52 45 39 35 32 26 25 25.5 23.8
70 13830 2L 105 86 82 77 73 67 62 58 54 52 42 36 32 30 27 22 22 25.7 23.7
80 13820 2L 107 90 86 82 80 74 70 66 63 59 48 40 36 32 30 23 23 25.6 23.7
90 13810 2L 99 83 80 76 73 67 63 58 55 53 43 36 32 30 27 22 22 25.6 23.7
100 13800 2L 126 106 102 96 92 84 78 72 67 63 50 43 37 34 31 24 24 25.5 23.7
0 13900 2B 108 90 87 83 79 73 68 64 60 57 46 38 32 28 25 18 18 25.6 23.7
10 13890 2B 128 107 101 94 90 81 75 69 64 60 48 39 31 28 24 19 19 25.5 23.7
20 13880 2B 125 102 96 90 85 77 71 65 61 58 46 39 34 32 28 22 22 25.5 23.8
30 13870 2B 177 163 140 122 113 99 88 79 72 67 53 44 38 34 30 25 24 25.7 24.0
40 13860 2B 197 160 146 131 118 104 94 84 76 71 55 46 39 36 32 26 24 25.7 23.9
50 13850 2B 148 125 117 109 101 91 84 77 71 67 53 45 39 35 32 26 23 25.7 24.0
60 13840 2B 170 142 132 121 113 102 94 86 79 74 59 50 43 38 33 26 24 25.7 24.0
70 13830 2B 128 113 107 99 93 84 78 71 66 62 50 42 37 33 29 24 22 25.7 23.9
80 13820 2B 124 102 97 92 89 82 77 72 68 65 52 44 38 34 30 25 22 25.7 23.9
90 13810 2B 113 95 91 86 83 75 70 65 61 58 47 40 34 31 28 23 21 25.8 24.2
100 13800 2B 127 108 104 99 95 87 81 73 68 64 51 44 38 34 31 25 23 25.8 24.1

Austroads 2020 | page 156


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure B 3: 1-year measured deflections

Normalised to 40 kN Deflection Results


Client ARRB Start Date of Testing 18-Jan-18 Target Load (kN) 40
Road Bruce Highway (Ayr to Townsville) Road No. 10L Start Reference FWD 0.0 = DTMR Chainage 13900 m
Comments 2B = Southbound Towards Ayr In Between Wheelpaths; 2L = Southbound Towards Ayr In Outer Wheelpath
FWD DTMR Deflections (microns) Air Surf
Chainage Chainage Lane Geophone Radius (mm) Temp. Temp.
(m) (m) D0 D200 D300 D400 D500 D600 D700 D800 D900 D1000 D1200 D1400 D1600 D1800 D2000 D2250 D2500 (oC) (oC)
0 13.90 2B 98 85 82 76 73 67 62 58 55 52 41 34 29 25 22 18 12 24.6 27.1
10 13.89 2B 117 98 94 87 83 76 70 65 60 56 45 37 30 27 24 19 13 24.5 26.9
20 13.88 2B 111 96 91 84 80 72 67 62 58 54 44 37 32 30 26 21 14 24.3 26.9
30 13.87 2B 138 123 120 109 102 92 82 75 70 65 52 44 38 34 30 25 17 24.2 27.0
40 13.86 2B 156 137 130 118 110 98 89 81 75 69 55 46 39 36 32 26 18 24.1 26.9
50 13.85 2B 149 126 120 109 102 92 84 77 71 66 53 44 38 35 31 25 17 24.0 26.8
60 13.84 2B 147 123 115 105 99 89 83 76 71 66 54 46 40 36 31 25 17 24.0 26.9
70 13.83 2B 126 114 108 99 93 84 77 71 66 62 49 42 36 33 29 23 16 24.0 26.9
80 13.82 2B 116 96 91 86 82 75 70 65 62 59 48 41 36 32 29 24 16 24.0 26.8
90 13.81 2B 112 96 92 86 83 75 69 65 61 57 46 40 34 32 28 23 16 23.9 26.9
100 13.80 2B 112 96 93 88 85 79 74 69 65 61 49 43 37 34 30 25 17 23.9 26.9
0 13.90 2L 104 90 86 81 78 71 66 61 58 54 43 36 31 27 22 18 12 23.8 26.6
10 13.89 2L 103 89 85 80 76 70 65 60 56 52 42 33 28 26 23 18 12 23.6 26.4
20 13.88 2L 104 90 86 80 76 69 64 59 55 52 42 36 32 30 26 21 14 23.7 26.5
30 13.87 2L 115 102 100 95 92 86 77 69 65 61 49 42 37 34 30 24 16 23.7 26.6
40 13.86 2L 124 120 113 98 93 84 78 71 66 62 50 42 37 34 31 25 17 23.8 26.7
50 13.85 2L 167 119 112 101 94 85 78 71 66 61 49 42 36 33 30 25 17 24.0 26.7
60 13.84 2L 102 88 86 80 78 72 67 63 60 57 47 41 36 33 29 24 16 24.2 26.7
70 13.83 2L 111 100 94 87 83 75 69 64 60 56 45 38 34 31 28 22 15 24.2 26.7
80 13.82 2L 97 83 80 75 73 67 63 59 56 53 43 37 33 31 27 22 15 24.2 26.7
90 13.81 2L 102 87 85 79 76 69 65 60 56 53 43 37 32 30 27 22 15 24.1 26.5
100 13.80 2L 112 100 97 92 88 80 75 69 65 61 49 42 37 34 30 25 17 24.1 26.5

Austroads 2020 | page 157


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Appendix C Surface Deflections Bruce Highway Barratta Creeks


Figure C 1: 28-day measured deflections

Normalised to 40 kN Deflection Results


Client ARRB Start Date of Testing 04-Apr-17 Target Load (kN) 40
Road Bruce Highway (Ayr to Townsville) Road No. 10L Start Reference FWD 0.0 m = DTMR 23.02 km
Comment 1B = Northbound Towards Townsville In Between Wheelpaths; 1L = Northbound Towards Townsville In Outer Wheelpath
FWD DTMR Deflections (microns) Surf
Air Temp.
Chainage Chainage Lane Geophone Radius (mm) Temp.
(oC)
(m) (m) D0 D200 D300 D400 D500 D600 D700 D800 D900 D1000 D1200 D1400 D1600 D1800 D2000 D2250 D2500 (oC)
0 23020 1B 235 205 189 176 165 149 138 121 105 96 73 61 53 47 42 35 32 25.3 25.6
10 23030 1B 631 474 384 325 282 243 210 183 162 144 109 88 73 64 57 48 45 25.3 25.7
20 23040 1B 634 507 420 365 321 273 237 200 178 157 116 92 77 68 61 52 49 25.4 25.8
30 23050 1B 634 436 351 297 257 219 192 167 145 130 98 81 69 62 56 48 45 25.4 25.7
40 23060 1B 616 453 354 299 258 216 191 163 144 127 95 78 64 57 50 43 39 25.4 25.7
50 23070 1B 529 436 363 307 265 222 192 165 144 127 97 77 64 56 48 40 37 25.2 25.7
60 23080 1B 481 383 319 277 244 212 189 166 148 133 102 81 66 56 48 37 34 25.2 25.6
70 23090 1B 429 330 276 238 211 183 165 144 128 115 88 72 58 52 46 38 37 25.2 25.5
80 23100 1B 559 421 323 271 235 200 175 151 134 120 91 74 64 56 49 41 39 25.2 25.5
90 23110 1B 675 444 316 245 201 164 140 120 107 99 76 63 55 50 45 38 36 25.2 25.6
100 23120 1B 608 428 322 258 215 177 152 130 115 104 80 67 57 53 47 40 38 25.3 25.6
0 23020 1L 301 247 211 184 163 140 125 111 100 92 73 62 55 50 45 37 39 25.5 25.6
10 23030 1L 803 621 496 424 363 304 263 226 198 176 132 107 88 79 70 60 63 26.1 25.8
20 23040 1L 776 604 493 420 355 298 256 218 189 166 125 101 86 76 70 60 62 25.8 25.7
30 23050 1L 667 491 402 346 297 255 223 189 165 147 111 90 76 69 63 54 57 25.9 25.6
40 23060 1L 515 415 340 294 256 219 190 163 144 129 98 80 66 59 53 45 47 25.9 25.8
50 23070 1L 595 497 416 357 308 257 222 190 167 148 113 92 77 66 58 49 45 25.8 25.9
60 23080 1L 516 406 356 322 289 251 222 195 172 155 118 97 79 69 57 45 40 26.0 26.0
70 23090 1L 494 402 339 297 263 228 201 172 153 136 103 85 71 61 54 45 42 25.8 26.0
80 23100 1L 614 488 395 343 298 259 226 194 169 147 108 88 73 64 57 48 44 25.8 25.9
90 23110 1L 466 349 279 234 197 162 139 120 106 96 75 63 55 51 47 40 37 25.8 25.9
100 23120 1L 488 380 306 261 219 182 156 134 117 104 79 66 57 52 47 39 36 25.9 26.0

Austroads 2020 | page 158


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure C 2: 90-day measured deflections

Normalised to 40 kN Deflection Results


Client ARRB Start Date of Testing 02-Jun-17 Target Load (kN) 40
Road Bruce Highway (Ayr to Townsville) Road No. 10L Start Reference FWD 0.0 m = DTMR 23.02 km
Comments 1B = Northbound Towards Townsville In Between Wheelpaths; 1L = Northbound Towards Townsville In Outer Wheelpath
FWD DTMR Deflections (microns) Air Surf
Chainage Chainage Lane Geophone Radius (mm) Temp. Temp.
(m) (m) D0 D200 D300 D400 D500 D600 D700 D800 D900 D1000 D1200 D1400 D1600 D1800 D2000 D2250 D2500 (oC) (oC)
0 23020 1B 172 157 149 141 135 126 121 114 101 85 68 55 49 44 38 32 29 19 19
10 23030 1B 560 418 331 280 244 212 185 163 147 132 101 82 68 59 53 44 39 19 19
20 23040 1B 576 466 380 322 279 238 205 176 155 138 105 83 71 62 56 47 41 19 19
30 23050 1B 521 382 293 247 216 185 162 140 126 114 87 72 63 56 50 43 38 20 19
40 23060 1B 496 392 308 260 224 194 169 146 130 117 89 72 61 54 48 40 34 20 20
50 23070 1B 531 449 374 318 276 235 204 174 153 135 100 78 65 55 49 41 36 20 20
60 23080 1B 423 338 275 242 218 195 179 160 144 130 100 81 67 57 47 38 33 20 20
70 23090 1B 565 437 358 309 274 243 215 189 168 152 117 93 76 65 54 46 41 20 20
80 23100 1B 462 334 260 219 195 169 150 132 119 109 84 68 58 51 42 37 33 20 20
90 23110 1B 519 370 267 213 178 149 128 112 101 91 72 59 52 47 42 35 32 20 20
100 23120 1B 494 367 278 228 195 163 141 123 110 99 77 63 54 48 43 37 33 20 20
0 23020 1L 232 218 195 159 142 128 114 103 97 87 69 58 51 47 42 35 33 19 20
10 23030 1L 758 600 457 374 320 272 236 205 181 161 123 100 84 74 67 57 50 19 20
20 23040 1L 785 614 471 393 333 277 234 197 175 155 115 95 80 72 65 57 54 19 20
30 23050 1L 691 528 394 326 277 231 195 166 149 132 101 82 71 64 58 50 46 19 20
40 23060 1L 557 438 339 286 245 208 178 151 133 119 91 74 64 57 51 43 41 19 20
50 23070 1L 653 534 435 367 310 259 223 186 161 142 109 86 73 64 57 48 44 19 20
60 23080 1L 531 434 365 319 284 243 212 185 164 146 111 88 73 63 54 46 40 19 20
70 23090 1L 637 537 447 379 317 260 218 186 163 142 105 85 71 64 57 50 45 19 19
80 23100 1L 586 479 383 315 267 225 193 166 149 133 97 79 66 58 53 45 40 19 20
90 23110 1L 428 337 264 215 183 153 129 112 100 91 74 60 53 49 44 37 34 19 20
100 23120 1L 518 402 314 261 215 175 148 126 111 100 78 65 57 52 46 40 36 19 20

Austroads 2020 | page 159


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure C 3: 1-year measured deflections

Normalised to 40 kN Deflection Results


Client ARRB Start Date of Te 15-Mar-18Target Load (kN) 40
Road Bruce Highway (Ayr to Townsville) Road No. 10L Start Reference FWD 0.0 m = DTMR 23.02 km
Commen1B = Northbound Towards Townsville in Between Wheelpaths; 1L = Northbound Towards Townsville In Outer Wheelpath
FWD DTMR Deflections (microns) Air Surf
Chaina Chain Lane Geophone Radius (mm) Tem Tem
ge (m) age D0 D200 D300 D400 D500 D600 D700 D800 D900D1000D1200D1400D1600D1800D2000D2250D2500 p. p.
0 23020 1B 177 161 156 149 144 137 132 125 112 92 70 58 50 46 41 33 23 24.3 24.9
10 23030 1B 574 429 351 294 255 217 188 167 148 132 100 82 69 61 55 46 32 24.2 25.0
20 23040 1B 585 463 392 332 287 242 209 177 156 137 104 85 71 64 57 48 35 24.2 24.9
30 23050 1B 588 409 318 263 227 192 167 145 130 116 90 75 64 58 52 45 31 24.1 25.0
40 23060 1B 746 577 474 396 344 294 252 215 188 164 120 95 78 68 60 49 35 24.1 24.9
50 23070 1B 602 478 403 341 294 253 220 189 166 147 111 90 75 66 58 48 32 24.1 24.8
60 23080 1B 555 449 376 327 290 253 223 194 172 151 115 93 77 66 58 47 33 24.1 24.8
70 23090 1B 639 510 424 355 309 264 226 193 171 150 110 88 73 64 58 50 35 24.1 24.5
80 23100 1B 588 447 370 311 269 226 197 170 152 133 101 81 67 59 52 42 30 24.2 24.6
90 23110 1B 576 407 302 232 190 157 134 116 103 94 74 61 53 49 44 37 26 24.2 24.6
100 23120 1B 540 388 304 243 203 168 147 127 113 101 79 66 56 50 45 38 26 24.3 24.7
0 23020 1L 299 246 222 187 163 137 119 104 94 86 70 60 53 49 44 37 26 24.3 25.5
10 23030 1L 899 680 523 408 334 279 241 204 178 158 121 98 84 74 67 57 41 24.3 25.3
20 23040 1L 896 701 547 433 347 282 242 203 177 155 118 97 83 75 69 58 41 24.3 25.1
30 23050 1L 758 560 425 330 275 227 193 165 147 131 102 85 73 67 60 52 38 24.2 25.1
40 23060 1L 732 562 456 367 297 244 213 178 153 138 106 88 76 69 62 52 37 24.2 25.0
50 23070 1L 794 668 557 458 384 310 250 204 180 157 117 93 80 71 64 54 39 24.2 25.1
59 23079 1L 729 601 512 434 376 314 262 217 193 168 123 100 83 73 64 53 38 24.2 25.0
70 23090 1L 766 623 524 443 372 304 261 224 195 166 122 101 82 72 65 56 41 24.1 25.0
80 23100 1L 780 646 520 403 338 277 237 203 172 150 114 92 77 70 62 51 36 24.2 25.0
90 23110 1L 547 417 325 248 194 153 131 115 104 96 76 66 57 53 48 40 29 24.4 25.2
100 23120 1L 564 442 351 274 225 181 152 129 115 102 80 67 59 54 49 42 29 24.4 25.4

Austroads 2020 | page 160


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Appendix D Material Splitting and Mixing


Procedure

D.1 Introduction
This Appendix describes the method to split and mix samples for wheel tracking fatigue testing (Section 8).

D.2 Splitting
The bulk untreated material was split into representative samples using a motorised rotary splitter shown in
Figure D 1. The splitter has a large funnel into which the bulk material can be loaded. A conveyor belt then
delivers this material at a constant speed to 12 removable canisters which rotate underneath the belt, also at
a constant speed. The process involved splitting two or more batches of 12 numbered (10 L) buckets each.
Corresponding numbers from each of the batches would then be combined for a second (final) round of
splitting to improve uniformity between samples.

Figure D 1: Motorised rotary splitter and splitting process

Source: Austroads (2014).

D.3 Mixing Procedure


As the compacted slab are 700 mm long x 500 mm wide x 100 mm deep, the volume of material in each slab
is 0.035 m3. As the target density ratio was 95%, the target dry densities were 2.19 t/m3 and 2.20 t/m3 for
1.5% and 3.0% mixes. Hence about 80 kg of each mix was required.

Batches of aggregate material were weighed and preconditioned (addition of moisture) 24 hours prior to
mixing. The preconditioned mix was then combined with the remaining water and cement in a motor-driven
planetary concrete mixer with a tank size of 800 mm diameter and 350 mm high (Figure D 2).

The mixing process was as follows:


• the bulk preconditioned (moisture prepared) host material was placed into the mixing tank
• the mixer was run for 15 seconds to spread the material evenly in the tank

Austroads 2020 | page 161


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

• the required amount of type GB cement was added


• the final amount of water was added to reach the target moisture content of modified Proctor optimum
moisture content
• the mixer was run for 120 seconds then let stand for 120 seconds
• the mixer was run for a further 120 seconds.

Figure D 2: Planetary concrete mixer used to mix cemented material

Source: Austroads (2014).

Following mixing, the material was placed in containers and covered with a plastic sheet and let stand for
some time before compaction. The intention was to make allowance for the commencement of the cement
binder reaction and also to replicate the field placement of cemented materials which can involve quarry
mixing and then the delivery time before placement. The standing time was at most 30 minutes as beyond
this time samples were too difficult to compact due to the initial ‘set’ of the cement binder.

Austroads 2020 | page 162


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Appendix E Compaction of Wheel-tracking Slabs


Section 8 describes the testing undertaken with the Austrack wheel-tracking device. The following process
was used to compact 100 mm thick slab for testing using the roller segment fitted to Austrack:
a. Remove the skateboard base plate from the wheel tracker and rest on the support arms.
b. Roll the trolley into place between the support arms.
c. Raise the trolley so that the base plate can be removed.
d. Take the first and second mould assemblies and attach them to the base plate.
e. Place the low-density polyethylene (LDPE) in the assembly to a depth 200 mm below the top of mould
assembly.
f. Take the mould assembly fence and attach it to the top of the mould assembly.
g. Using the trolley lower the mould assembly and base plate onto the support arms.
h. Transfer the half of the treated material into the compaction mould.
i. Manually spread the material uniformly throughout the mould assembly (avoiding segregation where
possible).
j. Using the hand-held tamping device, compact the material ensuring even distribution.
k. Transfer the remaining half of the treated material into the compaction mould.
l. Manually spread the material uniformly throughout the mould assembly (avoiding segregation where
possible).
m. Place the plastic sheet on top of the material.
n. Push the mould assembly filled with treated material into the wheel-tracker and lock in place.
o. Set the target height to 100 mm above the LDPE.
p. Using the roller segment, compact the material using the compaction steps shown in Table E 1. The
compaction sequence is automatically terminated when the measured height is equal to the targeted
height. This may occur before completion of all the compaction steps in Table E 1.
q. As the LDPE is elastic, the layer height needs to be checked at a minimum applied load: the so-called
unloaded height. Reduce the applied load to 2 kN, apply five loading cycles and record the unloaded
height of the top of the slab above the LDPE.
r. If the unloaded height ≤ 100 mm go to step (t).
s. If the unloaded height > 100 mm, then repeat step (p) using a corrected target compaction height. For the
correction, subtract the difference between the height measured after step (p) and the initial target of
100 mm (example, if 102 mm is measured then the next compaction stage, will target 100 – 2 = 98 mm).
t. Assign compacted specimen an identifying number.

Note: After completion of the compaction, the thickness of the slab at all points of the measuring zone must
not within 2 mm of the 100 mm target slab thickness.

Austroads 2020 | page 163


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Table E 1: Typical layer compaction sequence

Load
Compaction phases Compaction passes number
(kN)
Pre-compaction Compaction load gradually increased from 1 to 9 kN in increments of 2 kN every two
passes (i.e. one cycle)
Compaction 10 6
14 6
18 6
20 6
25 6
30 20

Austroads 2020 | page 164


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Appendix F Wheel-tracking Fatigue Test Method


Section 8 describes the fatigue testing undertaken with the Austrack wheel-tracking device. The following
test procedure was used:
a. Roll the trolley into place between the support arms.
b. Raise the trolley so that the final compacted material assembly can be removed.
c. Using the trolley remove the final compacted material assembly.
d. Using the compaction segmented plate separation device ensures that the compaction device is secured
to the wheel tracker.
e. Use the trolley to replace the final compacted material assembly onto the support arms.
f. Push the final compacted material assembly into the wheel tracker and lock in place.
g. Set the displacement measuring device such that there is sufficient downward and upward range to cater
for material rutting if any.
Note: Refer to the manufacturer’s operating instructions for appropriate settings.
h. Record the tracked surface profile of the wheelpath.
Note: The profiles are recorded across five cross-sections distributed symmetrically on each side of the
centre of the specimen as per the pattern shown in Figure F 1. They are recorded at given stages of the
test (i.e. number of cycles). The mean rut depth can be calculated as the average of the specimen
deformation.
i. Commence wheel-tracking test using a 20 kN wheel load for a maximum of 100 000 load cycles.
j. On a clear plastic sheet map the length of cracking every 10 000 to 20 000 cycles (morning and evening).
Use a different colour ink on each occasion to provide a record of the crack progression.
k. Readings of the specimen deformation are taken every 10 000 to 20 000 cycles.
l. The test on a specimen is completed when one of the following termination conditions is reached:
♦ The required number of cycles is reached.
♦ The overall mean deformation exceeds 10 mm.

Figure F 1: Location of rut depth measuring points

Source: Austroads (2013b).

Austroads 2020 | page 165


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Appendix G Wheel-tracking Test Deflection Plots

Figure G 1: Slab 5555 deflection change with loading cycles

Figure G 2: Slab 5568 deflection change with loading cycles

Austroads 2020 | page 166


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure G 3: Slab 5688 deflection change with loading cycles

Figure G 4: Slab 4833 deflection change with loading cycles

Austroads 2020 | page 167


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure G 5: Slab 4866 deflection change with loading cycles

Figure G 6: Slab 4932 deflection change with loading cycles

Austroads 2020 | page 168


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure G 7: Slab 5029 deflection change with loading cycles

Figure G 8: Slab 5383 deflection change with loading cycles

Austroads 2020 | page 169


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure G 9: Slab 5492 deflection change with loading cycles

Figure G 10: Slab 5694 deflection change with loading cycles

Austroads 2020 | page 170


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure G 11: Slab 5922 deflection change with loading cycles

Figure G 12: Slab 5988 deflection change with loading cycles

Austroads 2020 | page 171


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure G 13: Slab 6028 deflection change with loading cycles

Figure G 14: Slab 5945 deflection change with loading cycles

Austroads 2020 | page 172


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure G 15: Slab 5951 deflection change with loading cycles

Figure G 16: Slab 4811 deflection change with loading cycles

Austroads 2020 | page 173


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure G 17: Slab 4876 deflection change with loading cycles

Figure G 18: Slab 5469 deflection change with loading cycles

Austroads 2020 | page 174


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure G 19: Slab 5541 deflection change with loading cycles

Figure G 20: Slab 5928 deflection change with loading cycles

Austroads 2020 | page 175


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Figure G 21: Slab 6095 deflection change with loading cycles

Austroads 2020 | page 176


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Appendix H Vibrating Table Results


Table H 1: Measured masses of loose material after trafficked slabs were vibrated

Mass Mass
Shaking Mass under Total
under under
Slab ID time wheelpath mass Graph of results
left side right side
(minutes) (kg) (kg)
(kg) (kg)
5555 2 0.668 0.554 0.57 1.792
4 0.216 0.382 0.206 0.804 Slab 5555 (14 dyas, 1% GP, Hornfells)
6 0.168 0.382 0.112 0.662 6000
8 0.412 0.618 0.056 1.086
10 0.410 0.118 0.034 0.562
5000
15 0.380 0.314 0.024 0.718
Total 2.254 2.368 1.002 5.624

Loose Material Mass (g)


4000
Left of Wheel Path

3000 Centre (Under Wheel


Path)
Right of Wheel Path
2000
Sum of All Sections

1000

0
0 5 10 15 20
Shaking Time (Minutes)

Austroads 2020 | page 177


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Mass Mass
Shaking Mass under Total
under under
Slab ID time wheelpath mass Graph of results
left side right side
(minutes) (kg) (kg)
(kg) (kg)
5568 2 0.422 1.06 0.526 2.008
4 0.18 0.418 0.226 0.824 Slab 5568 (14 days, 1% GP, Hornfells)
6 0.086 0.41 0.1 0.596 5000
8 0.052 0.248 0.032 0.332
4500
10 0.058 0.276 0.036 0.37
15 0.094 0.426 0.078 0.598
4000
Total 0.892 2.838 0.998 4.728
3500

Loose Material Mass (g)


Left of Wheel Path
3000

2500 Centre (Under Wheel


Path)
2000 Right of Wheel Path

1500 Sum of All Sections

1000

500

0
0 5 10 15 20
Shaking Time (Minutes)

Austroads 2020 | page 178


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Mass Mass
Shaking Mass under Total
under under
Slab ID time wheelpath mass Graph of results
left side right side
(minutes) (kg) (kg)
(kg) (kg)
5688 2 2.844 3.840 6.100 12.784
4 1.946 0.988 3.442 6.376 Slab 5688 (28 days, 1% GB, Hornfells)
30000
6 1.590 0.926 1.402 3.918
8 0.718 0.436 0.332 1.486
10 1.078 0.638 0.136 1.852
15 1.218 0.23 0.282 1.730 25000

Total 9.394 7.058 11.694 28.146

20000

Loose Material Mass (g)


Left of Wheel Path

Centre (Under Wheel


15000
Path)
Right of Wheel Path

10000 Sum of All Sections

5000

0
0 5 10 15 20
Shaking Time (Minutes)

Austroads 2020 | page 179


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Mass Mass
Shaking Mass under Total
under under
Slab ID time wheelpath mass Graph of results
left side right side
(minutes) (kg) (kg)
(kg) (kg)
5694* 2 0.338 0.728 0.240 1.306
4 0.030 0.078 0.030 0.138 Slab 5694 (14 days, 1.5% GP, Hornfells)
6 0.026 0.120 0.016 0.162 2000
8 0.034 0.080 0.016 0.130
10 0.020 0.024 0.004 0.048 1800

15 0.022 0.068 0.024 0.114


1600
Total 0.47 1.098 0.33 1.898

1400

Loose Material Mass (g)


1200 Left of Wheel Path

1000 Centre (Under Wheel


Path)
800 Right of Wheel Path

600 Sum of All Sections

400

200

0
0 5 10 15 20
Shaking Time (Minutes)

Austroads 2020 | page 180


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Mass Mass
Shaking Mass under Total
under under
Slab ID time wheelpath mass Graph of results
left side right side
(minutes) (kg) (kg)
(kg) (kg)
5922 2 0.022 0.020 0.032 0.074
4 0.004 0.002 0.004 0.010 Slab 5922 (1 day, 2% GB, Barratta)
6 0.002 0.002 0.002 0.006 140
8 0.002 0.002 0.006 0.010
10 0.002 0.002 0.008 0.012
15 0.002 0.002 0.006 0.010 120
Total 0.034 0.03 0.058 0.122

100

Loose Material Mass (g)


Left of Wheel Path
80

Centre (Under
Wheel Path)
60 Right of Wheel
Path
Sum of All Sections
40

20

0
0 5 10 15 20
Shaking Time (Minutes)

Austroads 2020 | page 181


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Mass Mass
Shaking Mass under Total
under under
Slab ID time wheelpath mass Graph of results
left side right side
(minutes) (kg) (kg)
(kg) (kg)
6012 2 0.034 0.026 0.040 0.100
4 0.004 0.002 0.002 0.008 Slab 6012 (1 day, 2% GB, Barratta)
6 0.030 0 0.002 0.032 180
8 0.002 0 0.002 0.004
10 0 0 0 0 160
15 0.004 0.002 0.004 0.01
Total 0.074 0.03 0.05 0.154
140

120

Loose Material Mass (g)


Left of Wheel Path
100
Centre (Under
Wheel Path)
80
Right of Wheel Path

60 Sum of All Sections

40

20

0
0 5 10 15 20
Shaking Time (Minutes)

Austroads 2020 | page 182


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Mass Mass
Shaking Mass under Total
under under
Slab ID time wheelpath mass Graph of results
left side right side
(minutes) (kg) (kg)
(kg) (kg)
5998 2 0.010 0.016 0.022 0.048
4 0.002 0.004 0.010 0.016 Slab 5998 (28 days, 2% GB, Barratta)
6 0.002 0.004 0.002 0.008 80
8 0 0 0 0
10 0 0 0 0
70
15 0 0 0 0
Total 0.014 0.024 0.034 0.072
60

Loose Material Mass (g)


50
Left of Wheel Path

40 Centre (Under Wheel


Path)
Right of Wheel Path
30
Sum of All Sections

20

10

0
0 5 10 15 20
Shaking Time (Minutes)

Austroads 2020 | page 183


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Mass Mass
Shaking Mass under Total
under under
Slab ID time wheelpath mass Graph of results
left side right side
(minutes) (kg) (kg)
(kg) (kg)
6028 2 0.098 3.618 0.072 3.788
4 0.062 0.540 0.010 0.612 Slab 6028 (28 days, 2% GB, Barratta)
6 0.064 0.404 0.026 0.494 7000
8 0.062 0.266 0.032 0.36
10 0.012 0.158 0.064 0.234
15 0.054 0.306 0.068 0.428 6000
Total 0.352 5.292 0.272 5.916

5000

Loose Material Mass (g)


Left of Wheel Path
4000

Centre (Under Wheel


Path)
3000 Right of Wheel Path

Sum of All Sections


2000

1000

0
0 5 10 15 20
Shaking Time (Minutes)

Austroads 2020 | page 184


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Mass Mass
Shaking Mass under Total
under under
Slab ID time wheelpath mass Graph of results
left side right side
(minutes) (kg) (kg)
(kg) (kg)
5945 2 0.048 0.036 0.016 0.100
4 0.014 0.004 0.004 0.022 Slab 5945 (90 days, 2% GB, Barratta)
6 0.004 0.004 0.002 0.01 200
8 0.010 0.002 0.006 0.018
10 0.002 0.002 0.004 0.008 180
15 0.008 0.002 0.004 0.014
Total 0.086 0.05 0.036 0.172 160

140

Loose Material Mass (g)


120 Left of Wheel Path

100 Centre (Under Wheel


Path)
Right of Wheel Path
80

Sum of All Sections


60

40

20

0
0 5 10 15 20
Shaking Time (Minutes)

Austroads 2020 | page 185


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Mass Mass
Shaking Mass under Total
under under
Slab ID time wheelpath mass Graph of results
left side right side
(minutes) (kg) (kg)
(kg) (kg)
5951 2 0.218 1.218 0.262 1.698
4 0.014 0.092 0.032 0.138 Slab 5951 (90 days, 2% GB, Barratta)
6 0.002 0.030 0.028 0.060 3000
8 0.008 0.296 0.012 0.316
10 0.002 0.104 0.008 0.114
15 0.004 0.305 0.015 0.324 2500
Total 0.248 2.045 0.357 2.65

2000

Loose Material Mass (g)


Left of Wheel Path

1500 Centre (Under Wheel


Path)
Right of Wheel Path

1000
Sum of All Sections

500

0
0 5 10 15 20
Shaking Time (Minutes)

Austroads 2020 | page 186


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Mass Mass
Shaking Mass under Total
under under
Slab ID time wheelpath mass Graph of results
left side right side
(minutes) (kg) (kg)
(kg) (kg)
5541* 2 0.008 0.002 0.018 0.028
4 0.018 0 0.006 0.024 Slab 5541 (14 days, 3% GP, Hornfells)
6 0.010 0.002 0.008 0.020 140
8 0.010 0 0 0.010
10 0.008 0 0.002 0.010
15 0.028 0.002 0.008 0.038 120
Total 0.082 0.006 0.042 0.130

100

Loose Material Mass (g)


Left of Wheel Path
80

Centre (Under Wheel


Path)
60 Right of Wheel Path

Sum of All Sections


40

20

0
0 5 10 15 20
Shaking Time (Minutes)

Austroads 2020 | page 187


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Mass Mass
Shaking Mass under Total
under under
Slab ID time wheelpath mass Graph of results
left side right side
(minutes) (kg) (kg)
(kg) (kg)
5928 2 0.018 0.016 0.018 0.052
4 0 0 0.002 0.002 Slab 5928 (1 day, 4% GB, Barratta)
6 0.002 0.002 0.002 0.006 70
8 0 0 0 0
10 0 0 0 0
15 0 0 0.002 0.002 60

Total 0.020 0.018 0.024 0.062

50

Loose Material Mass (g)


Left of Wheel Path
40
Centre (Under Wheel
Path)
30
Right of Wheel Path

Sum of All Sections


20

10

0
0 5 10 15 20
Shaking Time (Minutes)

Austroads 2020 | page 188


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Mass Mass
Shaking Mass under Total
under under
Slab ID time wheelpath mass Graph of results
left side right side
(minutes) (kg) (kg)
(kg) (kg)
6095 2 0.024 0.114 0.046 0.184
4 0.010 0.01 0.004 0.024 Slab 6095 (28 days, 4% GB, Barratta)
6 0.006 0.004 0.004 0.014 300
8 0.004 0.004 0.008 0.016
10 0.002 0.002 0.002 0.006
15 0.004 0.022 0.002 0.028 250
Total 0.050 0.156 0.066 0.272

200

Loose Material Mass (g)


Left of Wheel Path

150 Centre (Under Wheel


Path)
Right of Wheel Path

100
Sum of All Sections

50

0
0 5 10 15 20
Shaking Time (Minutes)

*These specimens had a slurry layer applied to their surface before trafficking in the XL-WT device.

Austroads 2020 | page 189


Development of Design Procedures for Lightly Bound Cemented Materials in Flexible Pavements

Austroads 2020 | page 190

You might also like