Download as pdf or txt
Download as pdf or txt
You are on page 1of 191

Mechanical and durability properties of

ECO-UHPCs incorporating industrial


by-products

Tanvir Ahmed

MSc (Eng), BSc (Eng)

This thesis is presented for the degree of

Doctor of Philosophy

The University of Western Australia

Department of Civil, Mining and Environmental Engineering

2022
i
THESIS DECLARATION

I, Tanvir Ahmed, certify that:

This thesis has been substantially accomplished during enrolment in this degree.

This thesis does not contain material which has been submitted for the award of any other
degree or diploma in my name, in any university or other tertiary institution.

In the future, no part of this thesis will be used in a submission in my name, for any other
degree or diploma in any university or other tertiary institution without the prior approval
of The University of Western Australia and where applicable, any partner institution
responsible for the joint-award of this degree.

This thesis does not contain any material previously published or written by another person,
except where due reference has been made in the text and, where relevant, in the Authorship
Declaration that follows.

This thesis does not violate or infringe any copyright, trademark, patent, or other rights
whatsoever of any person.

This thesis contains published work and/or work prepared for publication, some of which
have been co-authored.

Signature:

Date: 8 March 2022

ii
ABSTRACT

Ultra-high performance concrete (UHPC) is a relatively novel class of concrete, compared


with normal-strength concrete (NSC) and high-strength concrete (HSC). UHPC exhibits
extraordinary compressive strength, durability, and even very high ductility when fibres are
incorporated. Despite its superior performance, in terms of both strength and durability, the
application of this material is still very limited mainly because of its high material CO2
footprint and cost. Use of high-volume of Portland cement (≥700 kg/m3 of cementitious
mix) and very high-quality fine quartz sand (quartz content ≥ 95%, maximum particle size
≤ 600 μm) are two factors contributing to the high material CO2 footprint and cost of the
material. The inclusion of industrial by-product based materials as partial and/or full
substitutes of cement and quartz sand in UHPC may address the foregoing drawbacks.

Ground granulated blast-furnace slag (GGBS) and Class-F fly ash are two supplementary
cementitious materials (SCMs) that are widely used as partial substitutes of cement in NSC
or HSC to reduce the CO2 footprint of the cementitious mix. Nevertheless, such materials
are hardly used in proprietary UHPCs. This is because the knowledge of UHPCs
incorporating GGBS and Class-F fly ash is still too limited to understand the true potentials
of these cementitious mixes and adopt them for construction works.

In Western Australia (WA), gold mines produce millions of tonnes of tailings each year.
Many of these tailings remain unused and are required to be stored in large storage facilities.
Gold mine tailings, being very fine, can be potential substitute for quartz sand. The use of
tailings, instead of quartz sand, can reduce the cost of UHPC. Several environmental
concerns can also be addressed, at least partially, if tailings are used as UHPC aggregate.
For instance, CO2 emission associated with the extraction and grinding of quartz sand,
hazards associated with the mismanagement of tailings (e.g., risk of chemical and toxic
metal pollution), etc. However, a study addressing the properties of UHPC with gold mine
tailings is scarce in the current literature to instil confidence in using such UHPC.

In order to address the gaps, as indicated above, in the existing literature to a certain extent,
the mechanical and durability properties of UHPCs incorporating GGBS and Class-F fly,
mechanical properties and resistance to long-term chloride ingress of ultra-high performance
fibre reinforced concrete (UHPFRC) with GGBS, and mechanical properties, durability
properties, and toxic leaching behaviour of UHPC with gold mine tailings (sourced from a
mine in WA) have been aimed to be investigated. The results suggest that 28-day
compressive strength beyond 150 MPa can be attained for up to 60% and 40% replacements

iii
of cement with GGBS and fly ash, respectively, without the need for any special curing or
fibres. UHPCs with up to 60% fly ash show similar durability to that of the UHPC without
fly ash in terms of corrosion risk, water absorption, initial rate of water absorption, while
UHPCs with up to 60% GGBS show better durability than the UHPC without GGBS. For
UHPCs with up to 60% GGBS or 70% fly ash content, carbonation depths remain below
the detection limit. UHPFRC with 30% GGBS exhibits better flexural performance than
UHPFRC without GGBS. The resistance to chloride penetration of UHPFRC with 30%
GGBS can be categorized as ‘extremely high’. UHPC exhibiting 28-day compressive
strength of greater than 120 MPa, water tightness better than UHPC incorporating quartz
sand, and carbonation resistance similar to conventional UHPC can be produced using gold
mine tailings as aggregate. The leachability of the toxic metals from UHPC, irrespective of
its tailings content, lies well below the regulatory thresholds.

iv
ACKNOWLEDGEMENTS

This research project was supported by a University Postgraduate Award and an Australian
Government Research Training Program (RTP) Fees Offset Scholarship. I acknowledge the
facilities and technical assistance of the Australian Microscopy & Microanalysis Research
Facility at the Centre for Microscopy, Characterisation & Analysis, The University of
Western Australia, a facility funded by the University, State and Commonwealth
Governments.

I am deeply grateful to my supervisory team. Thank you, Mohamed Elchalakani and Ali
Karrech, for your sincere guidance, encouragement, and support to carry out this research.
I express my heartfelt gratitude to Mohamed Elchalakani, who has not only been an
inspiring academic to me but also a true mentor and a caring guardian to both myself and
my family.

I owe a great debt of gratitude to my research team, who helped me generously throughout
the journey. Thank you Huiyuan Liu, Minhao Dong, Alexandra Meek, Waleed Nawaz,
Ehsan Sadrossadat, Thumitha Higgoda, and Xin Lyu.

I thank Jim Waters, Brad Rose, Andrew van de Ven, Stephen Naulls, Shane Robinson, and
others in and around the Structures Lab for their support and technical assistance. I also
thank the staff of Civil, Environmental and Mining Engineering (CEME) Administration,
graduate research coordinators, and staff of Graduate Research School (GRS) for their
valuable guidance and support.

Thanks are extended to all the Master’s students who assisted me in the Structures Lab in
various projects. I gratefully acknowledge the support provided by Adelaide Brighton
Cement, Simcoa Microsilica Pty Ltd., Mapei Australia, and Sika Australia.

To my parents, son, brother, parents-in-law, siblings-in-law, and all my family members:


thank you and I love you for your continuous encouragement and support. Last but not
least, to my wife, this journey would have been difficult without your love and support, your
proofreading and excellent cooking skills, and of course, your bearing with me in hard times,
thank you and I love you.

v
AUTHORSHIP DECLARATION:
CO-AUTHORED PUBLICATIONS

This thesis contains work that has been published and/or prepared for publication.

Chapter 2
Development of ECO-UHPC with very-low-C3A cement and ground granulated blast-furnace slag

Status: published in Construction and Building Materials

Student contribution to work: 85%

The candidate planned the work. He conducted all the laboratory tests with the assistance of
master’s students. The candidate carried out all the microstructure analyses. He analysed all the
experimental data, prepared figures, wrote the original draft of the manuscript.

Co-author signatures and dates:

28/10/2021 28/10/2021 28/10/2021 29/10/2021


Mohamed Elchalakani Ali Karrech M.S. Mohamed Ali Lanhui Guo

Chapter 3
Mechanical properties and chloride penetration resistances of very-low-C3A cement based SC-
UHP-SFRCs incorporating metakaolin and slag

Status: under review

Student contribution to work: 85%

The candidate planned the work. He conducted all the laboratory tests with the assistance of
master’s students. The candidate carried out all the microstructure analyses. He analysed all the
experimental data, prepared figures, wrote the original draft of the manuscript.

Co-author signatures and dates:

28/10/2021 28/10/2021 28/10/2021 29/10/2021


Mohamed Elchalakani Ali Karrech Riyadh Al-Ameri Bo Yang

vi
Chapter 4
ECO-UHPC with high-volume Class-F fly ash: new insight into mechanical and durability
properties

Status: published in Journal of Materials in Civil Engineering

Student contribution to work: 85%

The candidate planned the work. He conducted all the laboratory tests with the assistance of
master’s students. The candidate carried out the microstructure analyses with the support of the
fourth author. He analysed all the experimental data, prepared figures, wrote the original draft of
the manuscript.

Co-author signatures and dates:

28/10/2021 28/10/2021 28/10/2021 28/10/2021 29/10/2021


M.S. Mohamed
Mohamed Elchalakani Ali Karrech Minhao Dong Hua Yang
Ali

Chapter 5
Development of ECO-UHPC utilizing gold mine tailings as quartz sand alternative

Status: published in Cleaner Engineering and Technology

Student contribution to work: 85%

The candidate planned the work. He conducted all the laboratory tests. The candidate carried out
all the microstructure analyses. He analysed all the experimental data, prepared figures, wrote the
original draft of the manuscript.

Co-author signatures and dates:

28/10/2021 29/10/2021 28/10/2021 28/10/2021 29/10/2021


Ehsan
Mohamed Elchalakani Hakan Basarir Ali Karrech Bo Yang
Sadrossadat

vii
Appendix A
Shrinkage of UHPC incorporating very-low-C3A cement and ground granulated blast-furnace slag

Status: published in Proceedings of Concrete 2021, The 30th Biennial National Conference of the
Concrete Institute of Australia

Student contribution to work: 85%

The candidate planned the work. He conducted all the laboratory tests. The candidate carried out
all the microstructure analyses. He analysed all the experimental data, prepared figures, wrote the
original draft of the manuscript.

Co-author signatures and dates:

28/10/2021 28/10/2021 29/10/2021 29/10/2021


Mohamed
Ali Karrech Waleed Nawaz Huiyuan Liu
Elchalakani

Student signature:

Date: 30/10/2021

I, Mohamed Elchalakani, certify that the student’s statements regarding the contribution to each
of the works listed above are correct.

Coordinating supervisor signature:

Date: 30/10/2021

viii
NOTES ON VARIATIONS TO PUBLICATION
TEXTS

Figure, table, section, and equation numbers have been reformatted to be consistent with
the thesis structure.

Referencing styles have been reformatted in some chapters to maintain a consistent style.

Minor changes have been made in a few places in the text, without any significant alteration,
to improve the coherence between the chapters.

The final published version of the paper which is still under review (Chapter 3) may slightly
differ from the version included in this thesis based on the reviewers’ comments.

The published version of Appendix A contains a typographical error, which has been fixed
in the version included in this thesis (the unit in Equation A.1 was changed from µm/mm
to µm/m).

ix
CONTENTS

THESIS DECLARATION ........................................................................................... ii

ABSTRACT ................................................................................................................ iii

ACKNOWLEDGEMENTS ......................................................................................... v

AUTHORSHIP DECLARATION: CO-AUTHORED PUBLICATIONS ................... vi

NOTES ON VARIATIONS TO PUBLICATION TEXTS ........................................... ix

CONTENTS ................................................................................................................ x

LIST OF FIGURES .................................................................................................. xiv

LIST OF TABLES................................................................................................... xviii

CHAPTER 1: Introduction .......................................................................................... 1

1.1 Research motivation ............................................................................................. 1


1.2 Research objectives .............................................................................................. 6
1.3 Research framework ............................................................................................. 7
1.4 Thesis structure .................................................................................................... 8
1.5 Novelty and major contributions ......................................................................... 10
References ............................................................................................................... 10
CHAPTER 2: Development of ECO-UHPC with very-low-C3A cement and ground
granulated blast-furnace slag ...................................................................................... 14

Abstract................................................................................................................... 14
2.1 Introduction ....................................................................................................... 15
2.2 Materials and methods ....................................................................................... 17
2.2.1 Materials ....................................................................................................... 17
2.2.2 UHPC mixes .................................................................................................. 20
2.2.3 Mixing, specimen preparation, and curing ............................................................ 20
2.2.4 Test methods .................................................................................................. 23
2.3 Results and discussion ........................................................................................ 26
2.3.1 Fresh and hardened performances of moderate- and very-low-C3A cement based UHPCs 26
2.3.2 Development of control ECO-UHPC mix ............................................................. 28
2.3.3 Slump flow results of ECO-UHPC mixes.............................................................. 30
2.3.4 Mechanical properties of ECO-UHPC mixes ......................................................... 31
2.3.5 Durability properties of ECO-UHPC mixes .......................................................... 35
2.3.6 XRD analysis ................................................................................................. 38
2.3.7 TG-DTG analyses of ECO-UHPC samples .......................................................... 39
2.4 Conclusions ....................................................................................................... 42

x
Acknowledgements .................................................................................................. 42
References ............................................................................................................... 43
CHAPTER 3: Mechanical properties and chloride penetration resistances of very-low-
C3A cement based SC-UHP-SFRCs incorporating metakaolin and slag ...................... 47

Abstract................................................................................................................... 47
3.1 Introduction ....................................................................................................... 48
3.2 Materials and methods ....................................................................................... 51
3.2.1 Materials ....................................................................................................... 51
3.2.2 Mixes ............................................................................................................ 54
3.2.3 Mixing, specimen preparation, and curing ............................................................ 55
3.2.4 Test methods .................................................................................................. 56
3.3 Results and discussion ........................................................................................ 60
3.3.1 Flowability .................................................................................................... 60
3.3.2 Mechanical properties ....................................................................................... 62
3.3.3 Transport properties ......................................................................................... 66
3.3.4 Microstructure analyses .................................................................................... 71
3.4 Conclusions ....................................................................................................... 73
Acknowledgements .................................................................................................. 75
References ............................................................................................................... 75
CHAPTER 4: ECO-UHPC with high-volume Class-F fly ash: new insight into
mechanical and durability properties .......................................................................... 80

Abstract................................................................................................................... 80
4.1 Introduction ....................................................................................................... 81
4.2 Experimental program ........................................................................................ 83
4.2.1 Materials ....................................................................................................... 83
4.2.2 Mix proportions .............................................................................................. 86
4.2.3 Mixing, specimen preparation, and curing ............................................................ 86
4.2.4 Test methods .................................................................................................. 88
4.3 Results and discussion ........................................................................................ 93
4.3.1 Workability.................................................................................................... 93
4.3.2 Mechanical properties ....................................................................................... 94
4.3.3 Durability properties ...................................................................................... 105
4.3.4 Microstructure analyses .................................................................................. 108
4.3.5 CO2 emission and cost analysis ......................................................................... 109
4.4 Conclusions ..................................................................................................... 112
Acknowledgements ................................................................................................ 114
References ............................................................................................................. 114

xi
CHAPTER 5: Development of ECO-UHPC utilizing gold mine tailings as quartz sand
alternative ................................................................................................................ 121

Graphical abstract .................................................................................................. 121


Abstract................................................................................................................. 122
5.1 Introduction ..................................................................................................... 122
5.2 Materials and methods ..................................................................................... 124
5.2.1 Materials ..................................................................................................... 124
5.2.2 Mix proportions ............................................................................................ 128
5.2.3 Specimen preparation ..................................................................................... 130
5.2.4 Test methods ................................................................................................ 130
5.3 Results and discussion ...................................................................................... 132
5.3.1 Fresh behaviour ............................................................................................ 132
5.3.2 Compressive strength and splitting tensile strength ................................................ 133
5.3.3 Elastic modulus............................................................................................. 135
5.3.4 Durability .................................................................................................... 136
5.3.5 Leaching toxicity ........................................................................................... 137
5.3.6 Microstructure analysis ................................................................................... 139
5.3.7 Economic and environmental evaluation ............................................................ 139
5.4 Conclusion ...................................................................................................... 142
Acknowledgements ................................................................................................ 144
References ............................................................................................................. 144
Concluding remarks .................................................................................................. 148

6.1 Main findings ................................................................................................... 148


6.2 Future research recommendations ..................................................................... 150
APPENDIX A (Conference paper related to Chapter 2) ........................................... 152

Abstract................................................................................................................. 152
A.1 Introduction .................................................................................................... 152
A.2 Experimental program ..................................................................................... 154
A.2.1 Materials .................................................................................................... 154
A.2.2 Mix proportions............................................................................................ 154
A.2.3 Mixing and specimen preparation .................................................................... 154
A.2.4 Test methods................................................................................................ 154
A.3 Results and discussion ..................................................................................... 156
A.3.1 Shrinkage.................................................................................................... 156
A.3.2 Compressive strength ..................................................................................... 157
A.3.3 SEM analysis .............................................................................................. 158

xii
A.4 Conclusions .................................................................................................... 159
Acknowledgements................................................................................................ 160
References............................................................................................................. 160
APPENDIX B (Data related to Chapter 4) ............................................................... 162
APPENDIX C (Miscellaneous) ................................................................................ 169

xiii
LIST OF FIGURES

Figure 1.1. Research framework (VLAC = very-low-C3A cement, XRD = X-ray


diffractometry, TG-DTG = thermogravimetry and derivative thermogravimetry, SEM =
scanning electron microscopy, EDS = energy dispersive spectroscopy) ............................ 7

Figure 2.1. Particle size distributions of binders and sand ............................................. 17

Figure 2.2. XRD patterns of raw binders (G = gypsum, A = alite, B = belite, F = ferrite, At
= tricalcium aluminate, Ak = akermanite, Ge = gehlenite, Q = quartz, SC = silicon
carbide)....................................................................................................................... 19

Figure 2.3. Combined particle size distributions of ECO-UHPCs (MAE and RMSE are
respectively mean absolute error and root-mean-square error between target curve and
combined PSD of each mix)......................................................................................... 22

Figure 2.4. Curing regimes of carbonation test specimens .............................................. 23

Figure 2.5. Setup for initial rate of absorption test ......................................................... 24

Figure 2.6. Setup for half-cell potential test ................................................................... 25

Figure 2.7. Performances of M-SF0WB181BA139, SF0WB181BA139, and SF0WB166BA139 ...... 26

Figure 2.8. Development of hydration products in M-SF0WB181BA139, SF0WB181BA139, and


SF0WB166BA139 (sand particles are not shown for keeping demonstration simple) ............ 27

Figure 2.9. Performances of UHPCs with different silica fume contents ......................... 29

Figure 2.10. Performances of UHPCs with different b/a ratios....................................... 29

Figure 2.11. Slump flow results of ECO-UHPC mixes................................................... 30

Figure 2.12. Compressive strength gains of ECO-UHPCs with time............................... 31

Figure 2.13. (a) 28-day stress – strain curves of ECO-UHPCs; (b) 28-day elastic moduli of
ECO-UHPCs .............................................................................................................. 32

Figure 2.14. (a) 28-day splitting tensile strengths of ECO-UHPCs; (b) 28-day flexural
strengths of ECO-UHPCs ............................................................................................ 32

Figure 2.15. Comparison of (a) elastic modulus, (b) splitting tensile strength, and (c)
flexural strength of UHPC incorporating GGBS with model predictions (points in grey
circles represent data from UHPCs incorporating GGBS) .............................................. 35

Figure 2.16. (a) Water absorption, and (b) initial rate of absorption test results of ECO-
UHPCs ....................................................................................................................... 35

Figure 2.17. Half-cell potential test results of ECO-UHPCs ........................................... 37

Figure 2.18. Illustration of carbonation resistances of specimens (a) S0, (b) S30, and (c)
S60 after 28 of days exposure to curing regime A (300 ppm CO2), followed by 28 days of
exposure to curing regime B (10,000 ppm CO2) ............................................................. 37

xiv
Figure 2.19. XRD patterns of P-M-SF0WB181, P-SF0WB181 pastes at (a) 20 hours, and (b) 28
days (A = alite, B = belite, E = ettringite, P = portlandite) ............................................. 38

Figure 2.20. XRD patterns of P-S0, P-S15, and P-S60 pastes at 28 days (A = alite, B =
belite, E = ettringite, P = portlandite) ........................................................................... 40

Figure 2.21. (a) TG, and (b) DTG results of S0, S15, and S60 ........................................ 41

Figure 3.1. XRD patterns of as received binders [G = gypsum, A = alite, B = belite, F =


ferrite, Q = quartz, SC = silicon carbide, At = anatase, Ak = akermanite, Ge = gehlenite]
................................................................................................................................... 52

Figure 3.2. Particle size distributions of materials [d50 = median diameter] ..................... 52

Figure 3.3. Steel fibres used in this study....................................................................... 53

Figure 3.4. Combined particle size distributions (PSDs) of UHPCs [RMSE = root-mean-
square error between target curve and combined PSD] .................................................. 54

Figure 3.5. Specimen setup during wetting with 10% NaCl solution ............................... 57

Figure 3.6. Temperature and relative humidity during drying phase ............................... 57

Figure 3.7. Capillary water absorption test setup ........................................................... 59

Figure 3.8. Drilling depths of UHP-SFRC cubes for chloride penetration test ................. 59

Figure 3.9. Chloride concentration measurement for UHP-SFRC samples ..................... 60

Figure 3.10. Slump flow and relative sump flow results of UHPCs and UHP-SFRCs ...... 61

Figure 3.11. Typical spreads of (a) R-F2, (b) MK-F2, and (c) S-F2 ................................. 61

Figure 3.12. Compressive strength results of UHPCs and UHP-SFRCs .......................... 62

Figure 3.13. Load – deflection curves of (a) R-F0, MK-F0, S-F0, (b) R-F1, (c) R-F2, (d)
MK-F2, and (e) S-F2 ................................................................................................... 65

Figure 3.14. Water transport properties of UHP-SFRCs – (a) capillary water absorption,
(b) initial sorptivity (Si) in mm/s0.5, (c) secondary sorptivity (Ss) in mm/s0.5, and (d) water
absorption ................................................................................................................... 66

Figure 3.15. Chloride penetration test results of (a) R-F1, (b) R-F2, (c) MK-F2, (d) S-F2,
and apparent diffusivities and penetration resistances of (e) R-F1, R-F2, MK-F2, S-F2 ... 69

Figure 3.16. MK-F2 specimen after 216 d of exposure to cyclic drying and wetting with
10% NaCl solution ...................................................................................................... 70

Figure 3.17. SEM images of (a) R-F2, (b) MK-F2, and (c) S-F2 ..................................... 72

Figure 3.18. XRD patterns of R-F2, MK-F2, and S-F2 powders, collected from a depth of
7.5 mm, after 216 d of exposure to cyclic drying and wetting with 10% NaCl solution [A =
alite, B = belite, E = ettringite, Fs = Friedel’s salt, P = portlandite, Q = quartz] .............. 73

Figure 4.1. XRD patterns of raw binders (G = gypsum, A = alite, B = belite, F = ferrite, M
= mullite, H = hatrurite, Q = quartz, S = silicon carbide) .............................................. 84

xv
Figure 4.2. Particle size distributions of the binders and aggregates ................................ 85

Figure 4.3. Combined particle size distributions of F-UHPC mixes ................................ 87

Figure 4.4. Preparation of a typical half-cell potential test specimen ............................... 88

Figure 4.5. Curing environments of carbonation test specimens ..................................... 89

Figure 4.6. Four-point bending test setup ...................................................................... 90

Figure 4.7. Initial rate of absorption test setup ............................................................... 91

Figure 4.8. Half-cell potential test setup ........................................................................ 91

Figure 4.9. (a) Slump flow results of the ECO-UHPC mixes, (b) typical slump flow of F0,
and (c) typical slump flow of F60 ................................................................................. 93

Figure 4.10. 28-day mechanical test results of the ECO-UHPCs: (a) compressive strength,
(b) elastic modulus, (c) splitting tensile strength, and (d) flexural strength ....................... 94

Figure 4.11. Compressive strength gains of F-UHPCs with time .................................... 95

Figure 4.12. Relationship between 28-day compressive strength and reactivity moduli for
standard-cured UHPC with fly ash ............................................................................... 99

Figure 4.13. Relationship between elastic modulus and compressive strength of standard-
cured UHPC (points in black circles represent data from cement – silica fume – fly ash
based UHPCs) .......................................................................................................... 100

Figure 4.14. Relationship between splitting tensile strength and compressive strength of
standard-cured UHPC (points in black circles represent data from cement – silica fume –
fly ash based UHPCs) ................................................................................................ 103

Figure 4.15. Relationship between flexural strength and compressive strength of standard-
cured UHPC ............................................................................................................. 104

Figure 4.16. (a) Water absorption, and (b) initial rate of absorption test results of F-UHPCs
at 7 and 28 days ........................................................................................................ 106

Figure 4.17. Half-cell potential test results of F-UHPCs ............................................... 106

Figure 4.18. Illustration of carbonation resistances of specimens (a) F0, (b) F40, and (c)
F70 after 28 days of exposure to Environment A (300 ppm CO2), followed by 28 days of
exposure to Environment B (10,000 ppm CO2) ........................................................... 108

Figure 4.19. XRD patterns of F0, F40 and F70 (magnified intensity – left) (E = ettringite, P
= portlandite) ............................................................................................................ 109

Figure 4.20. SEM images and EDS results of F0, F40 and F70 (C = unreacted cement, F =
unreacted fly ash, H = partially reacted fly ash) ........................................................... 110

Figure 4.21. (a) CO2 footprints, and (b) CO2 intensities of different types of concrete .... 111

Figure 4.22. Costs of different types of concrete (A$ = Australian dollars) .................... 112

xvi
Figure 5.1. XRD pattern and scanning electron micrograph of WAGT1; [M = muscovite
(KAl3Si3O10(F,OH)2), A = albite (NaAlSi3O8), P = polyhalite (Ca2H4K2MgO18S4), H =
halite (NaCl), Q = quartz (SiO2)]................................................................................ 127

Figure 5.2. Particle size distributions of solid materials ................................................ 128

Figure 5.3. Combined particle size distributions of UHPC mixes (RMSE = root-mean-
square error, R2 = determination coefficient) ............................................................... 129

Figure 5.4. Leachate sample preparation for leaching toxicity test – (a) horizontal shaking
of UHPC pieces and leaching liquid, and (b) vacuum filtration of leachate ................... 132

Figure 5.5. (a) Compressive strength results, and (b) 28-day splitting tensile strength results
of UHPCs ................................................................................................................. 134

Figure 5.6. (a) 28-day elastic moduli of UHPCs, (b) Ec as a function of √(fc), and (c) Ec as a
function of √(fc)/φ; (Ec = elastic modulus, fc = compressive strength, φ = wet packing
density) ..................................................................................................................... 135

Figure 5.7. Water absorption and initial rate of absorption results of UHPCs ............... 136

Figure 5.8. Illustration of carbonation resistances of UHPCs ....................................... 137

Figure 5.9. Leaching behaviour of leachable heavy metals from tailings in – (a) low-
strength mortar, (b) normal-strength mortar, and (c) UHPC ........................................ 138

Figure 5.10. Backscattered electron micrographs of (a) T0, and (b) T100; (P = pore, S =
quartz sand, T = WAGT1) ........................................................................................ 140

Figure 5.11. (a) Costs, and (b) CO2 footprints of UHPC mixes ..................................... 141

Figure A.1. Mixing procedure .................................................................................... 156

Figure A.2. Length change measurement using comparator ........................................ 156

Figure A.3. Total shrinkage strains of UHPCs at 7 and 28 days ................................... 158

Figure A.4. Compressive strengths of UHPCs at 7 and 28 days.................................... 158

Figure A.5. SEM images and EDS results of C100 and C70-S30 (M = C-S-H rich matrix, P
= portlandite, S = GGBS, R = reaction products, G = gypsum, E = ettringite) ............. 159

Figure C.1. Compression test specimen at 28 days: (a) before failure, and after failure – (b)
S60 (Chapter 2), (c) S-F2 (Chapter 3), (d) F70 (Chapter 4), (e) T60 (Chapter 5)............. 170

Figure C.2. Flexural strength test specimen at 28 days: (a) before failure, (b) typical UHPC
specimen at peak load (Chapters 2, 3, 4, and 5), (c) UHPFRC specimen (S-F2) at around
peak load (Chapter 3), (d) UHPFRC specimen (S-F2) after peak load (Chapter 3) ........ 171

xvii
LIST OF TABLES

Table 1.1. Comparison between mix proportions and properties of NSC, HSC, and UHPC
..................................................................................................................................... 2

Table 1.2. Review of mechanical and durability properties of standard-cured UHPCs and
UHPFRCs incorporating GGBS or Class-F fly ash ......................................................... 4

Table 2.1. Chemical and physical properties of binders and sand ................................... 18

Table 2.2. Mix proportions of UHPCs and UHPC pastes (by mass) ............................... 21

Table 2.3. Mechanical relationships proposed by standards, and MAEs and RMSEs
between relationships and respective UHPC data .......................................................... 34

Table 2.4. Intensities (counts) of XRD peaks corresponding to AFt, CH, and C3S .......... 40

Table 2.5. TG results of S0, S15, and S60 ..................................................................... 41

Table 3.1. Chemical properties of binders and aggregates .............................................. 53

Table 3.2. Mix proportions of UHPCs and UHP-SFRCs (by mass) ................................ 55

Table 3.3. Flexural behaviours of UHPCs and UHP-SFRCs .......................................... 64

Table 4.1. Review of mechanical properties of standard-cured UHPCs and UHPC pastes
with cement – silica fume – Class-F fly ash based binder system .................................... 82

Table 4.2. Physicochemical properties of binders and aggregates ................................... 85

Table 4.3. Mix proportions .......................................................................................... 87

Table 4.4. 28-day mechanical test results of ECO-UHPCs ............................................. 94

Table 4.5. ANOVA outputs for 28-day mechanical test results ....................................... 95

Table 4.6. Elastic modulus – compressive strength models proposed by standards and
previous studies ......................................................................................................... 100

Table 4.7. Splitting tensile strength – compressive strength and flexural strength –
compressive strength models proposed by standards and previous studies .................... 103

Table 4.8. CO2 footprints and costs of different materials............................................. 112

Table 5.1. A review of mechanical properties of standard-cured structural concretes/


mortars incorporating different types of tailings as sand alternatives ............................. 125

Table 5.2. Chemical and physical properties of solid materials ..................................... 126

Table 5.3. Mix proportions of UHPC mixes................................................................ 129

Table 5.4. Fresh properties of UHPC mixes ................................................................ 132

Table 5.5. Concentrations (mg/L) of heavy metals in leachate samples ........................ 138

Table 5.6. CO2 footprints and costs of materials .......................................................... 141

xviii
Table A.1. Chemical compositions of binders ............................................................. 155

Table A.2. Mix proportions ....................................................................................... 155

Table B.1. Data used from this study for Figures 4.12, 4.13, 4.14, and 4.15 .................. 162

Table B.2. Data collected from studies on standard-cured UHPC with cement – silica fume
or cement – silica fume – fly ash binder system for Figures 4.12, 4.13, 4.14, and 4.15.... 162

Table B.3. Data collected from studies in the literature for Figures 4.13, 4.14, and 4.15 163

Table C.1. Constituents of UHPCs and UHPC pastes, presented in Chapter 2, in kg/m3
(based on a target density of 2425 kg/m3) ................................................................... 169

Table C.2. Constituents of UHPCs and UHPFRCs, presented in Chapter 3, in kg/m3


(based on a target density of 2425 kg/m3 for UHPC, 2450 kg/m3 for UHPFRC with 1%
steel fibre content, and 2500 kg/m3 for UHPFRC with 2% steel fibre content) .............. 170

xix
CHAPTER
1
Introduction

1.1 Research motivation

Concrete is the most widely used building material, and the second most consumed material
worldwide after water (Makul 2020). The practice of using concrete as a construction
material dates back to Roman times. Gypsum, quick lime, and volcanic ash were used as
binders, and brick and tuff pieces were used as aggregates in ancient Roman concrete (opus
caementicium) (Bajaber and Hakeem 2021; Giavarini et al. 2006). According to Giavarini
et al. (2006), the compressive strength of a good Roman concrete can be estimated to be 5–
6 MPa. Modern-day structural concretes, based on their compressive strengths, can be
categorized into three major classes: normal-strength concrete, high-strength concrete, and
ultra-high performance concrete (UHPC). Amongst these three types, UHPC is the most
novel one. Table 1.1 presents a comparison between the three types of structural concrete
based on their mix constituents and proportions, and mechanical and durability properties.
UHPC is mainly characterized by its “ultra-high strength”, which is 120 MPa according to
a group of researchers (Huang et al. 2019), and 150 MPa according to others (Courtial et al.
2013). To attain ultra-high strength, a high volume of cement (700–1100 kg/m3), silica fume,
a very low water to binder ratio (w/b ≤ 0.2), and fine quartz sand (with maximum aggregate
size ≤ 0.6 mm) are used in UHPC. Coarse aggregate is typically avoided. Steel fibres are
often incorporated to meet ductility requirements. Such composition of cementitious mix to
attain ultra-high strength was first proposed by Richard and Cheyrezy (1995, 1994).

If compared with NSC or HSC, UHPC exhibits extraordinary performance in terms of both
strength and durability. Consequently, slenderer but more durable structural members can
be built with UHPC. Fibre-reinforced UHPC, also known as ultra-high performance fibre
reinforced concrete (UHPFRC), can exhibit superior ductility and tensile strength. High
tensile strength of fibre-reinforced UHPC helps structural components made with it attain
substantial shear resistance (Tadros 2020). This feature may allow total elimination of
stirrups from a structural member and thus help reduce the size of the member. Ductal ®,
Cemtec®, DURA®, and SIFCON® are examples of some commercially available fibre-
reinforced UHPCs (Azmee and Shafiq 2018; Li and Wu 2018). Proprietary fibre-reinforced
Chapter 1

UHPCs are used for the construction of bridge systems, such as decks, girders, columns,
cast-in-place connections (Hung et al. 2021; Tadros and Voo 2016). Such UHPCs are also
used for constructing structural and architectural elements of buildings such as curved
panels, facades, floorings, columns, etc. (Yoo and Yoon 2016). UHPC without fibres is
brittle at the material scale. Nevertheless, RC (reinforced concrete) beams built with non-
fibre reinforced UHPC were observed to exhibit ductility indices comparable to or higher
than those of the RC beams built with fibre-reinforced UHPC (Yoo and Yoon 2015).
Commercial non-fibre reinforced UHPCs (e.g., DUCORIT®, UHPG-120®, etc.) are used as
grouting materials to strengthen and repair offshore structures (Mazaheri et al. 2021), to
connect wind turbines to gravity foundations (Jia 2018), etc. Such UHPCs are also used for
the construction of multifunctional walls, acoustic damping wall elements, thin slabs,
architectural elements, etc. (Gosselin et al. 2016; Lin et al. 2021).

Table 1.1. Comparison between mix proportions and properties of NSC, HSC, and UHPC
(Amin et al. 2021; Gu et al. 2015; Scheydt and Müller 2012; Shi et al. 2015; Voo and Foster
2010)

Concrete type NSC HSC UHPC


Constituents Cement ≤400 400–600 700–1100
(kg/m3) Silica fume - 40–50 100–300
Fine aggregate ≈700 500–700 1000–1300
Coarse aggregate ≈1000 800–1000 -
Chemical admixture - ≈5 10–70
Water ≈200 120–200 125–280
Fibre - - 0.5%–3%*
Parameters w/b ≈0.5 ≤0.3 ≤0.2
Maximum aggregate size (mm) 19.0–25.5 9.5–12.5 ≤0.6
Properties Compressive strength (MPa) 20–40 40–100 ≥120 or ≥150
Splitting tensile strength (MPa) 2–3.5 3.5–6 ≥6 or ≥8
Flexural strength (MPa) 2.5–4 4–8 ≥8 or ≥15
Water absorption (%) >3.5 1.5–3 <1.5
Carbonation depth (mm) 5–15 1–5 <0.5
NSC = normal-strength concrete, HSC = high-strength concrete, UHPC = ultra-high performance
concrete, w/b = water to binder ratio
*By volume (for fibre-reinforced UHPC or UHPFRC)

2
Chapter 1

Although several projects worldwide opted for UHPC to construct, repair, or strengthen
structural members, connections, and architectural elements, the use of the material is still
very limited if compared with NSC. Surely, the use of UHPC, in lieu of NSC, enables the
fabrication of slenderer structural components (Voo and Foster 2010). Nevertheless, the
overall material CO2 footprint of a structural component built with UHPC is usually still
higher than that of the component built with NSC (Joe et al. 2017). The high material CO2
footprint of UHPC is a major weakness that is worth addressing before its widespread use
to meet the pressing demand of slowing down global warming. The high cement content of
UHPC is one of the key reasons for its high CO2 footprint (≥700 kg/m3, as presented in
Table 1.1). Every tonne of Portland cement production generates nearly a tonne of CO2
(Carvalho et al. 2018). Moreover, the cement industry alone is responsible for up to 10% of
the annual global anthropogenic emissions, in particular CO2 (Carreño-Gallardo et al.
2018). One of the most practised approaches of reducing CO2 footprints of cementitious
mixes is to replace Portland cement with industrial by-product based binders or, more
commonly known as, supplementary cementitious materials (SCMs) (Elchalakani et al.
2017). Amongst different types of SCMs, ground granulated blast furnace slag (GGBS),
Class-F fly ash [AS/NZS 3582.1 (2016) Grade 1 fly ash], and silica fume are available in
Australia (Temuujin et al. 2013; VicRoads 1999). In proprietary UHPCs, silica fume is
usually used as the only SCM (Azmee and Shafiq 2018). However, silica fume being very
fine, compared to fly ash or GGBS, its inclusion increases the water and superplasticizer
demand of fresh cementitious mix. High volume inclusion of silica fume in a cementitious
mix is difficult as it negatively affects the workability of the mix (Sabet et al. 2013). A number
of studies were carried out in which properties of UHPCs, with/without fibres,
incorporating Class-F fly ash and/or GGBS were investigated (Gupta 2016; Li et al. 2015;
Liu et al. 2018; Mueller et al. 2016; Randl et al. 2014; Shaikh et al. 2018; Song et al. 2018;
Wang et al. 2016; Wu et al. 2017; Yu et al. 2015). Nevertheless, a few studies, as mentioned
in Table 1.2, systematically investigated the effects of replacing Portland cement with GGBS
or Class-F fly ash on the properties of UHPC, incorporating cement – silica fume binder
system and w/b ≤ 0.2, cured in standard curing condition (28 days of curing in water or fog
room at a relative humidity of 95% ± 5% and a temperature of 24 °C ± 6 °C). Most of these
studies focused on examining the compressive and flexural behaviours of UHPCs,
with/without fibres, incorporating GGBS and Class-F fly ash. Only Ganesh and Murthy
(2019) considered investigating some durability aspects namely initial sorptivity and rapid
chloride penetration resistance of fibre-reinforced UHPC with GGBS.

3
Chapter 1

Table 1.2. Review of mechanical and durability properties of standard-cured UHPCs and UHPFRCs incorporating GGBS or Class-F fly ash

SCM Reference Concrete Fibre w/b, Replacement† Mechanical properties Durability properties
type type, Vf sf/b†
GGBS Ganesh and UHPFRC Steel, 0.17, 0%, 20%, Compressive, direct tensile, and indirect tensile (flexural and splitting Initial sorptivity generally
Murthy (2019) 2% 0.13 40%, 60%, tensile) properties were investigated; 28-day compressive strength, reduced (from 0.123 down
(Ganesh and 80% direct tensile strength, flexural strength, and splitting tensile strength to 0.086 mm/s0.5) as GGBS
Murthy 2019) obtained for 20% replacement (120.9, 12.5, 32.5, 20.4 MPa, content was increased;
respectively) were slightly higher than those of the control mix (115.7, chloride permeability,
12, 30, 19.5 MPa, respectively); beyond 20% replacement, 28-day determined through RCPT,
compressive strength, direct tensile strength, flexural strength, and was found to be very low or
splitting tensile strength generally reduced (down to 101.2, 8, 27.3, 15 negligible for UHPFRC
MPa, respectively, at 80%) containing GGBS
Abdulkareem UHPC - 0.14– 0%, 28%, Compressive strengths, flexural strengths, and splitting tensile -
et al. (2018) 0.15, 48%, 79% strengths were reported; 28-day compressive strength obtained for 28%
(Abdulkareem 0.16– replacement (142 MPa) was comparable to that obtained for the
et al. 2018) 0.17 control mix (140 MPa); beyond 28% replacement, 28-day compressive
strength generally reduced (down to 108 MPa at 79%); 28-day flexural
strengths and splitting tensile strengths were in the ranges of 7.7–8.7
MPa and 14–19 MPa, respectively
Yazıcı et al. UHPFRC Steel, 0.13, 0%, 20%, Compressive and flexural behaviours were investigated; 28-day -
(2010)Γ (Yazıcı 3% 0.23 40%, 60% compressive strength obtained for 20% replacement (208 MPa) was
et al. 2010) higher than that obtained for the control mix (199 MPa); beyond 20%
replacement, 28-day compressive strength generally reduced (down to
191 MPa at 60%); 28-day flexural strength increased up to 40%
replacement; flexural strengths were in the range of 25.1–32.1 MPa
Class- Chen et al. UHPC - 0.2, 0%, 10%, Compressive and flexural properties were investigated; maximum 28- -
F fly (2018) (Chen 0.11 20%, 30% day compressive and flexural strengths were obtained for 20%
ash et al. 2018) replacement (126 MPa and 20 MPa, respectively); compressive and
flexural strengths were in the ranges of 114–126 MPa and 16.3–20
MPa, respectively
Peng et al. UHPC - 0.18, 0%, 12%, Compressive behaviour was investigated; compressive strength -
(2010) (Peng 0.15 18%, 24%, generally reduced as cement was replaced with fly ash, 28-day
et al. 2010) 35%, 47% compressive strengths were in the range of 85.8–113.1 MPa
Vf = fibre volume fraction, w/b = water to binder ratio, sf/b = silica fume to binder ratio, RCPT = rapid chloride penetration test
*Cured in lime saturated water
ΓBauxite was used as aggregate

By mass
4
Chapter 1

The production cost of a structural component built with UHPC, like its material CO 2
footprint, is also higher than the production cost of the component built with NSC (Joe et
al. 2017). According to Greybeal (2008), the high fabrication cost of a UHPC component is
the main reason that restricts the widespread application of the novel material. Total
exclusion of coarse aggregate from the material and use of very high-quality fine quartz sand
as the coarsest aggregate is one of the factors contributing to its high cost. In Western
Australia (WA), the mine sites annually produce millions of tonnes of tailings as by-
products. Dams occupying large areas are required for the storage of these tailings. Being
very fine, tailings may need little or no processing for their utilization as UHPC aggregate.
The use of tailings as quartz sand alternative may thus notably reduce the cost of UHPC.
Moreover, the use of tailings may reduce the hazards associated with its mismanagement in
the tailings dam. The CO2 emission associated with the extraction, crushing, and grinding
of quartz sand can also be reduced to some extent if quartz sand in UHPC is replaced with
tailings. Very few studies can be found in the literature which investigated the possibility of
utilizing iron ore tailings and quartz based mine tailings in UHPC as sand alternatives (Pyo
et al. 2018; Zhao et al. 2014; Zhu et al. 2015).

To encourage the application of UHPCs incorporating GGBS and Class-F fly ash as SCMs
and tailings as aggregate, and thus to reduce the CO2 footprint and/or cost of UHPC
components, further understanding of the mechanical properties and, especially, the
durability properties of the cementitious mixes is still required. Currently, there is a lack of
detailed studies in the literature addressing the relationships between the mechanical
properties of standard-cured UHPC incorporating GGBS or Class-F fly ash: such as elastic
modulus – compressive strength relationship, splitting tensile strength – compressive
strength relationship, flexural strength – compressive strength relationship, etc. More
importantly, there is a scarcity of a study addressing the relationships between the
mechanical properties of standard-cured UHPC in general. Relationships between the
mechanical properties can be beneficial tools in the case of structural design. Various design
codes specify relationships to estimate the elastic modulus, splitting tensile strength, and
flexural strength of NSC or HSC from its compressive strength (ACI 318 2014; ACI 363R
1992; AS 3600 2009). However, previous studies indicate that, as the strength of concrete
gets higher, its characteristics and mechanical properties become different (Alsalman et al.
2017; Rashid et al. 2002). An investigation to understand the relationships between various
mechanical properties of standard-cured UHPC is therefore salient. Again, a study exploring
the durability properties of UHPC incorporating GGBS or Class-F fly ash, such as water
absorption, initial rate of absorption, corrosion risk of embedded rebars, carbonation
resistance, resistance to long-term chloride ingress, etc., is scarce in the literature. Hence,

5
Chapter 1

investigating the durability properties of UHPCs incorporating GGBS and Class-F fly ash is
also essential. Most of the mines in WA are gold mines. Nevertheless, there is a scarcity of
a detailed study in the public literature addressing the properties of UHPC incorporating
gold mine tailings as aggregate. It is, therefore, necessary to investigate the properties of
UHPCs incorporating gold mine tailings, sourced from local mines in WA, to understand
the feasibility of using these tailings as UHPC aggregates.

1.2 Research objectives

Based on the research gaps highlighted in the preceding section, the objectives of this thesis
were set as follows:

1. To identify Portland cement type, silica fume to binder ratio, and binder to aggregate
ratio for the control UHPC mix to achieve good workability and ultra-high strength;
2. To investigate the influence of high-volume substitution of cement with GGBS on the
mechanical properties (e.g., compressive strength, elastic modulus, splitting tensile
strength, and flexural strength), durability properties (e.g., water absorption, initial rate
of absorption, corrosion risk of embedded rebars, and carbonation resistance), and
microstructure properties of UHPC;
3. To understand the effects of partially replacing cement with GGBS on the mechanical
properties (e.g, compressive strength, flexural strength, and load-deflection behaviour),
durability properties (e.g., water absorption, sorptivity, and resistance to long-term
chloride ingress), and microstructure properties of fibre-reinforced UHPC (i.e.,
UHPFRC);
4. To investigate the influence of high-volume substitution of cement with Class-F fly ash
on the mechanical properties (e.g., compressive strength, elastic modulus, splitting
tensile strength, and flexural strength), durability properties (e.g., water absorption,
initial rate of absorption, corrosion risk of embedded rebars, and carbonation resistance),
and microstructure properties of UHPC;
5. To develop a relationship for predicting the compressive strength of UHPC
incorporating fly ash from the contents and chemical compositions of the binders used;
to develop compressive strength – elastic modulus, compressive strength – splitting
tensile strength, and compressive strength – flexural strength relationships for standard-
cured UHPC incorporating fly ash and for standard-cured UHPC in general;
6. To explore the feasibility of using gold mine tailings, sourced from a mine site in WA,
as aggregate in UHPC by assessing the mechanical performance, durability
performance, and toxic leaching behaviour of UHPC incorporating the tailings.

6
Chapter 1

1.3 Research framework

The flowchart shown in Figure 1.1 illustrates the research framework adopted in this thesis.
As discussed in Section 1.1, the CO2 footprints and costs of commercial UHPCs are high,
as a high volume of cement and high-quality quartz sand are typically used in such UHPCs
and industrial by-products, apart from silica fume, are hardly used. One of the key reasons
behind the limited use of industrial by-products is the lack of knowledge on the mechanical
properties and especially the durability properties of UHPCs with by-products. In
order to address this concern to a certain extent, and to contribute to the current body
of literature, six research objectives, as mentioned in Section 1.2, were set. The
experimental program was subdivided into four phases to meet all the objectives.

Figure 1.1. Research framework (VLAC = very-low-C3A cement, XRD = X-ray


diffractometry, TG-DTG = thermogravimetry and derivative thermogravimetry, SEM =
scanning electron microscopy, EDS = energy dispersive spectroscopy)

7
Chapter 1

Descriptions of the test methods adopted to complete Phases 1, 2, 3, and 4 of the


experimental campaign are presented in Chapters 2, 3, 4, and 5, respectively. Data obtained
from each phase, additionally, in some cases, data collected from the literature, were
analysed to infer conclusions on the behaviours of UHPCs made with a certain type of by-
product (GGBS, Class-F fly ash, or gold mine tailings). Some future research directions were
also recommended to further enhance the understanding of UHPCs incorporating by-
products.

1.4 Thesis structure

This thesis is structured following the “Thesis as a Series of Papers (TASP)” format specified
by the Graduate Research School of The University of Western Australia. The thesis consists
of 6 chapters, 3 of which were published in journals and 1 is under review for publication.
Each chapter presents literature review, description of materials, and experimental methods
that are pertinent to the chapter.

Chapter 2: The scope of work for this chapter was outlined to meet Research objectives 1
and 2. Chapter 2 compares the performances of two different types of Portland cement,
namely moderate-C3A cement (MAC) and very-low-C3A cement (VLAC). Both the cements
comply with the requirements of ASTM C150 / C150M-16 (2016) Type I Portland cement.
The results presented in the chapter suggest that the latter cement exhibited better
performance in terms of compressive strength and workability. VLAC was therefore selected
for the rest of the work of this thesis. The results of the mixes prepared to identify the
optimum silica fume to binder ratio (for cement – silica fume binder system) and aggregate
to binder ratio are presented in Chapter 2. The identified silica fume to binder ratio and
aggregate to binder ratio were either kept similar or slightly changed (to meet strength or
workability requirement) in the control mixes prepared for the remaining research work.
The chapter also includes a discussion of the mechanical, durability, and microstructure test
results of UHPCs incorporating different GGBS contents. Chapter 2 further presents a
comparison of elastic modulus, splitting tensile strength, and flexural strength test results
with the predictions based on standard models for NSC and HSC (ACI 318 2014; ACI 363R
1992; AS 3600 2009). The comparison indicates that separate models are needed for UHPC
(especially, splitting tensile strength – compressive strength and flexural strength –
compressive strength models).

Chapter 3: This chapter presents the results obtained from the experimental investigation
carried out to address Research objective 3. As discussed above, the results presented in
Chapter 2 indicate that the use of VLAC may influence the workability and strength of

8
Chapter 1

UHPC positively. Several previous studies, however, suggest that the use of VLAC may
impair the chloride binding capacity of a cementitious mix (Liu et al. 2021; Oh et al. 2003).
In the marine environment, penetration of chloride ions into a structural component made
with UHPFRC beyond a threshold limit may result in fibre and/or rebar corrosion.
Therefore, Chapter 3 presents an investigation on the resistance of VLAC based UHPFRC,
with/without GGBS, to long-term chloride penetration. The chapter also includes a
discussion of the mechanical and microstructure test results of VLAC based UHPFRC
with/without GGBS.

Chapter 4: The mechanical, durability, and microstructure test results of UHPCs made by
replacing cement (VLAC) with Class-F fly ash at different ratios are presented in Chapter 4,
in order to address Research objective 4. Results presented in Chapter 2, as discussed above,
indicate that exclusive relationships are required to predict the elastic modulus, and
especially the splitting tensile strength and flexural strength of UHPC from its compressive
strength. To address this concern and to meet Research objective 5, the relationships for
predicting elastic modulus, splitting tensile strength, and flexural strength of standard-cured
UHPC with/without fly ash from the compressive strength of the cementitious mix are
presented in Chapter 4. A relationship for predicting the compressive strength of UHPC
incorporating fly ash from the contents and chemical compositions of the binders is also
presented in Chapter 4. Chapter 4 additionally presents a comparative discussion on the CO2
footprints and costs of conventional HSC and UHPCs incorporating fly ash (prepared in this
study and several previous studies). The results suggest that the inclusion of fly ash can
substantially reduce the material CO2 footprint of UHPC. The cost reduction, on the other
hand, may not be so pronounced. Nevertheless, the cost calculated for the UHPC mix with
fly ash and river sand reported in (Song et al. 2018) indicates that the use of nonconventional
low-cost material as aggregate in UHPC may be beneficial for its cost reduction.

Chapter 5: This chapter presents an investigation on the mechanical, durability, and


microstructure properties of UHPC mixes incorporating gold mine tailings as aggregate, to
meet Research objective 6. Chapter 5 also presents the CO2 footprints and costs of UHPCs
with tailings and reaffirms the conclusion drawn in Chapter 4 that the use of low-cost
material as aggregate in UHPC notably reduces the cost of the cementitious mix.

Chapter 6: This chapter presents the main findings that can be drawn from the scope of this
thesis. Some future research directions are also recommended in this chapter.

9
Chapter 1

1.5 Novelty and major contributions

The novel and original contribution from Chapter 2 of this thesis is an experimental
investigation on the mechanical, durability, and microstructure properties of UHPCs made
with different VLAC to GGBS ratios. An investigation on the mechanical, durability
(especially resistance to long-term chloride penetration), and microstructure properties of
UHPFRCs incorporating VLAC – silica fume – GGBS and VLAC – silica fume –
metakaolin binder systems is the major contribution from Chapter 3. Some of the original
contributions from Chapter 4 are: an experimental investigation on the mechanical,
durability, and microstructure properties of UHPCs made with high-volume replacement of
cement with Class-F fly ash (up to 70%); a relationship to predict the compressive strength
of standard-cured UHPC with fly ash from the chemical compositions of the binders; and
relationships to predict different mechanical properties from the compressive strength of
standard-cured UHPC. The novel and major contribution from Chapter 5 is an investigation
on the mechanical, durability, and toxic leaching behaviours of UHPCs made with gold
mine tailings as quartz sand substitute.

References
Abdulkareem, O. M., Ben Fraj, A., Bouasker, M., and Khelidj, A. (2018). “Effect of
chemical and thermal activation on the microstructural and mechanical properties of
more sustainable UHPC.” Construction and Building Materials, 169, 567–577.
ACI 318. (2014). Building Code Requirements for Structural Concrete. Farmington Hills, MI.
ACI 363R. (1992). State-of-the-Art Report on High-Strength Concrete.
Alsalman, A., Dang, C. N., Prinz, G. S., and Hale, W. M. (2017). “Evaluation of modulus
of elasticity of ultra-high performance concrete.” Construction and Building Materials, 153,
918–928.
Amin, M., Zeyad, A. M., Tayeh, B. A., and Saad Agwa, I. (2021). “Engineering properties
of self-cured normal and high strength concrete produced using polyethylene glycol and
porous ceramic waste as coarse aggregate.” Construction and Building Materials, 299.
AS/NZS 3582.1. (2016). Supplementary cementitious materials for use with portland and blended
cement - Fly ash.
AS 3600. (2009). Concrete structures.
ASTM C150 / C150M-16. (2016). Standard specification for portland cement. West
Conshohocken, PA.
Azmee, N. M., and Shafiq, N. (2018). “Ultra-high performance concrete: From
fundamental to applications.” Case Studies in Construction Materials, 9.
Bajaber, M. A., and Hakeem, I. Y. (2021). “UHPC evolution, development, and utilization
in construction: A review.” Journal of Materials Research and Technology, 10, 1058–1074.
Carreño-Gallardo, C., Tejeda-Ochoa, A., Perez-Ordonez, O. I., Ledezma-Sillas, J. E.,
Lardizabal-Gutierrez, D., Prieto-Gomez, C., Valenzuela-Grado, J. A., Robles
Hernandez, F. C., and Herrera-Ramirez, J. M. (2018). “In the CO2 emission

10
Chapter 1

remediation by means of alternative geopolymers as substitutes for cements.” Journal of


Environmental Chemical Engineering, 6(4), 4878–4884.
Carvalho, S. Z., Vernilli, F., Almeida, B., Oliveira, M. D., and Silva, S. N. (2018).
“Reducing environmental impacts: The use of basic oxygen furnace slag in portland
cement.” Journal of Cleaner Production, 172, 385–390.
Chen, T., Gao, X., and Ren, M. (2018). “Effects of autoclave curing and fly ash on
mechanical properties of ultra-high performance concrete.” Construction and Building
Materials, 158, 864–872.
Courtial, M., De Noirfontaine, M. N., Dunstetter, F., Signes-Frehel, M., Mounanga, P.,
Cherkaoui, K., and Khelidj, A. (2013). “Effect of polycarboxylate and crushed quartz in
UHPC: microstructural investigation.” Construction and Building Materials, 44, 699–705.
Elchalakani, M., Basarir, H., and Karrech, A. (2017). “Green concrete with high-volume fly
ash and slag with recycled aggregate and recycled water to build future sustainable
cities.” Journal of Materials in Civil Engineering, 29(2), 04016219.
Ganesh, P., and Murthy, A. R. (2019). “Tensile behaviour and durability aspects of
sustainable ultra-high performance concrete incorporated with GGBS as cementitious
material.” Construction and Building Materials, 197, 667–680.
Giavarini, C., Ferretti, A. S., and Santarelli, M. L. (2006). “Mechanical Characteristics of
Roman ‘Opus Caementicium.’” Fracture and failure of natural building stones: applications
in the restoration of ancient monuments, 107–120.
Gosselin, C., Duballet, R., Roux, P., Gaudillière, N., Dirrenberger, J., and Morel, P. (2016).
“Large-scale 3D printing of ultra-high performance concrete - a new processing route for
architects and builders.” Materials and Design, 100, 102–109.
Graybeal, B. (2008). “UHPC in the U.S. Highway Transportation System.” 2nd International
Symposium on Ultra High Performance Concrete, E. Fehling, M. Schmidt, and S. Stürwald,
eds., Kassel university press GmbH, Kassel, 11–18.
Graybeal, B., Brühwiler, E., Kim, B.-S., Toutlemonde, F., Voo, Y. L., and Zaghi, A. (2020).
“International Perspective on UHPC in Bridge Engineering.” Journal of Bridge
Engineering, 25(11), 04020094.
Gu, C. P., Ye, G., and Sun, W. (2015). “Ultrahigh performance concrete-properties,
applications and perspectives.” Science China Technological Sciences, 58(4), 587–599.
Gupta, S. (2016). “Effect of content and fineness of slag as high volume cement replacement
on strength and durability of ultra-high performance mortar.” J. Build. Mater. Struct., 3,
43–54.
Huang, H., Gao, X., and Jia, D. (2019). “Effects of rheological performance, antifoaming
admixture, and mixing procedure on air bubbles and strength of UHPC.” Journal of
Materials in Civil Engineering, 31(4), 04019016.
Hung, C.-C., El-Tawil, S., and Chao, S.-H. (2021). “A Review of Developments and
Challenges for UHPC in Structural Engineering: Behavior, Analysis, and Design.”
Journal of Structural Engineering, 147(9), 03121001.
Jia, J. (2018). “Grout Connections.” Soil Dynamics and Foundation Modeling, 625–628.
Joe, C. D., Moustafa, M. A., and Ryan, K. L. (2017). Cost and Ecological Feasibility of using
UHPC in Highway Bridges. Carson City, NV.
Li, J., and Wu, C. (2018). “Damage evaluation of ultra-high performance concrete columns
after blast loads.” International Journal of Protective Structures, 9(1), 44–64.
Li, W., Huang, Z., Cao, F., Sun, Z., and Shah, S. P. (2015). “Effects of nano-silica and
nano-limestone on flowability and mechanical properties of ultra-high-performance

11
Chapter 1

concrete matrix.” Construction and Building Materials, 95, 366–374.


Lin, W. T., Zhao, W. Q., Chang, Y. H., Yang, J. S., and Cheng, A. (2021). “The effect of
incorporating ultra-fine spherical particles on rheology and engineering properties of
commercial ultra-high-performance grout.” Crystals, 11(9).
Liu, H., Zhang, Y., Liu, J., Kong, S., and Feng, Z. (2021). “Chloride Binding Capacity of
Portland Cement System with Different Content of Tricalcium Aluminate.” IOP
Conference Series: Earth and Environmental Science, 719(2).
Liu, K., Yu, R., Shui, Z., Li, X., Ling, X., He, W., Yi, S., and Wu, S. (2018). “Effects of
Pumice-Based Porous Material on Hydration Characteristics and Persistent Shrinkage
of Ultra-High Performance Concrete (UHPC).” Materials, 12(1), 11.
Makul, N. (2020). “Advanced smart concrete - A review of current progress, benefits and
challenges.” Journal of Cleaner Production, 274.
Mazaheri, P., Asgarian, B., and Gholami, H. (2021). “Assessment of strengthening,
modification, and repair techniques for aging fixed offshore steel platforms.” Applied
Ocean Research, 110.
Mueller, U., Williams Portal, N., Chozas, V., Flansbjer, M., Larazza, I., da Silva, N., and
Malaga, K. (2016). “Reactive powder concrete for façade elements – A sustainable
approach.” Journal of Facade Design and Engineering, 4(1–2), 53–66.
Oh, B. H., Jang, S. Y., and Shin, Y. S. (2003). “Experimental investigation of the threshold
chloride concentration for corrosion initiation in reinforced concrete structures.”
Magazine of Concrete Research, 55(2), 117–124.
Peng, Y., Hu, S., and Ding, Q. (2010). “Preparation of reactive powder concrete using fly
ash and steel slag powder.” Journal Wuhan University of Technology, Materials Science
Edition, 25(2), 349–354.
Pyo, S., Tafesse, M., Kim, B. J., and Kim, H. K. (2018). “Effects of quartz-based mine
tailings on characteristics and leaching behavior of ultra-high performance concrete.”
Construction and Building Materials, 166, 110–117.
Randl, N., Steiner, T., Ofner, S., Baumgartner, E., and Mészöly, T. (2014). “Development
of UHPC mixtures from an ecological point of view.” Construction and Building Materials,
67(PART C), 373–378.
Rashid, M. A., Mansur, M. A., and Paramasivam, P. (2002). “Correlations between
Mechanical Properties of High-Strength Concrete.” Journal of Materials in Civil
Engineering, 14(3), 230–238.
Richard, P., and Cheyrezy, M. (1995). “Composition of reactive powder concretes.” Cement
and Concrete Research, 25(7), 1501–1511.
Richard, P., and Cheyrezy, M. H. (1994). “Reactive powder concretes with high ductility
and 200-800 mpa compressive strength.” American Concrete Institute, ACI Special
Publication, SP-144, 507–518.
Sabet, F. A., Libre, N. A., and Shekarchi, M. (2013). “Mechanical and durability properties
of self consolidating high performance concrete incorporating natural zeolite, silica fume
and fly ash.” Construction and Building Materials, 44, 175–184.
Scheydt, J., and Müller, H. (2012). “Microstructure of ultra high performance concrete
(UHPC) and its impact on durability.” 3rd International Symposium on on Ultra High
Performance Concrete, 349–356.
Shaikh, F. U. A., Nishiwaki, T., and Kwon, S. (2018). “Effect of fly ash on tensile properties
of ultra-high performance fiber reinforced cementitious composites (UHP-FRCC).”
Journal of Sustainable Cement-Based Materials, 7(6), 357–371.

12
Chapter 1

Shi, C., Wu, Z., Xiao, J., Wang, D., Huang, Z., and Fang, Z. (2015). “A review on ultra
high performance concrete: Part I. Raw materials and mixture design.” Construction and
Building Materials, 101, 741–751.
Song, Q., Yu, R., Shui, Z., Wang, X., Rao, S., and Lin, Z. (2018). “Optimization of fibre
orientation and distribution for a sustainable Ultra-High Performance Fibre Reinforced
Concrete (UHPFRC): Experiments and mechanism analysis.” Construction and Building
Materials, 169, 8–19.
Tadros, M. K. (2020). “Ultra-High-Performance Concrete-A Game Changer in Structural
Engineering.” Current Trends in Civil & Structural Engineering, 5(1).
Tadros, M. K., and Voo, Y. L. (2016). “Taking Ultra-High-Performance Concrete to New
Heights: The Malaysian Experience.” Aspire The Concrete Bridge Magazine, 10(3), 36–38.
Temuujin, J., Rickard, W., and Van Riessen, A. (2013). “Characterization of various fly
ashes for preparation of geopolymers with advanced applications.” Advanced Powder
Technology, 24(2), 495–498.
VicRoads. (1999). Supplementary Cementitious Materials (SCMs) in Concrete. Victoria,
Australia.
Voo, Y. L., and Foster, S. J. (2010). “Characteristics of ultra-high performance ‘ductile’
concrete and its impact on sustainable construction.” IES Journal Part A: Civil and
Structural Engineering, 3(3), 168–187.
Wang, D., Shi, C., Wu, Z., Wu, L., Xiang, S., and Pan, X. (2016). “Effects of nanomaterials
on hardening of cement–silica fume–fly ash-based ultra-high-strength concrete.”
Advances in Cement Research, 28(9), 555–566.
Wu, Z., Shi, C., and He, W. (2017). “Comparative study on flexural properties of ultra-high
performance concrete with supplementary cementitious materials under different curing
regimes.” Construction and Building Materials, 136, 307–313.
Yazıcı, H., Yardimci, M. Y., Yiǧiter, H., Aydin, S., and Türkel, S. (2010). “Mechanical
properties of reactive powder concrete containing high volumes of ground granulated
blast furnace slag.” Cement and Concrete Composites, 32(8), 639–648.
Yoo, D. Y., and Yoon, Y. S. (2015). “Structural performance of ultra-high-performance
concrete beams with different steel fibers.” Engineering Structures, 102, 409–423.
Yoo, D. Y., and Yoon, Y. S. (2016). “A Review on Structural Behavior, Design, and
Application of Ultra-High-Performance Fiber-Reinforced Concrete.” International
Journal of Concrete Structures and Materials, 10(2), 125–142.
Yu, R., Spiesz, P., and Brouwers, H. J. H. (2015). “Development of an eco-friendly Ultra-
High Performance Concrete (UHPC) with efficient cement and mineral admixtures
uses.” Cement and Concrete Composites, 55, 383–394.
Zhao, S., Fan, J., and Sun, W. (2014). “Utilization of iron ore tailings as fine aggregate in
ultra-high performance concrete.” Construction and Building Materials, 50, 540–548.
Zhu, Z., Li, B., and Zhou, M. (2015). “The influences of iron ore tailings as fine aggregate
on the strength of ultra-high performance concrete.” Advances in Materials Science and
Engineering, 2015, 412878.

13
CHAPTER
2
Development of ECO-UHPC with very-low-
C3A cement and ground granulated blast-
furnace slag

Tanvir Ahmeda, Mohamed Elchalakania, Ali Karrecha, M.S. Mohamed Alib, Lanhui Guoc,*
a
Department of Civil, Environmental and Mining Engineering, Faculty of Engineering, Computing and Mathematical
Sciences, The University of Western Australia, 35 Stirling Highway, Crawley, WA 6009, Australia
b
School of Civil, Environmental and Mining Engineering, The University of Adelaide, Adelaide, SA 5005, Australia
c
School of Civil Engineering, Harbin Institute of Technology, Harbin 150090, China

*Corresponding author: guolanhui@hit.edu.cn

Abstract

Substituting cement by supplementary cementitious material (SCM) is the most-practiced


approach of reducing CO2 footprint of concrete. However, in case of UHPC, for which ultra-
high performance in terms of both strength and durability is of utmost importance, high-
volume reduction of cement is often difficult or highly challenging. In this study, the
possibility of high-volume incorporation of ground granulated blast-furnace slag (GGBS) in
UHPC with very-low-C3A Portland cement (having 1.1% tricalcium aluminate or C3A) as
primary binder has been explored. Benefits of using very-low-C3A cement as primary binder
in UHPC, over cement with a moderate C3A content (9.3%), in terms of strength and
workability have also been investigated. Results of the UHPC mixes incorporating GGBS
suggest that 60% replacement of very-low-C3A cement by GGBS reduces the 28-day
compressive strength of UHPC by 16.1% with respect to the strength of UHPC without
GGBS. Nevertheless, ultra-high strength (>150 MPa) can still be achieved up to 60%
replacement of cement by GGBS, without the need for any special curing or fibres. The
water absorption and initial rate of absorption of UHPC, and the corrosion risk of rebar
embedded in UHPC reduce with the increase of very-low-C3A cement replacement by
GGBS. Up to 60% replacement of cement by GGBS, UHPC exhibits carbonation resistance
similar to that of the UHPC without GGBS.

Construction and Building Materials, 284, 122787


(https://doi.org/10.1016/j.conbuildmat.2021.122787)
Chapter 2

Keywords: carbonation, durability, ettringite, half-cell potential, high-volume SCM,


tricalcium aluminate, ultra-high performance concrete

2.1 Introduction

Ultra-high performance concrete (UHPC) is a relatively new class of concrete exhibiting


outstanding mechanical performance in comparison to normal-strength concrete (NSC) and
high-strength concrete (HSC). UHPC can exhibit compressive strength exceeding 150 MPa,
and elastic modulus and flexural strength exceeding 43 GPa and 15 MPa, respectively
(Ahmed et al. 2021; Soliman and Tagnit-Hamou 2017). In order to attain such ultra-high
strength, large volume of cement (800–1200 kg/m3) and a very low water to binder ratio
(≤0.2) are used for its production. Special curing techniques (e.g., steam-curing, heat-curing,
autoclaving, etc.) are sometimes also adopted to achieve ultra-high strength. It is, however,
worth mentioning that availing such techniques is difficult for cast-in-situ constructions.
UHPC is a brittle material. Nevertheless, reinforced concrete structural members built with
UHPC may achieve substantial ductility (Yoo and Yoon 2015). The ductility as well as the
strength of UHPC can be improved at material level by adding high-quality metallic fibres
to it. The resulting material is known as ultra-high performance fibre-reinforced concrete or
UHPFRC (Naaman 2011). However, UHPFRC is very costly compared with UHPC
(almost 68% costlier) (Kim 2018). UHPC, owing to its dense and homogenous
microstructure, also exhibits extraordinary durability performance. The water absorption of
UHPC is usually less than 1.5%, whereas the absorption of NSC is much higher than 3.5%
(Neville 2011; Scheydt and Müller 2012). UHPC significantly reduces the corrosion risk of
the reinforcing bars embedded in it, compared to NSC and HSC. The half-cell potential
(HCP) of the reinforcing bar embedded in UHPC was found to be around 150 mV and 140
mV less negative than the HCPs of those embedded in NSC and HSC, respectively (Jaafar
et al. 2018). Conventional UHPC incorporating high-volume cement content exhibits no or
insignificant (less than 0.5 mm) depth of carbonation after exposure to accelerated
carbonation curing (Scheydt and Müller 2012).

Although conventional UHPC exhibits superior mechanical and durability properties, its
large cement content imparts very high CO2 footprint (around 750–1200 kg/m3) to the
concrete (Yu et al. 2015). Several studies aimed to produce standard-cured UHPC (i.e.,
UHPC without special curing) with reduced cement content incorporating different types of
SCM, such as fly ash (Ahmed et al. 2021; Yu et al. 2015), rice husk ash (Tuan et al. 2011),
metakaolin (Norhasri et al. 2016), etc. GGBS is also used as an SCM. It is obtained from
slag which is a by-product of the steel industry. Unlike most of the SCMs, GGBS has both

15
Chapter 2

hydraulic and pozzolanic properties. This allows concrete to retain adequate strength and
durability even after high-volume of cement is substituted by GGBS. Few studies can be
found in the literature which investigated the effect of GGBS inclusion on the properties of
standard-cured UHPC with Portland cement, containing C3A content in the range of 4%–
8%, as primary binder (Abdulkareem et al. 2018; Yu et al. 2015). Mainly mechanical
properties were investigated in these studies. Incorporation of GGBS, especially beyond
30%, reduced the compressive strength of UHPC. More importantly, 28-day strength ≥ 150
MPa was not achieved for UHPC incorporating GGBS. Compressive strengths achieved in
these studies for UHPCs incorporating 30% GGBS were in the range of 110–142 MPa.
Detailed study on the durability properties of standard-cured UHPC incorporating GGBS is
scarce in the literature. Furthermore, study addressing the effects of GGBS incorporation on
the mechanical and durability properties of UHPC with very-low-C3A cement (C3A ≤ 3%)
as primary binder is unavailable.

Portland cements, especially the ones available in Australia, can be found to have C3A
content in four different ranges: high-C3A range (C3A > 10%), moderate-C3A range (6% <
C3A ≤ 10%), low-C3A range (3% < C3A ≤ 6%), and very-low-C3A range (C3A ≤ 3%). Among
these four groups, Portland cements with moderate C3A content are more commonly used
(Mohammadi and South 2016). Few studies investigated the fresh and hardened
performances of very-low-C3A cement (with 1.9%–2% C3A) based cementitious mixes
(Kadri et al. 2009; Mardani-Aghabaglou et al. 2017). These studies suggest that very-low-
C3A Portland cement, when compared with low- or moderate-C3A cement, reduces the
water demand of fresh cementitious mix containing high-range water reducer and improves
its rheological performance. The previous studies were conducted for cementitious mixes
with water to binder ratios (w/b) in the range of 0.3–0.6. However, the use of very-low-C3A
cement can be more beneficial to the fresh performance of UHPC, as UHPC is designed
with a very low w/b ratio, usually less than 0.2. Again, to attain a target workability, UHPC
with very-low-C3A cement can be designed with a lower w/b ratio than that of the moderate-
C3A cement based UHPC. Lower w/b ratio may allow UHPC to exhibit better performance
in hardened state. A study is therefore essential to comprehensively explore the benefits of
using very-low-C3A cement to the fresh and hardened performances of UHPC.

The present study aimed to develop ECO-UHPC mixes incorporating high-volume GGBS
with compressive strength greater than 150 MPa, utilizing the benefits of using very-low-
C3A cement (with 1.1% C3A) as primary binder, without employing any special curing or
fibres. The study also aimed to understand the performance variation between moderate-
C3A cement and very-low-C3A cement based UHPCs in terms of strength and workability.

16
Chapter 2

Mechanical properties such as, compressive strength, compressive stress-strain behaviour,


elastic modulus, splitting tensile strength, flexural strength, and durability properties such
as, water absorption, initial rate of absorption, corrosion risk (in terms of half-cell potential),
carbonation resistance were investigated for ECO-UHPC mixes with different GGBS to
very-low-C3A cement ratios (e.g., 0/100, 15/85, 30/70, 45/55, 60/40). X-ray diffraction
(XRD) analysis, and thermogravimetric and derivative thermogravimetric (TG-DTG)
analyses were conducted to better understand the properties of different ECO-UHPCs.

2.2 Materials and methods

2.2.1 Materials

Two types of Portland cement, one with a moderate C3A content (9.3%) and the other with
a very low C3A content (1.1%), were used in this study. For the convenience of discussion,
the cements will be termed respectively as MAC and VLAC in the following sections. Both
the cements conform to the requirements of ASTM C150 / C150M-16 Type I Portland
cement (ASTM C150 / C150M-16 2016). Supplementary binders used in this study, such as
GGBS and silica fume, conform to AS/NZS 3582.2 (2016) and AS/NZS 3582.3 (2016),
respectively. Quartz sand with a median diameter (d50) of 237.3 µm was used as aggregate.
The moisture content and absorption capacity of the quartz sand, measured in accordance
with ASTM C566-97 (1997) and ASTM C128-15 (2015), are 0.018% and 0.38%,
respectively. The particle size distributions (PSD) of the binders and the sand are shown in
Figure 2.1.

100
MAC
90
VLAC
80 Silica fume
70 GGBS
Quartz sand
60
% Finer

50
40
30
20
10
0
0.1 1 10 100 1000
Particle size (μm)

Figure 2.1. Particle size distributions of binders and sand

17
Chapter 2

Table 2.1. Chemical and physical properties of binders and sand

Material MAC VLAC Silica fume GGBS Quartz sanda


Oxides (mass %) Al2O3 5.13 3.18 0.13 13.04 0.017
CaO 63.53 64.48 0.33 42.80 0.002
Fe2O3 2.51 4.34 0.07 0.81 0.008
K2O 0.42 0.35 0.54 0.3 0.003
MgO 1.36 1.46 0.71 5.46 0.003
MnO 0.07 0.12 0.008 0.17 0.001
Na2O 0.34 0.14 0.29 0.28 0.003
P2O5 0.1 0.228 0.129 0.02 0.0
SiO2 20.49 21.4 94.6 31.14 99.88
SO3b 2.96 2.93 0.07 4.02 0.0
TiO2 0.25 0.192 0.005 0.56 0.025
Loss on 2.84 1.15 3.1 1.4 0.05
ignition

Cement phases C3S 56.4 63.7 - - -


(mass %)c C2S 16.2 13.4
C3A 9.3 1.1
C4AF 7.6 13.2

d50 (µm) 17.3 17.3 1.7 8.0 237.3


MAC = moderate-C3A cement, VLAC = very-low-C3A cement, C3S = Ca3SiO5 (alite), C2S = Ca2SiO4
(belite), C3A = Ca3Al2O6 (tricalcium aluminate), C4AF = Ca4(Al,Fe)2O7 (ferrite), d50 = median diameter
a
As per manufacturer
b
All oxides of S are noted here as SO3
c
Based on Bogue’s equations (ASTM C150 / C150M-16 2016)

Table 2.1 presents the chemical and physical properties of the binders and the sand. The
oxide compositions of the binders in Table 2.1 are based on their elemental concentrations
determined using inductively coupled plasma - optical emission spectrometry (ICP-OES).
Each binder sample was fused with 12:22 Norrish Flux (lithium metaborate/ lithium
tetraborate) in platinum crucibles at 1050 °C to prepare glass beads. The beads were then
dissolved in 1 M hydrochloric acid (HCl) for analysis of elemental concentrations using
Perkin Elmer Optima 5300 DV. Elemental concentrations were converted into oxide
compositions based on molar ratios.

The XRD patterns of the as-received binders are shown in Figure 2.2. Crystalline phases of
the MAC were alite (Ca3SiO5 or, C3S), belite (Ca2SiO4 or, C2S), tricalcium aluminate
(Ca3Al2O6 or, C3A), ferrite [Ca4(Al,Fe)2O7 or, C4AF], and gypsum (CaSO4.2H2O). Whereas
the crystalline phases of VLAC were mainly alite, belite, ferrite, and gypsum. The broad
hump in the 2θ range of 25–35° for as-received GGBS indicates that it was mostly in
amorphous state. Some crystalline peaks, corresponding to gypsum, akermanite
(Ca2MgSi2O7), gehlenite (Ca2Al2SiO7), were also observed in the XRD pattern of GGBS.
Higher gypsum peak intensity at around 11.5° in the XRD pattern of GGBS is indicative of

18
Chapter 2

its higher gypsum content than those in MAC and VLAC. The raw silica fume was also in
amorphous state mostly. However, few crystalline peaks corresponding to quartz (SiO2) and
silicon carbide (SiC) were identified in the XRD pattern of silica fume.

Chemical admixtures used in this study include a polycarboxylate based high-range water
reducer (HRWR) and a defoaming agent with 36% and 5.5% solid contents by mass,
respectively. Grade 500N deformed reinforcement bars (10 mm diameter), complying with
the requirements of AS/NZS 4671 (2001), were used for the preparation of half-cell potential
test specimens.

Figure 2.2. XRD patterns of raw binders (G = gypsum, A = alite, B = belite, F = ferrite, At =
tricalcium aluminate, Ak = akermanite, Ge = gehlenite, Q = quartz, SC = silicon carbide)

19
Chapter 2

2.2.2 UHPC mixes

The mix proportions of several preliminary UHPC mixes prepared to achieve the control
ECO-UHPC mix are presented in Table 2.2. Mix M-SF0WB181BA139 was synthesised with
MAC so as to compare between the performances of MAC and VLAC based UHPCs. Table
2.2 further presents the mix proportions of the ECO-UHPC mixes with different GGBS to
VLAC ratios, and the mix proportions of UHPC pastes prepared to conduct XRD analysis.
It should be noted that all the mix proportions in Table 2.2 are based on the masses of binders
and sand at air-dry condition. The 1-day bulk densities of the ECO-UHPC mixes, namely
S0, S15, S30, S45, and S60, were 2421 kg/m3, 2419 kg/m3, 2413 kg/m3, 2397 kg/m3, and
2412 kg/m3, respectively.

Figure 2.3 shows the combined PSDs of the ECO-UHPCs. The target PSD, based on
modified Andreasen and Andersen (A&A) model (Equation 2.1), is also shown in Figure
2.3. The value of q in Equation 2.1 was selected to be 0.23 as per the recommendations of
previous studies (Song et al. 2018; Yu et al. 2015). If compared with the target PSD, the
particle packing of UHPC mix became better with increasing VLAC replacement by GGBS.

𝐷𝑞 − 𝐷𝑚𝑖𝑛 𝑞
𝑝(𝐷) = (2.1)
𝐷𝑚𝑎𝑥 𝑞 − 𝐷𝑚𝑖𝑛 𝑞

D is particle size (µm), p(D) represents cumulative percent finer than size D, Dmin and Dmax
are respectively the minimum and maximum sizes of particle (µm).

2.2.3 Mixing, specimen preparation, and curing

In order to prepare each UHPC mix, first, the air-dry constituents such as cement, silica
fume, GGBS, and quartz sand were mixed for 5 minutes at a low speed of 60 rpm in a mixer.
After that, 100% of the water, 100% of the defoaming agent, and 80% of the HRWR were
added to the dry mix. The mixing was continued at 60 rpm for another 3 minutes. Then, the
rest of the 20% of HRWR was added to the mix. When the mix attained moist appearance,
the speed of the mixer was increased to 180 rpm. Mixing was continued until the mix
became flowable. Finally, the mixer speed was increased to 300 rpm and the mixing was
continued for 3 minutes. Fresh UHPC mix was then transferred to a workbench for
performing slump flow test and pouring into moulds. The mixing protocol mentioned above
was also adopted for the preparation of UHPC pastes, except that the quartz sand was
excluded from the process of paste preparation.

20
Chapter 2

Table 2.2. Mix proportions of UHPCs and UHPC pastes (by mass)

Mix type Mix ID. Constituents Mix design parameters


MAC VLAC SF S QS W HRWR* DA* b/a w/b
Preliminary UHPC M-SF0WB181BA139 1 - - - 0.72 0.181 0.013 0.0011 1.39 0.181
SF0WB181BA139 - 1 - - 0.72 0.181 0.013 0.0011 1.39 0.181
SF0WB166BA139 - 1 - - 0.72 0.166 0.013 0.0011 1.39 0.166
SF10WB166BA139 - 0.9 0.1 - 0.72 0.166 0.013 0.0011 1.39 0.166
SF20WB166BA139 - 0.8 0.2 - 0.72 0.166 0.013 0.0011 1.39 0.166
SF30WB166BA139 - 0.7 0.3 - 0.72 0.166 0.013 0.0011 1.39 0.166
SF20WB166BA125 - 0.8 0.2 - 0.8 0.166 0.013 0.0011 1.25 0.166
SF20WB166BA114† - 0.8 0.2 - 0.88 0.166 0.013 0.0011 1.14 0.166
SF20WB166BA104 - 0.8 0.2 - 0.96 0.166 0.013 0.0011 1.04 0.166

ECO-UHPC S0† - 1 0.25 - 1.1 0.208 0.016 0.0014 1.14 0.166


S15 - 0.85 0.25 0.15 1.1 0.208 0.016 0.0014 1.14 0.166
S30 - 0.7 0.25 0.3 1.1 0.208 0.016 0.0014 1.14 0.166
S45 - 0.55 0.25 0.45 1.1 0.208 0.016 0.0014 1.14 0.166
S60 - 0.4 0.25 0.6 1.1 0.208 0.016 0.0014 1.14 0.166

MAC or VLAC P-M-SF0WB181 1 - - - - 0.181 0.013 0.0011 - 0.181


based paste P-SF0WB181 - 1 - - - 0.181 0.013 0.0011 - 0.181

ECO-UHPC paste P-S0 - 1 0.25 - - 0.208 0.016 0.0014 - 0.166


P-S15 - 0.85 0.25 0.15 - 0.208 0.016 0.0014 - 0.166
P-S60 - 0.4 0.25 0.6 - 0.208 0.016 0.0014 - 0.166
MAC = moderate-C3A cement, VLAC = very-low-C3A cement, SF = silica fume, S = GGBS, QS = quartz sand, W = water, HRWR = high-range water reducer, DA =
defoaming agent, b/a = binder to aggregate ratio, w/b = water to binder ratio
*Solid content

SF20WB166BA114 and S0 represent same mix

21
Chapter 2

Figure 2.3. Combined particle size distributions of ECO-UHPCs (MAE and RMSE are
respectively mean absolute error and root-mean-square error between target curve and
combined PSD of each mix)

50 × 50 × 50 mm cube moulds, 50 × 100 mm cylindrical moulds, and 40 × 40 × 160 mm


prism moulds were used to prepare specimens for compressive strength test, splitting tensile
strength test, and four-point bending test, respectively. 100 × 100 × 100 mm cube moulds
were used to prepare the durability test specimens. In order to prepare half-cell potential
(HCP) test specimens, an activated titanium (Ti) electrode and a steel rebar (10 mm
diameter, 90 mm length) were embedded inside each 100 × 100 × 100 mm mould before
pouring UHPC mix. On each end of the steel rebar, 20 mm was covered with sealant tape.
The moulded UHPC samples were vibrated for a minute and then kept in a humidity-
controlled curing room (temperature 21 °C, 95% relative humidity). Both the mechanical
and durability test specimens were demoulded after 24 hours. The mechanical test
specimens were fully submerged in water (21 °C) until the time of testing, while the
durability specimens were left inside the humidity-controlled curing room. After 28 days,
half of the carbonation test specimens were moved to a CO2-controlled chamber and kept at
an elevated CO2 concentration of 10,000 ppm, a temperature of 21°C, and a relative
humidity of 60% for another 28 days. The curing conditions of the specimens for
carbonation test are illustrated in Figure 2.4.

22
Chapter 2

12000

CO2 concentration (ppm)


10000
Regime B
8000 (10000 ppm CO2,
21 °C, 60% RH)
6000
4000
Regime A
2000 (300 ppm CO2,
21 °C, 95% RH)
0
0 7 14 21 28 35 42 49 56
Curing time (days)

Figure 2.4. Curing regimes of carbonation test specimens

2.2.4 Test methods

2.2.4.1 Slump flow test

Slump flow tests of the fresh UHPC mixes were conducted according to the guidelines given
by AS 1012.3.5 (2015). A mini slump cone (height 116 mm, top and bottom diameters
respectively 38 mm and 76 mm) was used to conduct slump test. The aspect ratio of the cone
was similar to that of the standard cone specified by AS 1012.3.5 (2015).

2.2.4.2 Mechanical tests

A compression testing machine with a capacity of 600 kN was used to conduct the
mechanical tests. The compressive strength tests were performed as per ASTM C109 /
C109M-16a (2016). Strain gauges were attached to the 28-day ECO-UHPC specimens for
measuring their strains under compression. The elastic moduli of the specimens were
calculated according to AS 1012.17 (1997). The splitting tensile strength tests and four-point
bending tests were conducted for the ECO-UHPCs as per the guidelines of AS 1012.10
(2000) and AS 1012.11 (2000), respectively. For each mechanical property, three specimens
were tested.

2.2.4.3 Durability tests

The water absorption (WA) and the initial rate of absorption (IRA) tests were carried out
for the ECO-UHPC mixes, using 100 × 100 × 100 mm plain cubes, according to the
guidelines given by IS 3495 (Part 2) (1992) and AS/NZS 4456.17 (2003), respectively.
Equations 2.2 and 2.3 were used to measure the WA and IRA of each specimen,

23
Chapter 2

respectively. Three specimens prepared from every ECO-UHPC mix were tested for both
WA and IRA.

100(𝑊2 − 𝑊1 )
𝑊𝐴 = (%) (2.2)
𝑊1

1000(𝑊3 − 𝑊1 )
𝐼𝑅𝐴 = (kg/m2/min) (2.3)
𝐴

W1 is the mass of oven-dried specimen (g), W2 is the mass of specimen after 24 hours of
immersion in water (g), W3 is the mass of specimen after immersion of bottom 3 mm under
water for a minute (g) (Figure 2.5), A is the bottom surface area of specimen (mm2). Damp
cloth was used to wipe off the excess water on the specimen surface prior to measuring W2
or W3.

Half-cell potential (HCP) tests were performed for the ECO-UHPCs using a modified
version of the method mentioned in ASTM C876-15 (2015). An illustration of the test setup
is presented in Figure 2.6. The HCP of each cube specimen was calculated based on the
potential difference between the rebar and a reference saturated calomel electrode (SCE)
placed on top of the cube specimen. For each mix three specimens were tested. Method
specified in ASTM C876-15 (2015) does not include the use of Ti electrode. Nevertheless,
Ti electrodes were incorporated in the HCP test specimens of this study, to cross-check the
accuracy of the obtained HCP readings. Equation 2.4 was used to assess the accuracy of the
measured values.

Erebar/SCE = ETi/rebar + ETi/SCE (mV) (2.4)

UHPC cube

Rollers

3 mm
Water

Figure 2.5. Setup for initial rate of absorption test

24
Chapter 2

Figure 2.6. Setup for half-cell potential test

Erebar/SCE, ETi/rebar, and ETi/SCE in Equation 2.4 represent respectively the potential differences
between rebar and SCE, Ti electrode and rebar, and Ti electrode and SCE. As per the
suggestions given by ASTM C876-15 (2015), the HCP readings obtained using SCE were
converted to copper-copper sulphate electrode equivalent HCP readings in accordance with
the guidelines mentioned in (Elchalakani et al. 2018; Meek et al. 2018).

Carbonation tests were conducted according to EN 13295 (2004), using 100 × 100 × 100
mm plain specimens prepared from S0, S30, and S60. Half of the specimens were tested after
28 days of exposure in curing regime A (Figure 2.4). The remaining specimens were tested
after being cured in regime B for another 28 days. At the time of testing, each specimen was
split into two halves and phenolphthalein indicator (mixture of 1 g phenolphthalein powder,
70 ml ethanol, and 30 ml water) was sprayed over the split section of each half. The colour
change of the split section was observed to assess the carbonation resistance of each mix.

2.2.4.4 XRD and TG-DTG analyses

XRD analysis was conducted using powdered UHPC paste samples. For each sample, XRD
pattern in the 2θ range of 5–70° was recorded using an X-ray diffractometer with Cu K
radiation (λ = 1.5404 Å), 40 kV, and 40 mA at a rate of 2°/min. For P-M-SF0WB181 and P-
SF0WB181 pastes, XRD patterns were recorded at the ages of 20 h and 28 days. XRD patterns
were also recorded for P-S0, P-S15, and P-S60 pastes at 28 days.

25
Chapter 2

Figure 2.7. Performances of M-SF0WB181BA139, SF0WB181BA139, and SF0WB166BA139

TG-DTG analyses were conducted using powdered S0, S15, and S60 samples. Prior to
conducting TG-DTG analysis, each powdered sample having mass of 10–15 mg was
completely dried in an oven at 40 °C until constant mass was attained. The oven-dried
sample was then heated initially from 40 °C to 105 °C in nitrogen atmosphere at a heating
rate of 10 °C/min. After that, the sample was kept in isothermal condition at 105 °C for 10
minutes. Then the sample was again heated at a rate of 10 °C/min until it reached a
temperature of 1000 °C. The change in mass of each sample, with increasing temperature,
was recorded to construct TG and DTG diagrams.

2.3 Results and discussion

2.3.1 Fresh and hardened performances of moderate- and very-low-C3A cement based UHPCs

The compressive strength and slump flow diameters of M-SF0WB181BA139 and


SF0WB181BA139 mixes, synthesised respectively with MAC and VLAC, are presented in
Figure 2.7. The w/b and b/a ratios were kept constant at 0.181 and 1.39 for both the mixes,
and silica fume was not used in the mixes. The slump of SF0WB181BA139 was 34% higher
with respect to that of M-SF0WB181BA139. Hydration of C3A phase in presence of gypsum
(CaSO4.2H2O) starts at a very early age (as early as few seconds) (Meier et al. 2015). A
notable portion of water is consumed during the hydration of C3A. Very-low-C3A cement
causes substantially low early ettringite (AFt) formation compared with moderate-C3A
cement [as illustrated in Figures 2.8(a) and 2.8(b)], and thus demands less water to impart
workability to the fresh cementitious mix. Hence, at a given w/b ratio, very-low-C3A cement

26
Chapter 2

imparts better flowability to cementitious mix than moderate-C3A cement (Kadri et al.
2009). Higher slump of SF0WB181BA139 than that of M-SF0WB181BA139 can therefore be
explained by the lower early age AFt formation in the former. It is worth noting that the
PSDs of MAC and VLAC used in this study were similar (Figure 2.1). Therefore, it can be
assumed that the PSDs of these materials had little influence on the slump variation between
M-SF0WB181BA139 and SF0WB181BA139 mixes.

Figure 2.8. Development of hydration products in M-SF0WB181BA139, SF0WB181BA139, and


SF0WB166BA139 (sand particles are not shown for keeping demonstration simple)

27
Chapter 2

At the age of 3 days, M-SF0WB181BA139 had slightly higher strength than SF0WB181BA139 (by
3.9%). However, the 28-day strength of SF0WB181BA139 was 7.4% higher, with respect to M-
SF0WB181BA139. AFt formation at early age most likely helped M-SF0WB181BA139 gain higher
early strength. On the other hand, water consumption during the early AFt formation
possibly left slightly less water available for the hydration of C3S (alite) and C2S (belite) in
M-SF0WB181BA139 than that in SF0WB181BA139. Again, lower C3S content in MAC than that
in VLAC perhaps caused lower formation of calcium silicate hydrate (C-S-H) in M-
SF0WB181BA139 than that in SF0WB181BA139 [Figures 2.8(a) and 2.8(b)]. Consequently, at 28
days, higher strength was observed for SF0WB181BA139 than that for M-SF0WB181BA139 (Bach
2019; Goñi et al. 2010). The w/b ratio of VLAC based mix SF0WB181BA139 was reduced from
0.181 to 0.166, and mix SF0WB166BA139 was prepared, so as to get comparable slump to that
of MAC based M-SF0WB181BA139. Reduced w/b ratio possibly allowed SF0WB166BA139 to
have slightly denser microstructure than that of SF0WB181BA139 [Figures 2.8(b) and 2.8(c)],
which therefore helped SF0WB166BA139 attain slightly higher 28-day strength (by 2.1%). It
was also observed that SF0WB166BA139 exhibited 9.7% higher 28-day strength, with respect
to M-SF0WB181BA139.

2.3.2 Development of control ECO-UHPC mix

VLAC in mix SF0WB166BA139 was substituted by 10%, 20%, and 30% silica fume to prepare
mixes SF10WB166BA139, SF20WB166BA139, and SF30WB166BA139, respectively. The slump flow
and compressive strength results of the mixes are shown in Figure 2.9. The maximum 28-
day strength was achieved for 20% substitution of cement by silica fume (i.e., for
SF20WB166BA139). This was expected as silica fume in UHPC helps portlandite (CH) crystals
to convert into C-S-H (Andrade et al. 2019). However, the results also show that replacing
VLAC by silica fume beyond 20% impaired the strength. The strength drop in UHPC with
30% silica fume can be attributed to the dilution effect (i.e., reduction of binder with
hydraulic reactivity) (Bonavetti et al. 2000), and/or to the presence of insufficient CH
content inducing limited pozzolanic activity (Tironi et al. 2013). The slump of fresh UHPC
reduced with increasing replacement of VLAC by silica fume. This phenomenon can be
related with the high fineness of the silica fume compared to that of VLAC (d50 of silica fume
= 1.7 µm and d50 of VLAC = 17.3 µm). The silica fume, owing to its higher fineness, had
higher surface-area per unit volume than the cement. As a result, incorporation of silica
fume, substituting certain percentage of cement, increased the water demand to lubricate the
particles present in fresh UHPC mix and reduced its workability (Sabet et al. 2013).

28
Chapter 2

Figure 2.9. Performances of UHPCs with different silica fume contents

Figure 2.10. Performances of UHPCs with different b/a ratios

The binder to aggregate ratio (b/a) of SF20WB166BA139 was changed from 1.39 to 1.25, 1.14,
and 1.04 in order to prepare mixes SF20WB166BA125, SF20WB166BA114, and SF20WB166BA104,
respectively. Figure 2.10 shows the performances of the mixes with different b/a ratios.
Higher 28-day strength was achieved for SF20WB166BA114, with respect to other mixes.
SF20WB166BA114, with b/a ratio of 1.14, was therefore selected as the control ECO-UHPC
mix in this study. High binder content and low w/b ratio of UHPC, compared to NSC or
HSC, lead to its very high strength. However, the results in Figure 2.10 suggest that opting
for b/a ratio higher than a certain value (1.14 in the present study) may not lead to a higher
strength. Similar results were reported by Xie et al. (2018). Xie et al. (2018) mentioned that

29
Chapter 2

use of very high b/a ratio causes excessive amount of binder to remain unreacted and
weakens the interfacial transition zone (ITZ) between aggregate and binder. It should be
noted that the mixes S0 and SF20WB166BA114 are identical.

2.3.3 Slump flow results of ECO-UHPC mixes

The slump flow results of the fresh ECO-UHPC mixes are presented in Figure 2.11. The
slump of the ECO-UHPC mixes decreased with cement replacement by GGBS. The GGBS
used in this study was much finer than VLAC (from Table 2.1, d50 of VLAC = 17.3 µm, d50
of GGBS = 8 µm). As a result, GGBS incorporation increased the water demand of the fresh
UHPC and reduced its workability (Ting et al. 2019). It is interesting to note that some
researchers observed that the workability of NSC or HSC improved when moderate-C3A
cement was replaced by GGBS (Adjoudj et al. 2014; Johari et al. 2011). There may be a
number of reasons behind the difference between the present and previous observations.
First, substituting moderate-C3A cement by GGBS reduces the C3A present in a
cementitious mix and thereby reduces early AFt formation, which eventually enhances the
workability of the mix (Samet and Chaabouni 2004). However, this effect becomes
insignificant when very-low-C3A cement is replaced by GGBS. Second, the GGBS used in
this study most likely contained higher gypsum content than that in VLAC (Figure 2.2). In
presence of gypsum, GGBS grains are likely to take part in the formation of early age AFt
and reduce workability (Matschei et al. 2005). Third, unlike the present study, the GGBS
used in (Adjoudj et al. 2014) was slightly coarser than the cement used in the study; hence
the PSD of the GGBS had either no effect or positive impact on the workabilities of the
mixes with GGBS investigated in the study.

Figure 2.11. Slump flow results of ECO-UHPC mixes

30
Chapter 2

Figure 2.12. Compressive strength gains of ECO-UHPCs with time

2.3.4 Mechanical properties of ECO-UHPC mixes

2.3.4.1 Compressive strength

The compressive strength gains of the ECO-UHPCs with time are presented in Figure 2.12.
Irrespective of age, the compressive strengths of ECO-UHPCs with GGBS were generally
lower than that of S0. However, the strength gain rates of S15 and S30 were slightly high in
between 14 and 28 days, compared to S0. Such behaviour can be explained by slow nature
of pozzolanic reaction (Oner and Akyuz 2007). The strength gain rates of S45 and S60 were
higher than that of S0 in between 3 and 7 days. Very fine particles of GGBS possibly induced
filler effect (Xu et al. 2017) and heterogenous nucleation effect (i.e., fine GGBS particles
acting as nucleation sites for hydrates) (Lawrence et al. 2003), which possibly helped S45
and S60 attain higher strength gain rates in between 3 and 7 days. Nevertheless, the strength
gain rates of S45 and S60 became low after 14 days compared to other UHPCs. This possibly
resulted from the dilution effect due to substantial reduction of cement (Elchalakani et al.
2017).

The 28-day strengths of S15, S30, S45, and S60 were respectively 1.3%, 6.6%, 10.5%, and
16.1% lower with respect to that of S0. However, at 28 days, all the ECO-UHPCs attained
ultra-high strengths (i.e., strengths exceeding 150 MPa). Some researchers observed that low
replacement of cement by GGBS (e.g., 10%–30%) improved the 28-day strength of NSC
[52,56]. NSC usually does not contain any silica fume. On the other hand, UHPC contains
a binary cement – silica fume binder system. Therefore, the increment or decrement in the
compressive strength of UHPC upon replacing cement by GGBS is dependent on its silica
fume content. Since S0 was developed with optimal substitution of cement by silica fume,
substitution of cement further by GGBS did not lead to a higher strength.

31
Chapter 2

(a) (b)

Figure 2.13. (a) 28-day stress – strain curves of ECO-UHPCs; (b) 28-day elastic moduli of
ECO-UHPCs

(a) (b)

Figure 2.14. (a) 28-day splitting tensile strengths of ECO-UHPCs; (b) 28-day flexural strengths
of ECO-UHPCs

2.3.4.2 Elastic modulus, splitting tensile strength, and flexural strength

Figure 2.13(a) shows the stress-strain curves of ECO-UHPCs under compression at 28 days.
The compressive stress of each of the ECO-UHPC specimens increased almost linearly with
strain until failure, without any notable inelastic softening. This type of stress-strain
behaviour was expected for highly brittle material. The 28-day elastic moduli of the ECO-
UHPCs are presented in Figure 2.13(b). The elastic modulus of ECO-UHPCs with GGBS
were generally lower than that of S0. The elastic moduli of S15, S30, S45, and S60 were
respectively 2.4%, 7.6%, 9.9%, and 11.5% lower, with respect to the elastic modulus of S0.

The 28-day splitting tensile strengths and flexural strengths of ECO-UHPCs are shown in
Figures 2.14(a) and 2.14(b), respectively. The splitting tensile strengths of all the UHPC

32
Chapter 2

mixes were higher than 8 MPa, without incorporation of fibres. The splitting tensile strength
of ECO-UHPCs with GGBS were generally lower than that of S0. The splitting tensile
behaviour of ECO-UHPC agrees with its compressive strength behaviour, although the
former seems to be slightly less sensitive to the GGBS to VLAC replacement ratio. This may
be because the presence of fine GGBS improved the ITZ between matrix and aggregate (Gao
et al. 2005). The 28-day splitting tensile strengths of S15, S30, S45, and S60 were respectively
0.7%, 9%, 9.6%, and 14.1% lower with respect to that of S0. The flexural strengths of ECO-
UHPCs with GGBS were also generally lower than that of S0. The flexural strengths of S15,
S30, S45, and S60 were respectively 0.6%, 3.9%, 8.4%, and 14.8% lower with respect to the
flexural strength of S0, at 28 days.

2.3.4.3 Comparison of mechanical properties with model predictions

The elastic modulus, splitting tensile strength, and flexural strength results of the ECO-
UHPC mixes and the data reported in previous studies on GGBS based UHPC
(Abdulkareem et al. 2018; Yu et al. 2015) are plotted against their respective compressive
strength results in Figures 2.15(a), 2.15(b), and 2.15(c), respectively. Figure 2.15 furthermore
presents the relationships suggested by different standards for predicting elastic modulus,
splitting tensile strength, and flexural strength of concrete (ACI 318 2014; ACI 363R 1992;
AS 3600 2009). Table 2.3 shows the equations representing the relationships. Table 2.3 also
presents the mean absolute error and root-mean-square error between each relationship and
the respective experimental data of the ECO-UHPCs and the UHPCs in (Abdulkareem et
al. 2018; Yu et al. 2015).

Figure 2.15(a) shows that, given an error of ±20%, the elastic modulus – compressive
strength relationships proposed by ACI 318 (2014) and AS 3600 (2009) tend to overestimate
the elastic moduli of the ECO-UHPC mixes. This is because the relationships were proposed
for concrete with coarse aggregate (i.e., NSC and/or HSC). The coarse aggregate in concrete
is usually much stiffer than the matrix. But coarse aggregate is typically not used in UHPC.
Hence, the elastic modulus to compressive strength ratio of UHPC is lower than that of
concrete with coarse aggregate (Ma et al. 2004). The ACI 363R (1992) model, proposed for
HSC, seemingly provides better estimations of the elastic moduli of the ECO-UHPCs than
other models, with an error of ±10%.

Figures 2.15(b) and 2.15(c) respectively show that, given an error of ±20%, the splitting
tensile strength – compressive strength relationships proposed by ACI 318 (2014) and AS
3600 (2009), and the flexural strength – compressive strength relationships proposed by ACI
318 (2014), ACI 363R (1992), and AS 3600 (2009) underestimate the splitting tensile

33
Chapter 2

strengths and the flexural strengths of the ECO-UHPCs and the UHPCs in (Abdulkareem
et al. 2018; Yu et al. 2015). In NSC and HSC, cracks form around the ITZ between matrix
and coarse aggregate under external load. Absence of coarse aggregate in UHPC
significantly reduces the formation of such cracks and therefore increases the splitting tensile
strength and the flexural strength of the concrete (Richard and Cheyrezy 1995).
Furthermore, the ITZs between matrix and aggregate particles in UHPC are usually much
denser and thinner than those in NSC or HSC due to the use pozzolanic silica fume and
very fine sand (Ghafor et al. 2020; Zhu et al. 2015).

From the discussion above, it can be understood that the currently available relationships in
(ACI 318 2014; ACI 363R 1992; AS 3600 2009), except the elastic modulus – compressive
strength relationship in ACI 363R (1992), do not provide satisfactory estimations of the
elastic moduli, splitting tensile strengths, and flexural strengths of the ECO-UHPCs and
other standard-cured GGBS based UHPCs in the literature (Abdulkareem et al. 2018; Yu et
al. 2015). Because of limited datapoints, the current study did not propose new relationships.
Further investigation is essential with more experimental data to develop relationships that
will be more representative of the mechanical properties of UHPC incorporating GGBS.

Table 2.3. Mechanical relationships proposed by standards, and MAEs and RMSEs between
relationships and respective UHPC data

Mechanical Standard Equation Concrete type MAE RMSE


relationship
Elastic modulus – ACI 318 (2014) 𝐸𝑐 = 4.7√𝑓𝑐 NSC 13.46 13.47
compressive strength
ACI 363R
𝐸𝑐 = 3.32√𝑓𝑐 + 6.9 HSC 2.44 2.64
(1992)
AS 3600 (2009) 𝐸𝑐 = 5.05√𝑓𝑐 NSC and HSC 18.00 18.01
Splitting tensile ACI 318 (2014) 𝑓𝑠𝑝 = 0.56√𝑓𝑐 NSC 1.92 1.95
strength – compressive
ACI 363R
strength 𝑓𝑠𝑝 = 0.59√𝑓𝑐 HSC 1.55 1.59
(1992)
AS 3600 (2009) 𝑓𝑠𝑝 = 0.4√𝑓𝑐 NSC and HSC 3.88 3.90
Flexural strength – ACI 318 (2014) 𝑓𝑟 = 0.62√𝑓𝑐 NSC 12.52 12.83
compressive strength
ACI 363R
𝑓𝑟 = 0.94√𝑓𝑐 HSC 8.82 9.17
(1992)
AS 3600 (2009) 𝑓𝑟 = 0.6√𝑓𝑐 NSC and HSC 12.75 13.06
Ec = Elastic modulus (GPa), fsp = splitting tensile strength (MPa), fr = flexural strength (MPa), fc =
compressive strength (MPa), MAE = mean absolute error, RMSE = root-mean-square error

34
Chapter 2

Figure 2.15. Comparison of (a) elastic modulus, (b) splitting tensile strength, and (c) flexural
strength of UHPC incorporating GGBS with model predictions (points in grey circles represent
data from UHPCs incorporating GGBS)

(a) (b)

Figure 2.16. (a) Water absorption, and (b) initial rate of absorption test results of ECO-UHPCs

2.3.5 Durability properties of ECO-UHPC mixes

2.3.5.1 Water absorption (WA) and initial rate of absorption (IRA)

Figures 2.16(a) and 2.16(b) present the WA and IRA results of the ECO-UHPCs,
respectively. Both the 7-day and 28-day WAs and IRAs of the ECO-UHPCs incorporating

35
Chapter 2

GGBS were generally lower than those of S0. The 28-day WAs of S15, S30, S45, and S60
reduced by 15.3%, 30.4%, 25.3%, and 51.9%, and their 28-day IRAs reduced by 36.6%,
62.8%, 53.4%, and 90.8%, with respect to the 28-day WA and IRA of S0, respectively.
Substantially fine particles of GGBS (from Table 2.1, d50 of VLAC = 17.3 µm, d50 of GGBS
= 8 µm) are likely to fill the gaps between larger particles. This can facilitate better particle
packing in UHPC mix with GGBS (as can be seen from Figure 2.3). Better particle packing,
owing to the incorporation of fine GGBS, possibly reduced the capillary porosities of the
ECO-UHPCs with GGBS and thus reduced their WAs and IRAs with respect to those of S0
(Abd Elrahman and Hillemeier 2014). It is also worth noting that the WAs of all the ECO-
UHPCs, both at 7 and 28 days, were substantially less than 2%, which is the maximum limit
for concrete to be considered having ample durability for construction works with fifty-year
design life (Elchalakani et al. 2017).

2.3.5.2 Half-cell potential (HCP)

The HCP test results of the ECO-UHPCs are presented in Figure 2.17. The HCPs of all the
ECO-UHPCs were less negative than −350 mV, at 3 days. Irrespective of age, the HCPs of
S0 were generally more negative and the HCPs of S60 were generally less negative than the
HCPs of other ECO-UHPCs. The 28-day HCPs of S15, S30, S45, and S60 were respectively
41.2 mV, 90.9 mV, 60.7 mV, and 96.6 mV less negative than that of S0. Less negative HCPs
of the ECO-UHPCs incorporating GGBS imply lower corrosion risks in these UHPCs than
that in S0. The observation complies with the results reported by Cheng et al. (2005) for
NSC with GGBS. Cheng et al. (2005) observed that NSC with 60% GGBS, despite having
lower alkalinity, resulted in less negative HCP than that of the NSC without GGBS. The
researchers concluded that the porosity of concrete is the dominating factor that influences
its HCP. Hence, for UHPC, it can also be inferred that partial substitution of cement by
GGBS (up to 60%) reduces the porosity of UHPC (based on the WA and IRA results), and
consequently reduces the corrosion risk of the rebar embedded in it.

2.3.5.3 Carbonation resistance

Upon spraying phenolphthalein solution, dark pink or purple colour was observed
throughout the cross-sections of S0, S30, and S60 specimens after 28 days of exposure to
curing regime A, or after 28 days of exposure to regime A and followed by 28 days exposure
to regime B (Figure 2.18). No visible peripheral carbonation front was detected in any of the
samples. This implies, irrespective of exposure condition, the depth of carbonation was
below the detection limit (<0.5 mm) for S0, S30, and S60. These observations on UHPCs

36
Chapter 2

with different GGBS contents are in agreement with the results observed by Scheydt and
Müller (2012) in their investigations on conventional UHPC without GGBS.

GGBS incorporation tends to reduce the CH content in concrete (because of the pozzolanic
reactivity of GGBS) (Saillio et al. 2019). Lower CH content increases the possibility of its
conversion into calcium carbonate (CaCO3) within a shorter period of time when it comes
in contact with external CO2. This again increases the chance of concrete with GGBS
(especially with 50% or more) exhibiting higher depth of carbonation (Han-Seung and Wang
2016). However, such phenomenon was not observed in the current study for S30 or S60,
even after its exposure to elevated CO2 environment. This reflects comparable durability of
S30 and S60 to that of S0, in elevated CO2 environment. Most likely the dense
microstructures of the ECO-UHPCs, owing to their very low w/b (= 0.166), substantially
reduced the diffusion of CO2 into them and resulted in carbonation depths below the
detection limit (Paul et al. 2018).

Figure 2.17. Half-cell potential test results of ECO-UHPCs

Figure 2.18. Illustration of carbonation resistances of specimens (a) S0, (b) S30, and (c) S60
after 28 of days exposure to curing regime A (300 ppm CO2), followed by 28 days of exposure
to curing regime B (10,000 ppm CO2)

37
Chapter 2

Figure 2.19. XRD patterns of P-M-SF0WB181, P-SF0WB181 pastes at (a) 20 hours, and (b) 28
days (A = alite, B = belite, E = ettringite, P = portlandite)

2.3.6 XRD analysis

2.3.6.1 XRD analysis of moderate- and very-low-C3A cement based UHPC pastes

The XRD patterns of P-M-SF0WB181 and P-SF0WB181 pastes (i.e., pastes of MAC and VLAC
cement based UHPCs, respectively), at 20 h and 28 days, are presented in Figure 2.19. For
each paste, at both 20 h and 28 days, the intensities of AFt peaks at around 9° and 16°, CH
peak at around 18°, and C3S peaks at around 34.3°, 41.2°, 51.7°, 56.5°, and 62.3° are
presented in Table 2.4. At the age of 20 h, the AFt intensity at around 9° and 16° were
substantially low for P-SF0WB181 paste when compared with P-M-SF0WB181. AFt at an early
age, such as 20 h, is supposed to result mainly from the hydration of C3A in presence of
gypsum. AFt like product is also produced during the hydration of C4AF (ferrite phase) in
presence of gypsum; however, the reaction being much slower than that of C3A, insignificant
amount of AFt is expected to originate from C4AF at a very early age. Low early AFt
formation in P-SF0WB181, compared to P-M-SF0WB181, was therefore a result of very low
C3A content of the cement used in P-SF0WB181. This indicates that better flowability and
slightly lower 3-day strength of SF0WB181BA139 than those of M-SF0WB181BA139 were caused
by lower AFt formation in the former (Figure 2.7). At 28 days, the AFt peak intensities of
P-SF0WB181 became higher than the respective AFt intensities at 20 h. A significant portion
of AFt in P-SF0WB181 at 28 days possibly originated from the hydration of C4AF. C-S-H, the

38
Chapter 2

main strength imparting hydrate, could not be identified in the XRD patterns of P-M-
SF0WB181 and P-SF0WB181 pastes, because of the nano-crystalline structure of C-S-H (Tabet
et al. 2018). However, in between 20 h and 28 days, more pronounced reductions of the alite
peak intensities for P-SF0WB181 than those for P-M-SF0WB181 possibly indicate higher
formation of C-S-H in P-SF0WB181 at 28 days. Therefore, SF0WB181BA139 exhibited higher
28-day strength than that of M-SF0WB181BA139 (Figure 2.7).

2.3.6.2 XRD analysis of ECO-UHPC pastes

The XRD patterns of P-S0, P-S15, and P-S60 pastes at the age of 28 days are presented in
Figure 2.20. For each of the pastes, the intensities of AFt peaks at around 9° and 16°, CH
peak at around at around 18°, and C3S peaks at around 34.3°, 41.2°, 51.7°, 56.5°, and 62.3°
are presented in Table 2.4. The intensities of CH and clinker minerals were generally lower
for pastes with GGBS than those for P-S0. Interestingly, the AFt intensity at around 9° was
high for P-S15 compared to P-S0, and for P-S60 compared to P-S15. This is possibly because
the GGBS used in this study had higher gypsum content than VLAC (Figure 2.2), in
presence of which GGBS grains participated in the formation of AFt (Matschei et al. 2005).

2.3.7 TG-DTG analyses of ECO-UHPC samples

The 28-day TG and DTG results of powdered S0, S15, and S60 samples are shown in
Figures 2.21(a) and 2.21(b), respectively. As discussed in Section 2.2.4.4, before performing
the TG-DTG analysis, each powdered ECO-UHPC sample was dried at 40 °C until constant
mass was attained. Hence, mass loss in the range of 40–110 °C can be attributed mainly to
the dehydration of C-S-H and AFt (Ling et al. 2018; Sun et al. 2019; Tajuelo Rodriguez et
al. 2017). Mass losses in the ranges of 400–460 °C and 520–900 °C can be attributed mainly
to the decomposition of CH and to the decarbonation of CaCO3, respectively (Sun et al.
2019; Wu et al. 2016). Table 2.5 presents the mass losses due to dehydration of C-S-H and
AFt, decomposition of CH, and decarbonation of CaCO3 for S0, S15, and S60 samples. The
CH contents in S0, S15, and S60 were calculated based on Equation 2.5 (Gao et al. 2015).
The calculated CH contents are presented in Table 2.5.

𝑀𝐶𝐻
𝐶𝐻 = × ∆𝑚𝐶𝐻 (%) (2.5)
𝑀𝐻

MCH and MH are respectively the molar masses of CH and water, ∆mCH is the mass loss due
to the decomposition of CH (%).

39
Chapter 2

Table 2.4. Intensities (counts) of XRD peaks corresponding to AFt, CH, and C3S

Paste AFt CH C3S


2θ ≈ 2θ ≈ 2θ ≈ 2θ ≈ 2θ ≈ 2θ ≈ 2θ ≈ 2θ ≈
9° 16° 18° 34.3° 41.2° 51.7° 56.5° 62.3°
P-M- 1471 735 1356 7625 3774 2949 1439 1409
SF0WB181
(20-h)
P-SF0WB181 412 235 1120 6640 3613 4003 2285 2179
(20-h)
P-M- 1565 822 3181 5566 2938 2348 1199 1123
SF0WB181
(28-day)
P-SF0WB181 988 638 3359 4215 2481 1829 997 1128
(28-day)
P-S0 (28-day) 988 719 2012 4351 1885 1912 771 888
P-S15 (28- 1221 715 1481 3957 1772 1551 783 831
day)
P-S60 (28- 1552 738 1312 1725 870 724 440 420
day)

Figure 2.20. XRD patterns of P-S0, P-S15, and P-S60 pastes at 28 days (A = alite, B = belite, E
= ettringite, P = portlandite)

40
Chapter 2

° °

Figure 2.21. (a) TG, and (b) DTG results of S0, S15, and S60

Table 2.5. TG results of S0, S15, and S60

UHPC Mass loss from reaction product/ products (%) CH content (%)
C-S-H and AFt CH CaCO3
(40–110 °C) (400–460 °C) (520–900 °C)
S0 3.67 0.58 1.72 2.38
S15 3.60 0.48 1.59 1.96
S60 3.11 0.37 1.35 1.54

The mass loss of S0 due to the dehydration of C-S-H and AFt was slightly higher than S15.
On the other hand, the mass loss of S60 due to the dehydration of C-S-H and AFt was lower
than those of both S0 and S15. This implies possible presence of higher amount of C-S-H
and AFt in S0 than those in S15 and S60, and the lowest presence of C-S-H and AFt in S60.
The peak corresponding to CH, in the DTG – temperature diagram [Figure 2.21(b)], was
observed to be more pronounced for S0 than those for S15 and S60. Whereas the CH peak
of S60 was less pronounced compared to both S0 and S15. The observation is indicative of
higher CH content in S0 than those in S15 and S60, and the lowest CH content in S60. The
calculated CH contents in S0, S15, and S60 samples, presented in Table 2.5, support the
forgoing inference. The observation is also consistent with the XRD results of the ECO-
UHPC pastes. Higher presence of reaction products in S0 possibly helped it attain higher
compressive strength than other ECO-UHPCs (Figure 2.12). On the other hand, lower
strength of S60 can be attributed to its lower content of reaction products. Peaks
corresponding to CaCO3 were observed in the DTG – temperature diagrams of the ECO-
UHPCs. This is possibly because some carbonation of the hydration products took place
during the sample preparation process prior to conducting TG-DTG analysis (Wu et al.
2016). The mass loss due to the decarbonation of CaCO3 was higher for S0 than those for

41
Chapter 2

S15 and S60. This can be attributed to the presence of lower amount of carbonatable reaction
products in S15 and S60 than that in S0.

2.4 Conclusions

Based on the results obtained in this study, the following conclusions can be drawn:

1. At a w/b ratio of 0.181, VLAC imparts 34% higher slump flow to UHPC with respect to
that imparted by MAC.
2. UHPC incorporating VLAC may exhibit slightly low 3-day compressive strength
compared to UHPC with MAC. However, for a given slump flow, VLAC allows UHPC
to attain higher 28-day compressive strength than that of MAC based UHPC.
3. The compressive strength of UHPC, with optimum silica fume to VLAC ratio, generally
reduces with cement replacement by GGBS. 60% replacement of VLAC by GGBS
reduces the 28-day compressive strength of UHPC by 16.1% with respect to the strength
of UHPC without GGBS. Nevertheless, ultra-high strength (>150 MPa) can still be
achieved up to 60% substitution of VLAC by GGBS, without employing any special
curing or fibres. It is interesting to note that the strength attained in this study for ECO-
UHPC with GGBS to VLAC ratio of 60/40 is higher than previously reported strengths
for standard-cured UHPCs with lower ratio of GGBS to low- or moderate-C3A cement
(e.g., 30/70).
4. The elastic modulus, flexural strength, and splitting tensile strength of UHPC generally
reduce with the substitution of VLAC by GGBS. For 60% substitution of VLAC by
GGBS, the 28-day elastic modulus, flexural strength, and splitting tensile strength of
UHPC reduce respectively by 11.5%, 14.1%, and 14.8% with respect to the respective
mechanical properties of UHPC without GGBS.
5. The water absorption and initial rate of absorption of UHPC, and the corrosion risk of
rebar embedded in UHPC reduce with the increase of VLAC substitution by GGBS.
6. UHPCs made with up to 60% replacement of VLAC by GGBS exhibit carbonation
resistance similar to that of the UHPC without GGBS.

Acknowledgements

The authors express special gratitude to Mr. Bradley Connolly from Adelaide Brighton
Cement, Australia and Ms. Belinda Oelofse from Sika, Australia for providing the very-low-
C3A cement and the defoaming agent for the research, respectively. Authors also thank Mr.
Thomas Rossen, Mr. Liang Ming Lee, Mr. Beiming Zhang, and Mr. Joshua Ngo for their

42
Chapter 2

contributions during the laboratory work. The authors acknowledge the support provided
by Mr. Minhao Dong and the facilities given by Centre for Microscopy, Characterization &
Analysis, The University of Western Australia (UWA) for conducting the microstructure
analyses. The study was conducted while the first author was a recipient of a University
Postgraduate Award and the Australian Government Research Training Program
Scholarship at UWA.

References
Abd Elrahman, M., and Hillemeier, B. (2014). “Combined effect of fine fly ash and packing
density on the properties of high performance concrete: An experimental approach.”
Construction and Building Materials, 58, 225–233.
Abdulkareem, O. M., Ben Fraj, A., Bouasker, M., and Khelidj, A. (2018). “Effect of
chemical and thermal activation on the microstructural and mechanical properties of
more sustainable UHPC.” Construction and Building Materials, 169, 567–577.
ACI 318. (2014). Building Code Requirements for Structural Concrete. Farmington Hills, MI.
ACI 363R. (1992). State-of-the-Art Report on High-Strength Concrete.
Adjoudj, M., Ezziane, K., Kadri, E. H., Ngo, T. T., and Kaci, A. (2014). “Evaluation of
rheological parameters of mortar containing various amounts of mineral addition with
polycarboxylate superplasticizer.” Construction and Building Materials, 70, 549–559.
Ahmed, T., Elchalakani, M., Karrech, A., Dong, M., Mohamed Ali, M. S., and Yang, H.
(2021). “ECO-UHPC with High-Volume Class-F Fly Ash: New Insight into Mechanical
and Durability Properties.” Journal of Materials in Civil Engineering, 33(7), 0003726.
Andrade, D. da S., Rêgo, J. H. da S., Morais, P. C., Lopes, A. N. de M., and Rojas, M. F.
(2019). “Investigation of C-S-H in ternary cement pastes containing nanosilica and
highly-reactive supplementary cementitious materials (SCMs): Microstructure and
strength.” Construction and Building Materials, 198, 445–455.
AS/NZS 3582.2. (2016). Supplementary cementitious materials for use with portland and blended
cement - Slag—Ground granulated blastfurnace.
AS/NZS 3582.3. (2016). Supplementary cementitious materials for use with portland and blended
cement - amorphous silica.
AS/NZS 4456.17. (2003). Masonry units and segmental pavers and flags - methods of test:
determining initial rate of absorption (suction).
AS/NZS 4671. (2001). Steel reinforcing materials.
AS 1012.10. (2000). Methods of testing concrete: determination of indirect tensile strength of concrete
cylinders (“Brazil” or splitting test).
AS 1012.11. (2000). Methods of testing concrete: determination of the modulus of rupture.
AS 1012.17. (1997). Methods of testing concrete: determination of the static chord modulus of
elasticity and Poisson’s ratio of concrete specimens.
AS 1012.3.5. (2015). Methods of testing concrete: determination of properties related to the
consistency of concrete - slump flow, T500 and J-ring test.
AS 3600. (2009). Concrete structures.
ASTM C128 - 15. (2015). Standard Test Method for Relative Density (Specific Gravity) and
Absorption of Fine Aggregate. West Conshohocken, PA.
ASTM C150 / C150M-16. (2016). Standard specification for portland cement. West

43
Chapter 2

Conshohocken, PA.
ASTM C566 - 97. (1997). Standard test method for total evaporable moisture content of aggregate
by drying. West Conshohocken, PA.
ASTM C876-15. (2015). Standard Test Method for Corrosion Potentials of Uncoated Reinforcing
Steel in Concrete. West Conshohocken, PA.
ASTMC109 / C109M-16a. (2016). Standard test method for compressive strength of hydraulic
cement mortars (using 2-in. or [50-mm] cube specimens). West Conshohocken, PA.
Bach, Q. S. (2019). “Quantitative Study of Hydration of C3S and C2S in the Reactive
Powder Concrete together with its Strength Development.” Applied Mechanics and
Materials, 889, 294–303.
Bonavetti, V., Donza, H., Rahhal, V., and Irassar, E. (2000). “Influence of initial curing on
the properties of concrete containing limestone blended cement.” Cement and Concrete
Research, 30(5), 703–708.
Cheng, A., Huang, R., Wu, J. K., and Chen, C. H. (2005). “Influence of GGBS on durability
and corrosion behavior of reinforced concrete.” Materials Chemistry and Physics, 93(2–3),
404–411.
Elchalakani, M., Basarir, H., and Karrech, A. (2017). “Green concrete with high-volume fly
ash and slag with recycled aggregate and recycled water to build future sustainable
cities.” Journal of Materials in Civil Engineering, 29(2), 04016219.
Elchalakani, M., Dong, M., Karrech, A., Li, G., Mohamed Ali, M. S., Xie, T., and Yang,
B. (2018). “Development of Fly Ash- and Slag-Based Geopolymer Concrete with
Calcium Carbonate or Microsilica.” Journal of Materials in Civil Engineering, 30(12),
04018325.
EN 13295. (2004). Products and systems for the protection and repair of concrete structures - test
methods - determination of resistance to carbonation.
Gao, J. M., Qian, C. X., Liu, H. F., Wang, B., and Li, L. (2005). “ITZ microstructure of
concrete containing GGBS.” Cement and Concrete Research, 35(7), 1299–1304.
Gao, X., Yu, Q. L., and Brouwers, H. J. H. (2015). “Conceptual design and performance
evaluation of two-stage ultra-low binder ultra-high performance concrete.” Cement and
Concrete Research journal, 98(August), 397–406.
Ghafor, K., Qadir, S., Mahmood, W., and Mohammed, A. (2020). “Statistical variations
and new correlation models to predict the mechanical behaviour of the cement mortar
modified with silica fume.” Geomechanics and Geoengineering.
Goñi, S., Puertas, F., Hernández, M. S., Palacios, M., Guerrero, A., Dolado, J. S., Zanga,
B., and Baroni, F. (2010). “Quantitative study of hydration of C3S and C2S by thermal
analysis.” Journal of Thermal Analysis and Calorimetry, 102(3), 965–973.
Han-Seung, L., and Wang, X. Y. (2016). “Evaluation of compressive strength development
and carbonation depth of high volume slag-blended concrete.” Construction and Building
Materials, 124, 45–54.
IS 3495 (Part 2). (1992). Methods of tests of burnt clay building bricks, Part 2: determination of water
absorption. New Delhi.
Jaafar, M. F. M., Saman, H. M., Ariffin, N. F., Muthusamy, K., Ahmad, S. W., and Ismail,
N. (2018). “Corrosion monitoring on steel reinforced nano metaclayed-UHPC towards
strain modulation using fiber Bragg grating sensor.” IOP Conference Series: Materials
Science and Engineering, 431, 122006.
Johari, M. A. M., Brooks, J. J., Kabir, S., and Rivard, P. (2011). “Influence of
supplementary cementitious materials on engineering properties of high strength
concrete.” Construction and Building Materials, 25(5), 2639–2648.

44
Chapter 2

Kadri, E. H., Aggoun, S., and De Schutter, G. (2009). “Interaction between C3A, silica
fume and naphthalene sulphonate superplasticiser in high performance concrete.”
Construction and Building Materials, Elsevier Ltd, 23(10), 3124–3128.
Kim, Y. J. (2018). Development of Cost-Effective Ultra-High Performance Concrete (UHPC) for
Colorado’s Sustainable Infrastructure. Denver, CO.
Lawrence, P., Cyr, M., and Ringot, E. (2003). “Mineral admixtures in mortars: Effect of
inert materials on short-term hydration.” Cement and Concrete Research, 33(12), 1939–
1947.
Ling, G., Shui, Z., Sun, T., Gao, X., Wang, Y., Sun, Y., Wang, G., and Li, Z. (2018).
“Rheological behavior and microstructure characteristics of SCC incorporating
metakaolin and silica fume.” Materials, 11(12).
Ma, J., Orgass, M., Dehn, F., Schmidt, D., and Tue, N. V. (2004). “Comparative
Investigations on Ultra-High Performance Concrete with or without Coarse
Aggregates.” Proceedings of the international symposium on ultra high performance concrete,
205–212.
Mardani-Aghabaglou, A., Felekoğlu, B., and Ramyar, K. (2017). “Effect of cement C3A
content on properties of cementitious systems containing high-range water-reducing
admixture.” Journal of Materials in Civil Engineering, 29(8), 1–12.
Matschei, T., Bellmann, F., and Stark, J. (2005). “Hydration behaviour of sulphate-activated
slag cements.” Advances in Cement Research, 17(4), 167–178.
Meek, A. H., Beckett, C. T. S., Carsana, M., and Ciancio, D. (2018). “Corrosion protection
of steel embedded in cement-stabilised rammed earth.” Construction and Building
Materials, 187, 942–953.
Meier, M. R., Sarigaphuti, M., Sainamthip, P., and Plank, J. (2015). “Early hydration of
Portland cement studied under microgravity conditions.” Construction and Building
Materials, 93, 877–883.
Mohammadi, J., and South, W. (2016). “Measuring chromium in general purpose (GP)
cement.” Journal of Testing and Evaluation, 45(4), 1362–1377.
Naaman, A. E. (2011). “Half a century of progress leading to ultra-high performance fiber
reinforced concrete: Part 1-overall review.” 2nd Int. RILEM Conference on Strain Hardening
Cementitious Composites, Rio de Janeiro, Brazil, 12–14.
Neville, A. M. (2011). Properties of concrete. Pearson, London, UK.
Norhasri, M. S. M., Hamidah, M. S., Mohd Fadzil, A., and Megawati, O. (2016).
“Inclusion of nano metakaolin as additive in ultra high performance concrete (UHPC).”
Construction and Building Materials, 127, 167–175.
Oner, A., and Akyuz, S. (2007). “An experimental study on optimum usage of GGBS for
the compressive strength of concrete.” Cement and Concrete Composites, 29(6), 505–514.
Paul, S. C., Panda, B., Huang, Y., Garg, A., and Peng, X. (2018). “An empirical model
design for evaluation and estimation of carbonation depth in concrete.” Measurement:
Journal of the International Measurement Confederation, 124, 205–210.
Richard, P., and Cheyrezy, M. (1995). “Composition of reactive powder concretes.” Cement
and Concrete Research, 25(7), 1501–1511.
Sabet, F. A., Libre, N. A., and Shekarchi, M. (2013). “Mechanical and durability properties
of self consolidating high performance concrete incorporating natural zeolite, silica fume
and fly ash.” Construction and Building Materials, 44, 175–184.
Saillio, M., Baroghel-Bouny, V., Bertin, M., Pradelle, S., and Vincent, J. (2019). “Phase
assemblage of cement pastes with SCM at different ages.” Construction and Building
Materials, 224, 144–157.

45
Chapter 2

Samet, B., and Chaabouni, M. (2004). “Characterization of the Tunisian blast-furnace slag
and its application in the formulation of a cement.” Cement and Concrete Research, 34(7),
1153–1159.
Scheydt, J., and Müller, H. (2012). “Microstructure of ultra high performance concrete
(UHPC) and its impact on durability.” 3rd International Symposium on on Ultra High
Performance Concrete, 349–356.
Soliman, N. A., and Tagnit-Hamou, A. (2017). “Using glass sand as an alternative for quartz
sand in UHPC.” Construction and Building Materials, 145, 243–252.
Song, Q., Yu, R., Shui, Z., Wang, X., Rao, S., and Lin, Z. (2018). “Optimization of fibre
orientation and distribution for a sustainable Ultra-High Performance Fibre Reinforced
Concrete (UHPFRC): Experiments and mechanism analysis.” Construction and Building
Materials, 169, 8–19.
Sun, J., Shen, X., Tan, G., and Tanner, J. E. (2019). “Modification Effects of Nano-SiO2
on Early Compressive Strength and Hydration Characteristics of High-Volume Fly Ash
Concrete.” Journal of Materials in Civil Engineering, 31(6), 04019057.
Tabet, W. E., Cerato, A. B., Madden, A. S. E., and Jentoft, R. E. (2018). “Characterization
of hydration products’ formation and strength development in cement-stabilized
kaolinite using TG and XRD.” Journal of Materials in Civil Engineering, 30(10).
Tajuelo Rodriguez, E., Garbev, K., Merz, D., Black, L., and Richardson, I. G. (2017).
“Thermal stability of C-S-H phases and applicability of Richardson and Groves’ and
Richardson C-(A)-S-H(I) models to synthetic C-S-H.” Cement and Concrete Research, 93,
45–56.
Ting, L., Qiang, W., and Shiyu, Z. (2019). “Effects of ultra-fine ground granulated blast-
furnace slag on initial setting time, fluidity and rheological properties of cement pastes.”
Powder Technology, 345, 54–63.
Tironi, A., Trezza, M. A., Scian, A. N., and Irassar, E. F. (2013). “Assessment of pozzolanic
activity of different calcined clays.” Cement and Concrete Composites, 37(1), 319–327.
Tuan, N. Van, Ye, G., Breugel, K. Van, Fraaij, A. L. A., and Bui, D. D. (2011). “The study
of using rice husk ash to produce ultra high performance concrete.” Construction and
Building Materials, 25(4), 2030–2035.
Wu, Z., Shi, C., and Khayat, K. H. (2016). “Influence of silica fume content on
microstructure development and bond to steel fiber in ultra-high strength cement-based
materials (UHSC).” Cement and Concrete Composites, 71, 97–109.
Xie, T., Fang, C., Mohamad Ali, M. S., and Visintin, P. (2018). “Characterizations of
autogenous and drying shrinkage of ultra-high performance concrete (UHPC): An
experimental study.” Cement and Concrete Composites, 91, 156–173.
Xu, G., Tian, Q., Miao, J., and Liu, J. (2017). “Early-age hydration and mechanical
properties of high volume slag and fly ash concrete at different curing temperatures.”
Construction and Building Materials, 149, 367–377.
Yoo, D. Y., and Yoon, Y. S. (2015). “Structural performance of ultra-high-performance
concrete beams with different steel fibers.” Engineering Structures, 102, 409–423.
Yu, R., Spiesz, P., and Brouwers, H. J. H. (2015). “Development of an eco-friendly Ultra-
High Performance Concrete (UHPC) with efficient cement and mineral admixtures
uses.” Cement and Concrete Composites, 55, 383–394.
Zhu, Z., Li, B., and Zhou, M. (2015). “The influences of iron ore tailings as fine aggregate
on the strength of ultra-high performance concrete.” Advances in Materials Science and
Engineering, 2015, 412878.

46
CHAPTER
3
Mechanical properties and chloride penetration
resistances of very-low-C3A cement based SC-
UHP-SFRCs incorporating metakaolin and slag

Tanvir Ahmeda, Mohamed Elchalakania, Ali Karrecha, Riyadh Al-Amerib, Bo Yangc,d,*


a
Department of Civil, Environmental and Mining Engineering, Faculty of Engineering, Computing and Mathematical
Sciences, The University of Western Australia, 35 Stirling Highway, Crawley, WA 6009, Australia
b
School of Engineering, Deakin University-Waurn Ponds Campus, Geelong, VIC 3220, Australia
c
Key Laboratory of New Technology for Construction of Cities in Mountain Area, Chongqing University, Chongqing
400045, China
d
School of Civil Engineering, Chongqing University, Chongqing 400045, China

*Corresponding author: yang0206@cqu.edu.cn

Abstract

It can be observed from the results presented in Chapter 2, to attain a target flowability,
ultra-high performance concrete (UHPC) can be designed with lower water to binder ratio
(w/b) if cement with a very low C3A (tricalcium aluminate) content (≤3%) is used as the
primary binder instead of cement with a moderate C3A content (6%–10%). Hence, notably
better mechanical performance, as well as ample flowability, of UHPC can be achieved with
very-low-C3A cement (VLAC). Utilizing such beneficial features of VLAC, self-compacting
ultra-high performance steel fibre reinforced concrete (SC-UHP-SFRC) with a relative
slump flow of more than 5 (without jolting) and a 28-d compressive strength of more than
165 MPa (without special curing) has been synthesised in this study. In the marine
environment, however, VLAC based composite is not likely to exhibit a chemical chloride
binding capacity as good as that of a moderate-C3A cement (MAC) based composite. This
is because of the insufficient C3A content in VLAC, which limits the formation of Friedel’s
salt in the composite made with it. Therefore, the chloride penetration resistance of VLAC
based SC-UHP-SFRC has been aimed to be investigated in this study. The potentials of
supplementary binders which have higher alumina (Al2O3) contents than VLAC, such as
metakaolin and ground granulated blast-furnace slag (GGBS), to improve the chloride
Chapter 3

penetration resistance, as well as the mechanical properties, of VLAC based SC-UHP-SFRC


have also been aimed to be studied. Results indicate that inclusion of metakaolin replacing
40% of the silica fume in SC-UHP-SFRC with VLAC as the primary binder increases the
28-d compressive strength of the composite by 3.4%. Inclusion of GGBS substituting 30%
of the VLAC in SC-UHP-SFRC reduces the 28-d strength by 4.9%. The 28-d flexural
strength of SC-UHP-SFRC can be increased by 35.7% and 65.6% incorporating metakaolin
and GGBS, respectively. The level of chloride penetration resistance of VLAC based SC-
UHP-SFRC with metakaolin or GGBS can be ‘extremely high’, after 216 d of exposure to
cyclic drying and wetting with 10% NaCl solution. X-ray diffractometry (XRD) confirms
the formation of more Friedel’s salt in the mix with metakaolin or GGBS.

Keywords: apparent diffusivity, cyclic drying and wetting, fibre reinforced concrete,
Friedel’s salt, marine environment, tricalcium aluminate, ultra-high performance concrete

3.1 Introduction

Ultra-high performance steel fibre reinforced concrete (UHP-SFRC) is a relatively novel


class of steel fibre reinforced concrete. UHP-SFRC exhibits superior mechanical
performance compared to normal-strength steel fibre reinforced concrete (NS-SFRC) or
high-strength steel fibre reinforced concrete (HS-SFRC). NS-SFRC can have compressive
strength from 25 MPa to around 40 MPa, and flexural strength in the range of 4–7 MPa
(Khaloo and Kim 1996; Sakthivel and Vijay Aravind 2020). HS-SFRC can reach a
compressive strength of up to around 100 MPa, and its flexural strength typically ranges
from 7 MPa to as high as 15 MPa (Choi et al. 2019; Khaloo and Kim 1996; Song and Hwang
2004). UHP-SFRC, on the other hand, must have a compressive strength of at least 120 MPa
according to a group of researchers (Aghdasi and Ostertag 2018; Chu and Kwan 2019).
Some other researchers consider the minimum compressive strength to be 150 MPa (Cadoni
et al. 2019; Ngo et al. 2017). UHP-SFRC, moreover, can achieve a flexural strength of more
than 20 MPa (Kim et al. 2011; Song et al. 2018a; Yang et al. 2009). UHP-SFRC was also
observed to exhibit extraordinary performance in terms of durability. The water absorption
of UHP-SFRC was observed to be around 0.5% or less (Scheydt and Müller 2012), while
the water absorption of NS-SFRC is usually higher than 3.5% (Koushkbaghi et al. 2019;
Neville 2011). Specimens made from UHP-SFRC were observed to have a free chloride
concentration of less than 0.025% (by mass of concrete) at a depth of 4 mm from the surface,
after their exposure to 3% NaCl solution for 90 days, which was almost one order of
magnitude lower than the chloride concentration of normal-strength concrete specimens at
the same depth (Dobias et al. 2016).

48
Chapter 3

Several studies reported the benefits of using metakaolin and ground-granulated blast
furnace slag (GGBS) as partial substitutes of silica fume and cement, respectively, on the
compressive and flexural properties of ambient temperature cured UHP-SFRC
incorporating Portland cement with C3A content of more than 5% as the primary binder
(Ganesh and Murthy 2019; Song et al. 2018b; Wu et al. 2017; Yazıcı et al. 2010). Song et
al. (2018b) observed that the partial replacement of silica fume with metakaolin in UHP-
SFRC, having cement with 5.5% C3A content, improved both the 28-d compressive and
flexural strengths of the composite. The maximum strength increments (around 6% in
compression and 28% in flexure) were observed for 40% replacement. Ganesh and Murthy
(2019), Wu et al. (2017), and Yazıcı et al. (2010) investigated the influence of partially
replacing cement, containing C3A in the range of 6.8%–11.2%, with GGBS on the
compressive and flexural performances of UHP-SFRC. No notable improvement in the 28-
d compressive strength of UHP-SFRC was reported by the studies for partial replacement of
cement by GGBS. However, in the case of 28-d flexural strength, a maximum improvement
of around 7%–42% was reported for 20%–40% replacement. Few studies also investigated
the influence of partially replacing silica fume and cement with metakaolin and GGBS,
respectively, on the chloride ingress resistance of UHP-SFRC incorporating cement with
C3A content in the range of 5.5%–9.8% (Ganesh and Murthy 2019; Song et al. 2019). It is
worth noting that short-term electrically accelerated chloride migration tests, such as rapid
chloride migration (RCM) and rapid chloride penetration (RCP) tests, were conducted in
these studies to measure the chloride ingress resistance of the UHP-SFRCs incorporating
metakaolin and GGBS. Song et al. (2019) observed that the UHP-SFRC with cement – silica
fume – metakaolin binder system provided more resistance to rapid chloride migration than
the UHP-SFRC with cement – silica fume binder system. Ganesh and Murthy (2019) found
that the partial replacement of cement incorporating 9.8% C3A with GGBS improved the
resistance to rapid chloride penetration. The best performance was observed for 40%
replacement. Electrically accelerated chloride ingress tests are frequently used to measure
the chloride penetration resistance of concrete because of the considerably short time
required to conduct them. However, during conducting an electrical migration type test on
a steel fibre reinforced concrete specimen, the possibility of short-circuiting taking place by
the steel fibres cannot be ruled out (Stanish et al. 2001). In such a case, long-term chloride
ingress test may be more reliable for measuring the chloride penetration resistance of a steel
fibre reinforced concrete. Study addressing the resistance of UHP-SFRC with GGBS or
metakaolin to long-term chloride ingress is not available in the public literature.
Furthermore, there is a lack of a systematic study addressing the influence of using
metakaolin and GGBS as partial substitutes of silica fume and cement, respectively, on the

49
Chapter 3

mechanical properties and long-term chloride penetration resistance of UHP-SFRC with


VLAC as the primary binder.

Cements with moderate C3A content are more commonly used in Australia (Ahmed et al.
2021a; Mohammadi and South 2016). In Chapter 2 (Ahmed et al. 2021a), however, it can
be observed that, VLAC demands considerably less water than MAC to impart certain
flowability to UHPC [at a given water to binder ratio (w/b), VLAC based UHPC exhibits
34% higher slump flow than that of MAC based UHPC]. This allows VLAC based UHPC
to be designed with a lower w/b ratio and thus helps it attain improved mechanical
performance while maintaining good flowability at the fresh state. Utilizing such benefits of
VLAC, self-compacting UHP-SFRC (i.e., SC-UHP-SFRC) was prepared in this study
exhibiting a relative slump flow of more than 5 and a 28-d compressive strength of more
than 165 MPa. Although VLAC has the potential to impart good flowability and mechanical
performance to a cementitious composite, it may not impart a chloride binding capacity
similar to that imparted by MAC (Liu et al. 2021; Oh et al. 2003). C3A phase of MAC helps
reduce the amount of free chloride in a cementitious composite, incorporating MAC,
exposed to saline water by converting the chloride ions into Friedel’s salt
[Ca4Al2Cl2(OH)12·4H2O] (Liu et al. 2021). In a hydrated MAC based system, aluminium
bearing hydrate monosulphoaluminate (a hydrate resulting from C3A) transforms into
Friedel’s salt and Kuzel’s salt [Ca4Al2(SO4)0.5(Cl)(OH)12·6H2O] and thus helps bind the
chloride ions (Balonis et al. 2010). In contrast, VLAC lacks sufficient C3A which, as
discussed above, is an important phase for a cementitious composite to attain good chemical
chloride binding capacity. Silica fume is usually used as the only supplementary binder in
UHP-SFRC. However, silica fume may also not facilitate formation of any Friedel’s salt
(Guo et al. 2019). This is because silica fume mostly contains amorphous silica (SiO2) but
almost no aluminium and calcium. It is therefore necessary to investigate the chloride
penetration resistance of UHP-SFRC made with VLAC – silica fume binder system, in
addition to investigating its mechanical properties, so as to have an understanding of its
performance in the marine environment. Metakaolin and ground granulated blast furnace
slag (GGBS), owing to typically having higher alumina (Al2O3) contents than those of silica
fume and VLAC, may help produce more Friedel’s salt if they are incorporated in VLAC –
silica fume based SC-UHP-SFRC and thus play beneficial roles to improve the chloride
penetration resistance of the composite (Guo et al. 2019; Luo et al. 2003).

The present study investigated the effects of partial substitution of silica fume with
metakaolin (40%) and VLAC with GGBS (30%) on the flowability and mechanical
properties, such as compressive strength, flexural strength, and load-deflection behaviour,

50
Chapter 3

of SC-UHP-SFRC incorporating VLAC as the primary binder. The study further


investigated the influence of incorporating metakaolin and GGBS on the transport
properties of VLAC based SC-UHP-SFRC such as water absorption and chloride
penetration after exposure to cyclic drying and wetting with 10% NaCl solution for 216 d.
The microstructures of selected SC-UHP-SFRC samples were also investigated, by means
of scanning electron microscopy (SEM) and X-ray diffractometry (XRD), to better
understand their macro-behaviours.

3.2 Materials and methods

3.2.1 Materials

A very-low-C3A cement (VLAC), containing a C3A content of 1.1% and conforming to the
requirements of ASTM C150 / C150M-19a Type I Portland cement (ASTM C150 / C150M-
16 2016), was used as the primary binder in this study. The supplementary binders used in
this study include silica fume conforming to the requirements of AS/NZS 3582.3 (2016),
metakaolin, and GGBS conforming to AS/NZS 3582.2 (2016). Figure 3.1 presents the XRD
patterns of the binders. The crystalline phases present in the VLAC were mainly alite
(Ca3SiO5 or, C3S), belite (Ca2SiO4 or, C2S), ferrite [Ca4(Al,Fe)2O7 or, C4AF], and gypsum
(CaSO4·2H2O). The broad humps in the 2θ ranges of around 16–29°, 16–31°, and 25–35° in
the XRD patterns of silica fume, metakaolin, and GGBS, respectively, indicate that these
binders were mainly composed of amorphous phases. Nevertheless, some crystalline phases
like quartz (SiO2), silicon carbide (SiC) were identified in the silica fume, anatase (TiO2) was
identified in the metakaolin, and gypsum, akermanite (Ca2MgSi2O7), gehlenite (Ca2Al2SiO7)
were found in the XRD pattern of the GGBS. It is worth noticing that the gypsum peak
intensity at around 11.5° was higher for GGBS than for VLAC, which implies higher
gypsum content in the former. Quartz sand with a median diameter (d50) of 237.3 µm, and
quartz powder with a d50 of 41.5 µm were used as aggregates. The chemical properties of the
binders and the aggregates are presented in Table 3.1. Figure 3.2 shows the particle size
distributions of the materials. A polycarboxylate based high-range water reducer (HRWR)
(with a specific gravity of 1.07 and a solid content of 37% by mass) and a defoaming agent
(with a specific gravity of 1 and a solid content of 5.5%) were used as chemical admixtures.
Brass coated straight steel fibres with a diameter of around 0.22 mm and a length of around
13 mm were used to prepare the UHP-SFRC mixes (Figure 3.3). The tensile strength of the
fibres was around 2850 MPa.

51
Chapter 3

A, B VLAC Silica fume


A A Metakaolin GGBS
G A A
F A A F A A
AA A A

Q
Intensity

SC

At

G Ak, Ge

5 10 15 20 25 30 35 40 45 50 55 60 65 70
2θ (degree)

Figure 3.1. XRD patterns of as received binders [G = gypsum, A = alite, B = belite, F =


ferrite, Q = quartz, SC = silicon carbide, At = anatase, Ak = akermanite, Ge = gehlenite]

100
VLAC d50 = 3.2 μm
90
Silica fume
80 Metakaolin
70 GGBS
Quartz powder
60 d50 =
Quartz sand
% Finer

17.3 μm
50
40
d50 = 41.5 μm
30 d50 = 1.7 μm d50 =
237.3 μm
20
10 d50 = 8.0 μm

0
0.1 1 10 100 1000
Particle size (μm)

Figure 3.2. Particle size distributions of materials [d50 = median diameter]

52
Chapter 3

Figure 3.3. Steel fibres used in this study

Table 3.1. Chemical properties of binders and aggregates

Material VLAC Silica Metakaolin GGBS Quartz Quartz


fume powder sanda
Oxides (mass Al2O3 3.18 0.13 45.9 13.04 0.34 0.017
%) CaO 64.48 0.33 0.02 42.80 0.41 0.002
Fe2O3 4.34 0.07 0.49 0.81 0.10 0.008
K2O 0.35 0.54 0.15 0.3 0.085 0.003
MgO 1.46 0.71 0.04 5.46 0.027 0.003
MnO 0.12 0.008 N/M 0.17 0.001 0.001
Na2O 0.14 0.29 0.14 0.28 0.027 0.003
P2O5 0.228 0.129 N/M 0.02 0.003 0.0
SiO2 21.4 94.6 51.7 31.14 95.5 99.88
SO3 2.93 0.07 0.0 4.02 0.0 0.0
TiO2 0.192 0.005 0.92 0.56 0.013 0.025
Loss 1.15 3.1 0.54 1.4 3.45 0.05
on
ignition

Cement phases C3S 63.7 - - - - -


(mass %)b C2S 13.4
C3A 1.1
C4AF 13.2

VLAC = very-low-C3A cement, C3S = Ca3SiO5 (alite), C2S = Ca2SiO4 (belite), C3A = Ca3Al2O6
(tricalcium aluminate), C4AF = Ca4(Al,Fe)2O7 (ferrite), N/M = not measured
a
As per manufacturer
b
Based on Bogue’s equations (ASTM C150 / C150M-16 2016)

53
Chapter 3

3.2.2 Mixes

The proportions of the UHPC and UHP-SFRC mixes investigated in this study are presented
in Table 3.2. Mixes R-F0, R-F1, and R-F2 were synthesised with a binder system containing
VLAC and silica fume. To produce mixes MK-F0 and MK-F2, 40% of the silica fume was
replaced with the metakaolin; whereas to produce mixes S-F0 and S-F2, 30% of the VLAC
was replaced with the GGBS. The w/b was kept constant at 0.193 in all the mixes. For the
convenience of discussion, the combined Al2O3 content and the combined CaO content in
the binder system present in each mix are also presented in Table 3.2.

The combined PSDs of the UHPCs (i.e., R-F0, MK-F0, and S-F0) are shown in Figure 3.4.
Figure 3.4 also shows the target PSD based on modified Andreasen and Andersen (A&A)
model (Equation 3.1). If compared with the target PSD, the particle packing was better for
S-F0 than for R-F0 and MK-F0. The particle packing of R-F0 was slightly better than that
of MK-F0.

𝑑𝑞 − 𝑑𝑚𝑖𝑛 𝑞
𝑝(𝑑) = (3.1)
𝑑𝑚𝑎𝑥 𝑞 − 𝑑𝑚𝑖𝑛 𝑞

d is particle size (µm), p(d) represents cumulative percent finer than size d, dmin and dmax are
respectively the minimum and maximum sizes of particle (µm). q was selected to be 0.23 as
per the recommendations of previous studies (Song et al. 2018a; Yu et al. 2015).

100
R-F0
90
MK-F0
80 S-F0
Modified A&A
70

60
% Finer

50

40
RMSE = 3.427
30

20
RMSE = 4.622
10 RMSE = 5.052

0
0.1 1 10 100 1000
Particle size (μm)

Figure 3.4. Combined particle size distributions (PSDs) of UHPCs [RMSE = root-mean-square
error between target curve and combined PSD]

54
Chapter 3

Table 3.2. Mix proportions of UHPCs and UHP-SFRCs (by mass)

Mix ID. R-F0 R-F1 R-F2 MK- MK- S-F0 S-F2


F0 F2
Constituents (by VLAC 1 1 1 1 1 0.7 0.7
mass) SF 0.27 0.27 0.27 0.16 0.16 0.27 0.27
MK - - - 0.11 0.11 - -
S - - - - - 0.3 0.3
QP 0.22 0.22 0.22 0.22 0.22 0.22 0.22
QS 0.88 0.88 0.88 0.88 0.88 0.88 0.88
W 0.245 0.245 0.245 0.245 0.245 0.245 0.245
HRWRa 0.016 0.016 0.016 0.016 0.016 0.016 0.016
DAa 0.0014 0.0014 0.0014 0.0014 0.0014 0.0014 0.0014
Fb - 1% 2% - 2% - 2%
Mix design mk/sf 0/100 0/100 0/100 40/60 40/60 0/100 0/100
parameters s/c 0/100 0/100 0/100 0/100 0/100 30/70 30/70
w/b 0.193 0.193 0.193 0.193 0.193 0.193 0.193
Combined Al2O3 or Al2O3 2.53 2.53 2.53 6.50 6.50 4.86 4.86
CaO content in
CaO 50.84 50.84 50.84 50.81 50.81 45.72 45.72
binder system
(mass %)
VLAC = very-low-C3A cement, SF = silica fume, MK = metakaolin, S = GGBS, QP = quartz powder,
QS = quartz sand, W = water, HRWR = high-range water reducer, DA = defoaming agent, F = fibre,
mk/sf = metakaolin to silica fume ratio, s/c = GGBS to VLAC ratio, w/b = water to binder ratio
a
Solid content
b
Volume fraction

3.2.3 Mixing, specimen preparation, and curing

To prepare each UHPC mix, first, the binders and the aggregates were mixed for 5 min at a
speed of 60 rpm. Next, 100% of the water, 80% of the HRWR, and 100% of the defoaming
agent were added to the dry mix. The mixing was continued for another 3 min, and then the
remaining 20% of the HRWR was added. When the mix appeared to be moist, the mixing
speed was increased to 180 rpm. The mixing was continued until the mix became flowable.
After that the mixer speed was further increased to 300 rpm and the mixing was continued
for 3 min. To prepare each UHP-SFRC mix, following the aforementioned steps, fibres were
incorporated in the mix and then mixed for 1 min. The mix prepared was then used to
perform slump flow test and fill up moulds.

50 × 50 × 50 mm cube moulds and 40 × 40 × 160 mm prism moulds were filled to prepare


specimens for compressive strength test and four-point bending test. 100 × 100 × 100 mm
cube moulds were filled with selected UHP-SFRC mixes to prepare specimens for water
absorption and chloride penetration tests. The freshly poured UHPCs/ UHP-SFRCs were
self-consolidating. Therefore, the moulded mixes were not vibrated. The moulds were

55
Chapter 3

moved to a curing room and kept at a temperature of 21 °C and a relative humidity of 95%.
All the specimens were demoulded after 24 h.

The mechanical test specimens (i.e., the 50 × 50 × 50 mm cubes and 40 × 40 × 160 mm


prisms), after being demoulded, were kept submerged in water at a temperature of 21 °C
until the test day. The 100 × 100 × 100 mm specimens, on the other hand, were kept in the
curing room. After 26 d, the 100 × 100 × 100 mm cubes prepared for capillary water
absorption and water absorption tests were oven-dried at a temperature of 110 °C for 48 h,
and the cubes prepared for chloride penetration tests were oven-dried at a temperature of 40
°C for 48 h. The oven-dried specimens prepared for capillary water absorption and chloride
penetration tests were then sealed with epoxy resin and duct tape on all the faces (including
the unfinished top) except one, while the specimens prepared for water absorption tests were
left unsealed. After 28 d, the 100 × 100 × 100 mm cubes prepared for chloride penetration
tests were subjected to cyclic wetting and drying. The wetting phase of each cycle involved
48 h of immersion of the cubes in 10% NaCl solution (as shown in Figure 3.5), and the
drying phase involved 48 h of drying in the atmospheric condition. Figure 3.6 shows the
daily temperatures and relative humidities, sourced from the Australian Government
Bureau of Meteorology (2021), at which the specimens were dried. Equal durations were
chosen for wetting and drying because the ratio of flood tide duration to that of the ebb tide
is approximately 1:1. A total of 54 cycles were completed in a period of 216 d.

3.2.4 Test methods

Slump flow tests were conducted for the fresh UHPC and UHP-SFRC mixes as per the
guidelines in AS 1012.3.5 (AS 1012.3.5 2015). The fresh mixes were not tamped, jolted, or
vibrated before slump flow measurement. A mini cone, with a height of 54 mm, and top and
bottom diameters of respectively 70 mm and 100 mm, was used to perform the slump tests.
The relative slump flow of each mix was calculated using Equation 3.2 (EFNARC 2002;
Tuaum et al. 2018).

𝑑𝑠 2
𝛤=( ) −1 (3.2)
𝑑𝑏

ds is the slump flow diameter (mm), db is the bottom diameter of the slump cone.

The compressive strength tests of the UHPCs and UHP-SFRCs were carried out at 7 and 28
d according to the guidelines of ASTM C109 / C109M-16a (2016). The four-point bending
tests were performed at 28 d as per the guidelines of AS 1012.11 (2000). Three specimens
were tested to evaluate the mechanical property of each mix.

56
Chapter 3

Figure 3.5. Specimen setup during wetting with 10% NaCl solution

Figure 3.6. Temperature and relative humidity during drying phase

57
Chapter 3

Capillary water absorption (CWA) tests were carried out for the UHP-SFRCs at the age of
28 d as per the test method mentioned in ASTM C1585-04 (2004) using 100 × 100 × 100
mm cubes sealed on five faces. The setup used for the tests is shown in Figure 3.7. Equation
3.3 was used to determine the CWA of a UHP-SFRC cube at a certain time. Water
absorption (WA) tests were conducted using the unsealed 100 × 100 × 100 mm cubes in
accordance with the guidelines of IS 3495 (Part 2) (1992). The WA of each unsealed UHP-
SFRC cube was calculated using Equation 3.4. Three specimens were tested for evaluating
the CWA or WA of each mix.

(𝑚2 − 𝑚1 ) (mm3/mm2) (3.3)


𝐶𝑊𝐴 =
𝐴𝜌𝑤

100(𝑚4 − 𝑚3 ) (%) (3.4)


𝑊𝐴 =
𝑚3

m1 is the mass of specimen after being oven-dried and then sealed on five faces (g), m2 is the
mass of sealed specimen after immersion of the exposed face in the water (3 ± 1 mm) for a
certain time (g) (Figure 3.7), A is the area of the exposed face (mm2), ρw is the density of
water (g/mm3), m3 is the mass of unsealed specimen after being oven-dried (g), m4 is the
mass of unsealed specimen after 24 h of submersion in water (g). Damp cloth was used to
wipe off the excess water on the specimen surface prior to measuring m2 or m4.

Chloride penetration tests were conducted following a modified version of the procedure
mentioned in ASTM C1556-11a (2016). 100 × 100 × 100 mm UHP-SFRC cubes sealed on
five faces were oven-dried at a temperature of 40 °C for 48 h, after being subjected to cyclic
drying and wetting with 10% NaCl solution for 216 d (as mentioned in Section 3.2.3). After
that, powder samples were collected from different depths of each UHP-SFRC specimen by
drilling three holes along the diagonal of the specimen (as shown in Figure 3.8). It is worth
mentioning that the diagonal of each specimen, along which the holes were drilled, was kept
at a constant depth from the top of the 10% NaCl solution during each wetting phase (Figure
3.5). 1.5 g of concrete powder, collected from each of the 7 consecutive depths of each hole
(as shown in Figure 3.8), was mixed with a mixture of 100 mL deionized water and 10 mL
ionic strength adjuster solution (1 M NaNO3). The new mixture was then stirred for around
20 min and a chloride ion-selective electrode, connected to an ion meter, was immersed in
the mixture. The free chloride concentration reading obtained from the ion meter for the
mixture was converted to percentage chloride concentration per unit mass of concrete
powder. Figure 3.9 shows the chloride concentration measurement setup. The free chloride
concentrations measured in the aforementioned way for three powder samples collected

58
Chapter 3

from equal depths of the three holes of a UHP-SFRC specimen were averaged to determine
the chloride concentration of that UHP-SFRC at that particular depth.

Thin sections were prepared from selected UHP-SFRC samples, cured for 28 d, and then
thin layers of platinum coating were applied on the thin sections for SEM analysis. SEM
images were captured using a field emission scanning electron microscope (FE-SEM). XRD
analysis was conducted for UHP-SFRC powders collected from a depth of 7.5 mm of
selected specimens which were exposed to cyclic drying and wetting with 10% NaCl solution
for 216 d. For each powdered sample, XRD pattern in the 2θ range of 5–70° was recorded
using an X-ray diffractometer with Cu K radiation (λ = 1.5404 Å), 40 kV, and 40 mA at a
rate of 2°/min.

100 × 100 × 100 mm


UHP-SFRC cube
sealed on 5 faces

Supporting
bars
3 ± 1 mm
Water

Exposed surface

Figure 3.7. Capillary water absorption test setup

Figure 3.8. Drilling depths of UHP-SFRC cubes for chloride penetration test

59
Chapter 3

Figure 3.9. Chloride concentration measurement for UHP-SFRC samples

3.3 Results and discussion

3.3.1 Flowability

Figure 3.10 presents the slump flow results of the UHPC and UHP-SFRC mixes prepared
in this study. Inclusion of straight steel fibres in VLAC based UHPC mix up to 2% by
volume, so as to prepare UHP-SFRC mixes, marginally changed its flowability.
Nevertheless, the slump flow diameter of R-F1 was 2.1% higher than that of R-F0. The high
specific gravity of the steel fibres possibly enabled mix R-F1 to spread more than R-F0 (Hung
et al. 2020). Further inclusion of steel fibres, i.e., increasing the fibre content from 1% to 2%,
most likely increased the interaction between the fibres. Hence R-F2 was observed to have
a slump flow slightly lower (2.4%) than that of R-F1 and comparable to that of R-F0.

The slump flow diameters of MK-F0 and MK-F2 were higher by 6.6% and 6% with respect
to those of R-F0 and R-F2, respectively. Such observation can be attributed to the less
HRWR demand of metakaolin than that of silica fume (Ding and Li 2002; Kadri et al. 2011).
The slump flow of S-F0 was 3.5% lower with respect to R-F0. This can be attributed to the
finer size of the GGBS, than that of the VLAC, which possibly imparted higher yield stress
to fresh S-F0 (Bentz et al. 2012; Ting et al. 2019). The observation is consistent with the
results presented in Chapter 2 (Ahmed et al. 2021a). Marginal improvement (2.2%) was
observed in the slump flow of S-F2, with respect to S-F0. A possible reason could be that
higher fibre content increased the fresh unit weight of S-F2 and helped it spread more,
overcoming somewhat the yield stress induced by GGBS inclusion (Hung et al. 2020). With
respect to R-F2 and MK-F2, the slump flow of S-F2 was 1.1% and 6.7% lower, respectively.

60
Chapter 3

Figure 3.10 further presents the relative slump flow results of the UHPC and UHP-SFRC
mixes. The relative slump flow values of all the mixes were higher than 4.76, which is the
minimum relative slump flow requirement for self-compacting mortar (EFNARC 2002;
Tuaum et al. 2018). It is worth noting that the fresh mixes were not tamped or jolted before
flow measurement. The relative slump flow values of UHPC and UHP-SFRC mixes with
VLAC – silica fume binder system were in the range of 5.59–5.92. For the mixes with
metakaolin and GGBS, the relative slump flow values were in the ranges of 6.41–6.54 and
5.18–5.45, respectively. The photographs of typical spreads of R-F2, MK-F2, and S-F2 are
shown in Figure 3.11.

350 10
Slump flow
Relative slump flow 9
300

Relative slump flow


Slump flow (mm)

8
250
7
200
6

150
5

100 4
R-F0 R-F1 R-F2 MK-F0 MK-F2 S-F0 S-F2

Figure 3.10. Slump flow and relative sump flow results of UHPCs and UHP-SFRCs

Figure 3.11. Typical spreads of (a) R-F2, (b) MK-F2, and (c) S-F2

61
Chapter 3

200
7-d 28-d

Compressive strength (MPa)


160

120

80

40

0
R-F0 R-F1 R-F2 MK-F0 MK-F2 S-F0 S-F2

Figure 3.12. Compressive strength results of UHPCs and UHP-SFRCs

3.3.2 Mechanical properties

3.3.2.1 Compressive strength

The 7-d and 28-d compressive strengths of the UHPCs and UHP-SFRCs are presented in
Figure 3.12. The 7-d and 28-d compressive strengths of R-F0 were 115.1 MPa and 164.3
MPa, respectively. Mixes R-F1 and R-F2 exhibited higher compressive strengths, both at 7
and 28 d, with respect to those of R-F0. The 28-d strengths of R-F1 and R-F2 were
respectively 8.5% and 5.9% higher with respect to the 28-d strength of R-F0. Incorporation
of steel fibres most likely delayed the formation of micro-cracks in R-F1 and R-F2 under
compression, compared with R-F0, which resulted in improved compressive strengths in the
UHP-SFRCs (Yoo et al. 2014). Slightly lower compressive strength of R-F2 than that of R-
F1 was perhaps a result of increased interaction or agglomeration of fibres in the former
(Meng and Khayat 2018). The 7-d and 28-d compressive strengths, moreover, were higher
for UHP-SFRC mixes MK-F2 and S-F2 than for the UHPC mixes MK-F0 and S-F0,
respectively. The 28-d strengths of MK-F2 and S-F2 were respectively 6.3% and 3.8% higher
with respect to those of MK-F0 and S-F0. In general, increase in the compressive strength
of UHPC due to the inclusion of steel fibres was also confirmed by previous studies (Meng
and Khayat 2018; Ren et al. 2018; Wu et al. 2016).

Figure 3.12 further shows that the UHPC and UHP-SFRC mixes incorporating metakaolin
(i.e., mixes MK-F0 and MK-F2, respectively) exhibited higher compressive strengths than
the reference mixes (i.e., mixes R-F0 and R-F2, respectively). The increases in the
compressive strengths, owing to the incorporation of metakaolin, were found to be more
pronounced at 7 d than at 28 d. With respect to R-F0 and R-F2, the 28-d compressive

62
Chapter 3

strengths of MK-F0 and MK-F2 were 3% and 3.4% higher, while the 7-d compressive
strengths of the mixes were 7.9% and 7.8% higher, respectively. High early strength gains of
the mixes with metakaolin, compared to the respective reference mixes, can be attributed to
the higher pozzolanic reactivity of metakaolin than that of silica fume at early ages (Akcay
and Tasdemir 2018), and to the accelerated cement hydration in the presence of metakaolin
(Akcay and Tasdemir 2018; Kadri et al. 2011). The mixes containing GGBS exhibited lower
compressive strengths than the reference mixes. Such observation is consistent with the
results presented in Chapter 2 (Ahmed et al. 2021a). Nevertheless, the differences between
the compressive strengths were lower at 28 d than at 7 d. This can be attributed to the
increased pozzolanic reactivity of GGBS in S-F0 and S-F2 between 7 and 28 d compared to
its pozzolanic reactivity before 7 d (Ahmed et al. 2021a; Oner and Akyuz 2007). With
respect to R-F0 and R-F2, the 7-d compressive strengths of S-F0 and S-F2 were 6.1% and
10.2% lower, and their 28-d strengths were 2.9% and 4.9% lower, respectively. It was
interesting to notice that all the UHP-SFRCs attained strengths of more than 120 MPa at
the age of 7 d. At 28 d, the UHPCs attained strengths of more than 155 MPa, and the UHP-
SFRCs attained strengths of more than 165 MPa, without employing any special curing.

3.3.2.2 Load – deflection behaviour under flexure

Figure 3.13 shows the load versus mid-span deflection curves of the UHPC and UHP-SFRC
beams under four-point bending. Table 3.3 presents the flexural strength (i.e., the maximum
flexural stress or fr) of each UHPC or UHP-SFRC. The first cracking strength (fcr), deflection
capacity at the point of first cracking load (δcr), deflection capacity at the maximum load (δu),
and toughness indices ηu and ηL/40 are also presented in Table 3.3 for each of the mixes. ηu
and ηL/40 of each mix were calculated by dividing the areas up to the deflections δu and L/40
(i.e., span length/40 = 3 mm), respectively, by the area up to δcr (Wu et al. 2016).

Figure 3.13(a) shows that the load – deflection curve of any of the UHPCs had almost no
inelastic softening part, which is indicative of the very brittle nature of the UHPCs. The
flexural strength of mix S-F0 was lower than both R-F0 and MK-F0 (6.8% and 12.3% lower,
respectively). Incorporation of 1% steel fibre in the reference UHPC (i.e., mix R-F0), so as
to prepare R-F1, led to an increase of 11.2% in the flexural strength. The post-cracking
strength of R-F1 was almost similar to its first cracking strength [Figure 3.13(b)]. In other
words, a deflection hardening part was absent in the load – deflection curve of R-F1. This
observation is in agreement with the results obtained in (Ren et al. 2018; Wu et al. 2016) for
UHP-SFRCs made with 1% short straight steel fibre. The fr of R-F2 increased by 21.2% with
respect to the fcr of the composite. Nevertheless, the deflection hardening part of the load –

63
Chapter 3

deflection curve of R-F2 was not pronounced and the accompanied multiple cracks were
few, if compared with MK-F2 and S-F2 [Figures 3.13(c), 3.13(d), and 3.13(e)]. The fr, δu, ηu,
and ηL/40 of MK-F2 were respectively 35.7%, 84.6%, 127.9%, and 19.4% higher, with respect
to those of R-F2. The fr of MK-F2 was 56.3% higher with respect to its fcr. Better performance
of MK-F2 after the formation of the first crack than that of R-F2 was perhaps a result of
higher pozzolanic activity of the metakaolin at early ages than that of the silica fume (Akcay
and Tasdemir 2018). Higher pozzolanic activity of metakaolin at early ages tends to help
consume more CH (portlandite) crystals in the fibre – matrix interfaces of a composite at 28
d and convert them into C-S-H (calcium silicate hydrate) and/or C-A-S-H (calcium
aluminate silicate hydrate) (Song et al. 2019). Consequently, improved fibre – matrix
interfaces are likely to be obtained for a composite with cement – silica fume – metakaolin
binder system at 28 d, compared to a composite with cement – silica fume binder system.
The deflection hardening part of the load – deflection curve of S-F2 was even more
pronounced. The fr, δu, ηu, and ηL/40 of S-F2 were respectively 65.6%, 136.8%, 242.3%, and
30.1%, with respect to those of R-F2. Partial substitution of VLAC by finer GGBS possibly
improved the fibre – matrix interfaces of S-F2 by facilitating better particle packing (Figure
3.4), and converting more portlandite into C-S-H (Ahmed et al. 2021a; Zhou et al. 2010).
Furthermore, softer matrix of S-F2 (as indicated by the lower fcr of S-F2 than that of R-F2 in
Table 3.3) perhaps played a role to improve its behaviour in the deflection hardening zone,
as a larger margin between the minimum fibre bridging strength and the first cracking
strength tends to generate more cracks (Zhou et al. 2010).

Table 3.3. Flexural behaviours of UHPCs and UHP-SFRCs

Mix fcr (MPa) fr (MPa) δcr (mm) δu (mm) ηu ηL/40


R-F0 20.3 20.3 0.180 0.180 1 1
R-F1 22.6 22.6 0.213 0.213 1 16.23
R-F2 20.7 24.6 0.181 0.296 2.52 26.50
MK-F0 21.5 21.5 0.186 0.186 1 1
MK-F2 21.3 33.4 0.187 0.547 5.73 31.64
S-F0 18.9 18.9 0.210 0.210 1 1
S-F2 19.4 40.7 0.210 0.702 8.61 34.48
fcr = first cracking strength, fr = flexural strength, δcr = deflection at first cracking load, δu = deflection at
maximum load, ηu = toughness index at δu, L = span length, ηL/40 = toughness index at a deflection of
L/40

64
Chapter 3

Figure 3.13. Load – deflection curves of (a) R-F0, MK-F0, S-F0, (b) R-F1, (c) R-F2, (d) MK-
F2, and (e) S-F2

65
Chapter 3

Figure 3.14. Water transport properties of UHP-SFRCs – (a) capillary water absorption, (b)
initial sorptivity (Si) in mm/s0.5, (c) secondary sorptivity (Ss) in mm/s0.5, and (d) water
absorption

3.3.3 Transport properties

3.3.3.1 Water absorption

Figure 3.14(a) shows the capillary water absorption of each UHP-SFRC plotted against the
square root of time. The slope of the straight line that best fit the points from 1 min to 6 h
(i.e., from 7.75 to 146.97 s0.5) indicates the initial sorptivity, while the slope of the line best
fitting to the points from 1 to 7 d (i.e., from 293.94 to 777.69 s0.5) indicates the secondary
sorptivity. The calculated initial and secondary sorptivity values are presented in Figures
3.14(b) and 3.14(c), respectively. In each of Figures 3.14(b) and 3.14(c), the correlation
coefficients between the data points and the respective lines of best fit were higher than 0.98.

66
Chapter 3

The initial sorptivity of R-F2 was found to be lower than R-F1. Incorporation of higher fibre
content possibly interrupted the continuity of more connected pores in R-F2 and
consequently lower initial sorptivity was observed for R-F2 than for R-F1 (Abbas et al.
2015). The secondary sorptivity of R-F2 was also found to be lower than that of R-F1.
Among the UHP-SFRCs with 2% fibre content, S-F2 resulted in the lowest initial and
secondary sorptivities. This may be because better particle packing was achieved for S-F2
than for R-F2 or MK-F2 (inferred from the comparison between the target PSD and the
combined PSDs of R-F0, MK-F0, and S-F0 presented in Figure 3.4). Another reason could
be that VLAC replacement by GGBS reduced the formation of shrinkage cracks in UHP-
SFRC (Ahmed et al. 2021b). The initial sorptivity of MK-F2 was slightly lower than that of
R-F2, although Figure 3.4 indicates that slightly better particle packing was achieved for the
latter. This is possibly because partial replacement of the silica fume with the metakaolin
resulted in lower shrinkage induced microcracking in MK-F2 (Akcay and Tasdemir 2018;
Güneyisi et al. 2010; Staquet and Espion 2004). Less microcracking in MK-F2 caused water
to be absorbed less rapidly in the initial stage (Ghasemzadeh and Pour-Ghaz 2015).
However, the secondary sorptivity of MK-F2 increased slightly with respect to that of R-F2.
This may be attributed to the less dense packing of particles in MK-F2 (inferred from Figure
3.4) and to the lower pozzolanic reactivity of metakaolin at later ages (Curcio et al. 1998).

The water absorptions of the unsealed UHP-SFRC cubes after 24 h of complete immersion
in the water are presented in Figure 3.14(d). The results show that the water absorption of
R-F2 was 40.7% lower with respect to that of R-F1. The results are in agreement with the
sorptivity results of R-F2 and R-F1. The water absorption of S-F2 was 25.7% lower with
respect to R-F2, and the water absorption of MK-F2 was 8.6% higher than that of R-F2. It
is worth noticing that water absorption levels of all these UHP-SFRCs were well below 2%,
which is the maximum limit for concrete being considered as durable for construction works
with 50-year design life (Elchalakani et al. 2017).

3.3.3.2 Chloride penetration resistance

The chloride concentrations in the powders collected from different depths of UHP-SFRC
cubes, subjected to continuous drying and wetting with 10% NaCl solution for 216 d, are
presented in Figure 3.15. The chloride concentration profile for each UHP-SFRC was
constructed by adjusting the error function solution of the Fick’s second law of diffusion to
the measured chloride concentrations of the UHP-SFRC by means of non-linear regression.
Equations 3.5 and 3.6 represent the Fick’s second law of diffusion and its error function
solution, respectively.

67
Chapter 3

𝜕𝐶 𝜕2𝐶
= 𝐷𝑎 2 (3.5)
𝜕𝑡 𝜕𝑥

C is the chloride concentration, t is the time, Da is the apparent diffusivity (Da), and x is the
depth of penetration.

𝑥
𝐶(𝑥, 𝑡) = 𝐶𝑠 − (𝐶𝑠 − 𝐶𝑖 )erf ( ) (3.6)
2√𝐷𝑎 𝑡

C(x,t) is the chloride concentration at any time t and depth x, Cs is the surface chloride
concentration, Ci is the background chloride concentration, erf(z) is the Gaussian error
2 𝑧 −𝑢2
function defined as erf(𝑧) = ∫ 𝑒 𝑑𝑢
√𝜋 0

Based on the chloride concentration profile of S-F2, its chloride concentration at any depth
was lower than 0.045% by mass of concrete (1% by mass of binder), which according to (Oh
et al. 2003) is the threshold free chloride concentration for corrosion initiation. Hence, based
on the chloride profile of S-F2, no cover is theoretically required for the rebars embedded in
S-F2. In other words, the value of clear cover α (as shown in Figure 3.15) would be zero.
The theoretical values of α would be 6.1 mm, 0.6 mm, and 4.8 mm for R-F1, R-F2, and MK-
F2, respectively. Further investigation at structural scale is still essential to identify the actual
required αs for marine structures built with the UHP-SFRCs. The maximum free chloride
concentration near the surface was found for MK-F2 (0.098% of concrete, at the surface).
There could be few reasons behind such observation. First, the particles in MK-F2 were less
densely packed than those in R-F2 and S-F2 (Figure 3.4). Second, the pozzolanic reactivity
of the metakaolin was lower than that of the silica fume at later ages (Curcio et al. 1998),
which caused the process of pore structure refinement through the formation of secondary
C-S-H to be less pronounced in MK-F2, compared with the other UHP-SFRCs, between 28
d and 246 d (28 d of standard curing + 216 d of cyclic drying and wetting + 2 d of oven-
drying at 40 °C). Third, because of the high Al2O3 content of the metakaolin, it helped trap
more chloride ions near the surface of MK-F2 through the formation of Friedel’s salt
(Seleem et al. 2010). However, a portion of the bound chloride near the surface perhaps got
released due to carbonation during the drying process (Easterbrook et al. 2015). It is worth
mentioning that the surface chloride concentration of MK-F2 was at least 58.5% lower with
respect to the maximum free chloride concentrations reported in (Chand et al. 2020; Dobias
et al. 2016; Lu et al. 2015; Min et al. 2009; Wang et al. 2014) for normal-strength materials
exposed to less aggressive environments.

68
Chapter 3

Figure 3.15. Chloride penetration test results of (a) R-F1, (b) R-F2, (c) MK-F2, (d) S-F2, and apparent diffusivities and penetration resistances of (e) R-F1, R-
F2, MK-F2, S-F2

69
Chapter 3

Figure 3.16. MK-F2 specimen after 216 d of exposure to cyclic drying and wetting with 10%
NaCl solution

The steel fibres of all the UHP-SFRC specimens were observed to be corroded only at the
surface. The corrosion was not observed to progress inside the specimens. Such observation
can be attributed to the use of brass coated steel fibres in this study as well as to the very low
chloride concentration at any depth of each of the UHP-SFRC cubes. Figure 3.16 shows the
exposed surface and split section of MK-F2 after its exposure to cyclic drying and wetting.

The apparent diffusivity (Da) values of R-F1, R-F2, MK-F2, and S-F2, calculated based on
their chloride concentration profiles shown in Figures 3.15(a), 3.15(b), 3.15(c), and 3.15(d),
respectively, are presented in Figure 3.15(e). Higher Da values of R-F1 and R-F2 than that
of MK-F2 or S-F2 were expected owing to the low alumina contents in the VLAC – silica
fume binder systems of the former. Although MK-F2 had a higher chloride concentration
near the surface than R-F1, R-F2 and S-F2, the chloride concentration of MK-F2 decreased
more rapidly along the direction of chloride ingress. As a result, Da was found to be the
lowest for MK-F2. The VLAC – silica fume – metakaolin binder system in MK-F2 had
higher combined Al2O3 content, as well as comparable or higher combined CaO content,
than those in the binder systems present in other UHP-SFRCs (Table 3.2). Such chemical
composition of VLAC – silica fume – metakaolin binder system possibly enhanced the
chloride binding capacity of MK-F2 through the formation of Friedel’s salt and reduced the
amount of free chloride at a greater extent along the direction of ingress (Seleem et al. 2010).
The Da of S-F2 was lower than R-F2 but higher than MK-F2. This is possibly because the
binder system of S-F2 had higher alumina content than that of R-F2 but lower than that of
MK-F2. The low initial and secondary sorptivities of S-F2 compared to the other UHP-
SFRCs, as presented in Figure 3.14, perhaps also played a beneficial role to reduce the rate
of chloride transport into the composite. Based on the classification suggested by Gjørv
(1996), the chloride penetration resistances of R-F1 and R-F2 can be categorized as ‘very
high’, and the resistances of MK-F2 and S-F2 can be categorized as ‘extremely high’ [Figure
3.15(e)]

70
Chapter 3

3.3.4 Microstructure analyses

Figure 3.17 shows the SEM images of polished thin sections of R-F2, MK-F2, and S-F2,
after 28 d of curing. More noticeable microcracks were observed in the matrix of R-F2 than
in MK-F2 or S-F2. This is indicative of more pronounced shrinkage in the UHP-SFRC with
cement – silica fume binder system than that in the UHP-SFRC with cement – silica fume –
metakaolin or cement – silica fume – GGBS binder system (Ahmed et al. 2021b; Akcay and
Tasdemir 2018; Güneyisi et al. 2010; Staquet and Espion 2004). Higher extent of
microcracking in R-F2 is perhaps one of the factors which caused the composite to exhibit
higher initial sorptivity than those exhibited by MK-F2 and S-F2 [Figure 3.14(b)]. Owing to
the low w/b (= 0.193) of the UHP-SFRC mixes, interfaces between fibre and matrix in the
UHP-SFRCs were in general found to be dense. Nevertheless, S-F2 seemingly had denser
fibre – matrix interfaces than the others. This can be attributed to the lower slump flow of
fresh S-F2 than those of R-F2 and MK-F2 which restricted local bleeding around the fibres
in S-F2 (Abdallah et al. 2018), and to the pozzolanic reactivity of GGBS which helped form
secondary C-S-H around the fibres (Zhou et al. 2010). Such features of S-F2 most likely
helped it attain better performance in flexure than the others (Figure 3.13).

Figure 3.18 shows the XRD patterns of R-F2, MK-F2, and S-F2 powders collected from a
depth of 7.5 mm of respective UHP-SFRC cubes, after being subjected to cyclic drying and
wetting with 10% NaCl solution for 216 d (total curing time was 28 + 216 + 2 = 246 d). The
Friedel’s salt peak at around 23° was found to be the most pronounced for MK-F2. Higher
Friedel’s salt formation in MK-F2 most likely enabled it to attain lower Da than the other
UHP-SFRCs [Figure 3.15(e)]. The intensity of the Friedel’s salt peak of R-F2 was
substantially low, which was expected because of the lower Al2O3 content in the VLAC –
silica fume binder system than those in the binder systems of MK-F2 and S-F2 (Table 3.2).
The ettringite peak of R-F2 at around 9° was not notable. The intensity of the ettringite peak
was also very low for MK-F2. However, slightly more pronounced ettringite peak was
observed in the XRD pattern of S-F2. This is possibly because the GGBS had higher gypsum
content than that in the VLAC (Figure 3.1). A sulphate source like gypsum has a favourable
effect on the stability of ettringite, but it limits the formation of Friedel’s salt (Luo et al. 2003;
De Weerdt et al. 2014). Hence, lower intensity of the Friedel’s salt peak in the XRD pattern
of S-F2 than in that of MK-F2 can be attributed to the higher gypsum content in the former
as well as to the lower Al2O3 content in its binder system. The use of low-sulphate GGBS
can be more beneficial for improving the chloride binding capacity of a composite. A study
can be carried out in future to investigate the chloride penetration resistance of VLAC based
UHP-SFRC incorporating low-sulphate GGBS.

71
Chapter 3

Figure 3.17. SEM images of (a) R-F2, (b) MK-F2, and (c) S-F2

72
Chapter 3

Figure 3.18. XRD patterns of R-F2, MK-F2, and S-F2 powders, collected from a depth of 7.5
mm, after 216 d of exposure to cyclic drying and wetting with 10% NaCl solution [A = alite, B
= belite, E = ettringite, Fs = Friedel’s salt, P = portlandite, Q = quartz]

It was interesting to notice that the CH peak at around 18° was slightly higher for MK-F2
than for R-F2 or S-F2. This indicates that at later ages (between 28 d and 246 d), pore
refinement through the formation of secondary C-S-H slowed down in MK-F2 compared to
R-F2 or S-F2. Similar results were reported by Curcio et al. (1998). They observed that the
cementitious mixes prepared with four different types of metakaolin had higher CH
concentrations at 90 d than that of the mix prepared with silica fume, although some of the
mixes with metakaolin had lower CH concentrations at an age of less than 28 d. Lower
formation of secondary C-S-H in MK-F2 at later ages was perhaps a factor that caused
higher chloride concentration near the surface of the composite (Figure 3.15).

3.4 Conclusions

The main conclusions that can be drawn from the results presented in the preceding section
are as follows:

1. Self-compacting UHP-SFRC mixes having relative slump flow values of more than 5
(without jolting during flow test) and 28-d compressive strengths of more than 165 MPa
(without employing special curing) can be prepared with VLAC and a w/b ratio of 0.193.
2. The inclusion of straight steel fibres in VLAC based UHPC mix up to 2% by volume
marginally changes the flowability. The relative slump flow of the UHPC/ SC-UHP-
SFRC mixes with VLAC – silica fume binder system ranges from 5.59 to 5.92. The
relative slump flow values of the mixes with VLAC – silica fume – metakaolin and
VLAC – silica fume – GGBS binder systems lie in the ranges of 6.41–6.54 and 5.18–
5.45, respectively. 40% substitution of the silica fume in VLAC based SC-UHP-SFRC

73
Chapter 3

with metakaolin increases the slump flow diameter of the mix by 6%. The slump flow
of SC-UHP-SFRC marginally reduces by 1.1% when 30% of the VLAC is replaced with
GGBS.
3. Replacement of 40% of the silica fume with metakaolin increases the compressive
strength of SC-UHP-SFRC. Increment in the compressive strength is more pronounced
at 7 d than at 28 d. The 7-d and 28-d strengths of SC-UHP-SFRC with metakaolin are
7.8% and 3.4% higher with respect to those of the SC-UHP-SFRC without metakaolin.
The compressive strength of SC-UHP-SFRC reduces when 30% of the VLAC is replaced
with GGBS. The reduction is less pronounced at 28 d than at 7 d. The 7-d and 28-d
strengths of SC-UHP-SFRC incorporating GGBS reduce by 10.2% and 4.9%,
respectively, with respect to the SC-UHP-SFRC with VLAC – silica fume binder system.
4. Increment in the post-cracking flexural strength of VLAC based SC-UHP-SFRC with
respect to its first cracking strength is noticeable when a straight steel fibre content of 2%
is used (i.e., a deflection-hardening branch in the load-deflection curve is detectable at
2% fibre content). The 28-d flexural strengths of SC-UHP-SFRCs with metakaolin and
GGBS are respectively 35.7% and 65.6% higher, with respect to the 28-d flexural
strength of SC-UHP-SFRC with VLAC – silica fume binder system. The deflection
capacity at the point of maximum load is 84.6% higher for SC-UHP-SFRC incorporating
metakaolin and 136.8% higher for SC-UHP-SFRC incorporating GGBS, with respect to
the SC-UHP-SFRC without metakaolin and GGBS.
5. The 28-d water absorption of VLAC based SC-UHP-SFRC with 2% fibre is 40.7% lower
with respect to that of the SC-UHP-SFRC with 1% fibre. The initial and secondary
capillary sorptivities of VLAC based SC-UHP-SFRC also reduce when a fibre content
of 2% is used instead of 1%. The 28-d water absorption of SC-UHP-SFRC incorporating
metakaolin is 8.6% higher and the water absorption of SC-UHP-SFRC with GGBS is
25.7% lower with respect to the SC-UHP-SFRC with VLAC – silica fume binder system.
The initial and secondary sorptivities of SC-UHP-SFRC with GGBS are lower than
those of the SC-UHP-SFRC without GGBS. The inclusion of metakaolin slightly
reduces the initial sorptivity of the SC-UHP-SFRC with VLAC – silica fume binder
system. The secondary sorptivity, however, increases when metakaolin is incorporated.
6. The surface chloride concentration of the SC-UHP-SFRC mix with metakaolin,
subjected to cyclic drying and wetting with 10% NaCl solution for 216 d, can be higher
than the mix with VLAC – silica fume binder system. Nevertheless, the chloride
diffusivity of the mix with metakaolin is substantially low compared with the mix with
VLAC – silica fume binder system. Both the surface chloride concentration and the
chloride diffusivity can be lower for SC-UHP-SFRC with GGBS than for SC-UHP-

74
Chapter 3

SFRC with VLAC – silica fume binder system. Based on the diffusivity results obtained
in this study, the resistance to chloride penetration of SC-UHP-SFRC with VLAC –
silica fume binder system can be categorized as ‘very high’, and the resistances of SC-
UHP-SFRCs with metakaolin and GGBS can be categorized as ‘extremely high’.

Acknowledgements

The authors express gratitude to Adelaide Brighton Cement, Australia for providing the
very-low-C3A cement. Special thanks are given to Sika, Australia for providing the HRWR
and the defoaming agent used in this study. The authors thank Mr. Andrew van de Ven for
his contribution during conducting the chloride penetration test. The authors acknowledge
the facilities given by the Centre for Microscopy, Characterization & Analysis (CMCA),
The University of Western Australia (UWA) and School of Molecular Sciences, UWA for
conducting the SEM and XRD analyses, respectively. Thanks are also given to Mr. Jianfeng
Chen, Ms. May Khant Maw, Mr. Hanqing Zhang, and Mr. Congnan Zhang for their
contributions during the laboratory work. The research was undertaken while the first author
was in receipt of a University Postgraduate Award and the Australian Government Research
Training Program Scholarship at UWA.

References
Abbas, S., Soliman, A. M., and Nehdi, M. L. (2015). “Exploring mechanical and durability
properties of ultra-high performance concrete incorporating various steel fiber lengths
and dosages.” Construction and Building Materials, 75, 429–441.
Abdallah, S., Fan, M., and Rees, D. W. A. (2018). “Bonding Mechanisms and Strength of
Steel Fiber–Reinforced Cementitious Composites: Overview.” Journal of Materials in
Civil Engineering, 30(3), 04018001.
Aghdasi, P., and Ostertag, C. P. (2018). “Green ultra-high performance fiber-reinforced
concrete (G-UHP-FRC).” Construction and Building Materials, 190, 246–254.
Ahmed, T., Elchalakani, M., Karrech, A., Mohamed Ali, M. S., and Guo, L. (2021a).
“Development of ECO-UHPC with very-low-C3A cement and ground granulated blast-
furnace slag.” Construction and Building Materials, 284, 122787.
Ahmed, T., Elchalakani, M., Karrech, A., Nawaz, W., and Liu, H. (2021b). “Shrinkage of
UHPC Incorporating Very-Low-C3A Cement and Ground Granulated Blast-Furnace
Slag.” 30th Biennial National Conference of the Concrete Institute of Australia, Perth, WA.
Akcay, B., and Tasdemir, M. A. (2018). “Performance evaluation of silica fume and
metakaolin with identical finenesses in self compacting and fiber reinforced concretes.”
Construction and Building Materials, 185, 436–444.
AS/NZS 3582.2. (2016). Supplementary cementitious materials for use with portland and blended
cement - Slag—Ground granulated blastfurnace.
AS/NZS 3582.3. (2016). Supplementary cementitious materials for use with portland and blended
cement - amorphous silica.

75
Chapter 3

AS 1012.11. (2000). Methods of testing concrete: determination of the modulus of rupture.


AS 1012.3.5. (2015). Methods of testing concrete: determination of properties related to the
consistency of concrete - slump flow, T500 and J-ring test.
ASTM C150 / C150M-16. (2016). Standard specification for portland cement. West
Conshohocken, PA.
ASTM C1556-11a. (2016). Standard Test Method for Determining the Apparent Chloride Diffusion
Coefficient of Cementitious Mixtures by Bulk Diffusion. West Conshohocken, PA.
ASTM C1585-04. (2004). Standard Test Method for Measurement of Rate of Absorption of Water
by Hydraulic Cement Concretes. West Conshohocken, PA.
ASTMC109 / C109M-16a. (2016). Standard test method for compressive strength of hydraulic
cement mortars (using 2-in. or [50-mm] cube specimens). West Conshohocken, PA.
Australian Government Bureau of Meteorology. (2021). Climate Data Online.
Balonis, M., Lothenbach, B., Le Saout, G., and Glasser, F. P. (2010). “Impact of chloride
on the mineralogy of hydrated Portland cement systems.” Cement and Concrete Research,
40(7), 1009–1022.
Bentz, D. P., Ferraris, C. F., Galler, M. A., Hansen, A. S., and Guynn, J. M. (2012).
“Influence of particle size distributions on yield stress and viscosity of cement-fly ash
pastes.” Cement and Concrete Research, 42(2), 404–409.
Cadoni, E., Forni, D., Bonnet, E., and Dobrusky, S. (2019). “Experimental study on direct
tensile behaviour of UHPFRC under high strain-rates.” Construction and Building
Materials, 218, 667–680.
Chand, V. V. S., Rao, B. K., and Rao, C. H. (2020). “Investigation on chloride penetration
in concrete mixes of different cement replacement percentages with fly ash and silica
fume.” Materials Today: Proceedings, 33, 820–827.
Choi, W. C., Jung, K. Y., Jang, S. J., and Yun, H. Do. (2019). “The influence of steel fiber
tensile strengths and aspect ratios on the fracture properties of high-strength concrete.”
Materials, 12(13).
Chu, S. H., and Kwan, A. K. H. (2019). “Mixture design of self-levelling ultra-high
performance FRC.” Construction and Building Materials, 228.
Curcio, F., DeAngelis, B. A., and Pagliolico, S. (1998). “Metakaolin as a pozzolanic
microfiller for high-performance mortars.” Cement and Concrete Research, 28(6), 803–809.
Ding, J. T., and Li, Z. (2002). “Effects of metakaolin and silica fume on properties of
concrete.” ACI Materials Journal, 99(4), 393–398.
Dobias, D., Pernicova, R., and Mandlik, T. (2016). “Water transport properties and depth
of chloride penetration in ultra high performance concrete.” Key Engineering Materials,
711, 137–142.
Easterbrook, D., Geng, J., Li, L.-Y., and Liu, Q.-F. (2015). “Effect of carbonation on release
of bound chlorides in chloride-contaminated concrete.” Magazine of Concrete Research, 1–
11.
EFNARC. (2002). Specification and Guidelines for Self-Compacting Concrete. Hampshire, GU.
Elchalakani, M., Basarir, H., and Karrech, A. (2017). “Green concrete with high-volume fly
ash and slag with recycled aggregate and recycled water to build future sustainable
cities.” Journal of Materials in Civil Engineering, 29(2), 04016219.
Ganesh, P., and Murthy, A. R. (2019). “Tensile behaviour and durability aspects of
sustainable ultra-high performance concrete incorporated with GGBS as cementitious
material.” Construction and Building Materials, 197, 667–680.

76
Chapter 3

Ghasemzadeh, F., and Pour-Ghaz, M. (2015). “Effect of Damage on Moisture Transport in


Concrete.” Journal of Materials in Civil Engineering, 27(9), 04014242.
Gjørv, O. E. (1996). “Performance and Serviceability of Concrete Structures in the Marine
Environment.” Odd E. Gjørv Symposium on Concrete for Marine Structures, P. K. Mehta,
ed., CANMET/ACI, Ottawa, ON, 259–279.
Güneyisi, E., Gesolu, M., and Özbay, E. (2010). “Strength and drying shrinkage properties
of self-compacting concretes incorporating multi-system blended mineral admixtures.”
Construction and Building Materials, 24(10), 1878–1887.
Guo, Y., Zhang, T., Tian, W., Wei, J., and Yu, Q. (2019). “Physically and chemically
bound chlorides in hydrated cement pastes: a comparison study of the effects of silica
fume and metakaolin.” Journal of Materials Science, 54(3), 2152–2169.
Hung, C. C., Chen, Y. T., and Yen, C. H. (2020). “Workability, fiber distribution, and
mechanical properties of UHPC with hooked end steel macro-fibers.” Construction and
Building Materials, 260.
IS 3495 (Part 2). (1992). Methods of tests of burnt clay building bricks, Part 2: determination of water
absorption. New Delhi.
Kadri, E. H., Kenai, S., Ezziane, K., Siddique, R., and De Schutter, G. (2011). “Influence
of metakaolin and silica fume on the heat of hydration and compressive strength
development of mortar.” Applied Clay Science, 53(4), 704–708.
Khaloo, A. R., and Kim, N. (1996). “Mechanical Properties of Normal to High-Strength
Steel Fiber-Reinforced Concrete.” Cement, Concrete and Aggregates, 18(2), 92–97.
Kim, D. J., Park, S. H., Ryu, G. S., and Koh, K. T. (2011). “Comparative flexural behavior
of hybrid ultra high performance fiber reinforced concrete with different macro fibers.”
Construction and Building Materials, 25(11), 4144–4155.
Koushkbaghi, M., Kazemi, M. J., Mosavi, H., and Mohseni, E. (2019). “Acid resistance
and durability properties of steel fiber-reinforced concrete incorporating rice husk ash
and recycled aggregate.” Construction and Building Materials, 202, 266–275.
Liu, H., Zhang, Y., Liu, J., Kong, S., and Feng, Z. (2021). “Chloride Binding Capacity of
Portland Cement System with Different Content of Tricalcium Aluminate.” IOP
Conference Series: Earth and Environmental Science, 719(2).
Lu, C., Gao, Y., Cui, Z., and Liu, R. (2015). “Experimental Analysis of Chloride
Penetration into Concrete Subjected to Drying–Wetting Cycles.” Journal of Materials in
Civil Engineering, 27(12), 04015036.
Luo, R., Cai, Y., Wang, C., and Huang, X. (2003). “Study of chloride binding and diffusion
in GGBS concrete.” Cement and Concrete Research, 33(1), 1–7.
Meng, W., and Khayat, K. H. (2018). “Effect of Hybrid Fibers on Fresh Properties,
Mechanical Properties, and Autogenous Shrinkage of Cost-Effective UHPC.” Journal of
Materials in Civil Engineering, 30(4), 04018030.
Min, B., Ditao, N., and Xiong, W. (2009). “Experiment study on the chloride penetration
of steel fibre reinforced concrete.” Advanced Materials Research, 79–82, 1771–1774.
Mohammadi, J., and South, W. (2016). “Measuring chromium in general purpose (GP)
cement.” Journal of Testing and Evaluation, 45(4), 1362–1377.
Neville, A. M. (2011). Properties of concrete. Pearson, London, UK.
Ngo, T. T., Park, J. K., Pyo, S., and Kim, D. J. (2017). “Shear resistance of ultra-high-
performance fiber-reinforced concrete.” Construction and Building Materials, 151, 246–
257.
Oh, B. H., Jang, S. Y., and Shin, Y. S. (2003). “Experimental investigation of the threshold

77
Chapter 3

chloride concentration for corrosion initiation in reinforced concrete structures.”


Magazine of Concrete Research, 55(2), 117–124.
Oner, A., and Akyuz, S. (2007). “An experimental study on optimum usage of GGBS for
the compressive strength of concrete.” Cement and Concrete Composites, 29(6), 505–514.
Ren, G. M., Wu, H., Fang, Q., and Liu, J. Z. (2018). “Effects of steel fiber content and type
on static mechanical properties of UHPCC.” Construction and Building Materials, 163,
826–839.
Sakthivel, P. B., and Vijay Aravind, S. (2020). “Flexural strength and toughness of steel fiber
reinforced concrete beams.” Asian Journal of Civil Engineering, 21(8), 1309–1330.
Scheydt, J., and Müller, H. (2012). “Microstructure of ultra high performance concrete
(UHPC) and its impact on durability.” 3rd International Symposium on on Ultra High
Performance Concrete, 349–356.
Seleem, H. E. D. H., Rashad, A. M., and El-Sabbagh, B. A. (2010). “Durability and strength
evaluation of high-performance concrete in marine structures.” Construction and Building
Materials, 24(6), 878–884.
Song, P. S., and Hwang, S. (2004). “Mechanical properties of high-strength steel fiber-
reinforced concrete.” Construction and Building Materials, 18(9), 669–673.
Song, Q., Yu, R., Shui, Z., Wang, X., Rao, S., and Lin, Z. (2018a). “Optimization of fibre
orientation and distribution for a sustainable Ultra-High Performance Fibre Reinforced
Concrete (UHPFRC): Experiments and mechanism analysis.” Construction and Building
Materials, 169, 8–19.
Song, Q., Yu, R., Shui, Z., Wang, Y., Rao, S., Wu, S., and He, Y. (2019). “Physical and
chemical coupling effect of metakaolin induced chloride trapping capacity variation for
Ultra High Performance Fibre Reinforced Concrete (UHPFRC).” Construction and
Building Materials, 223, 765–774.
Song, Q., Yu, R., Wang, X., Rao, S., and Shui, Z. (2018b). “A novel Self-Compacting Ultra-
High Performance Fibre Reinforced Concrete (SCUHPFRC) derived from compounded
high-active powders.” Construction and Building Materials, 158, 883–893.
Stanish, K. D., Hooton, R. D., and Thomas, M. D. A. (2001). Testing the Chloride Penetration
Resistance of Concrete : A Literature Review. Prediction of Chloride Penetration in Concrete,
McLean, VA.
Staquet, S., and Espion, B. (2004). “Early-age autogenous shrinkage of UHPC incorporating
very fine fly ash or metakaolin in replacement of silica fume.” International Symposium
on UHPC, M. Schmidt, E. Fehling, and C. Geisenhanslüke, eds., Kassel university press
GmbH, Kassel, 587–599.
Ting, L., Qiang, W., and Shiyu, Z. (2019). “Effects of ultra-fine ground granulated blast-
furnace slag on initial setting time, fluidity and rheological properties of cement pastes.”
Powder Technology, 345, 54–63.
Tuaum, A., Shitote, S., and Oyawa, W. (2018). “Experimental study of self-compacting
mortar incorporating recycled glass aggregate.” Buildings, 8(2).
Wang, G. M., Kong, Y., Shui, Z. H., Li, Q., and Han, J. L. (2014). “Experimental
investigation on chloride diffusion and binding in concrete containing metakaolin.”
Corrosion Engineering Science and Technology, 49(4), 282–286.
De Weerdt, K., Orsáková, D., and Geiker, M. R. (2014). “The impact of sulphate and
magnesium on chloride binding in Portland cement paste.” Cement and Concrete Research,
65, 30–40.
Wu, Z., Shi, C., and He, W. (2017). “Comparative study on flexural properties of ultra-high
performance concrete with supplementary cementitious materials under different curing

78
Chapter 3

regimes.” Construction and Building Materials, 136, 307–313.


Wu, Z., Shi, C., He, W., and Wu, L. (2016). “Effects of steel fiber content and shape on
mechanical properties of ultra high performance concrete.” Construction and Building
Materials, 103, 8–14.
Yang, S. L., Millard, S. G., Soutsos, M. N., Barnett, S. J., and Le, T. T. (2009). “Influence
of aggregate and curing regime on the mechanical properties of ultra-high performance
fibre reinforced concrete (UHPFRC).” Construction and Building Materials, 23(6), 2291–
2298.
Yazıcı, H., Yardimci, M. Y., Yiǧiter, H., Aydin, S., and Türkel, S. (2010). “Mechanical
properties of reactive powder concrete containing high volumes of ground granulated
blast furnace slag.” Cement and Concrete Composites, 32(8), 639–648.
Yoo, D. Y., Shin, H. O., Yang, J. M., and Yoon, Y. S. (2014). “Material and bond properties
of ultra high performance fiber reinforced concrete with micro steel fibers.” Composites
Part B: Engineering, 58, 122–133.
Yu, R., Spiesz, P., and Brouwers, H. J. H. (2015). “Development of an eco-friendly Ultra-
High Performance Concrete (UHPC) with efficient cement and mineral admixtures
uses.” Cement and Concrete Composites, 55, 383–394.
Zhou, J., Qian, S., Beltran, M. G. S., Ye, G., Van Breugel, K., and Li, V. C. (2010).
“Development of engineered cementitious composites with limestone powder and blast
furnace slag.” Materials and Structures/Materiaux et Constructions, 43(6), 803–814.

79
CHAPTER
4
ECO-UHPC with high-volume Class-F fly ash:
new insight into mechanical and durability
properties

Tanvir Ahmeda, Mohamed Elchalakania, Ali Karrecha, Minhao Donga, M. S. Mohamed Alib,
Hua Yangc,*
a
Department of Civil, Environmental and Mining Engineering, Faculty of Engineering, Computing and Mathematical
Sciences, The University of Western Australia, 35 Stirling Highway, Crawley, WA 6009, Australia
b
School of Civil, Environmental and Mining Engineering, The University of Adelaide, Adelaide, SA 5005, Australia
c
Key Laboratory of Structures Dynamic Behaviour and Control (Harbin Institute of Technology), Ministry of Education,
Heilongjiang, Harbin 150090, China; Professor, School of Civil Engineering, Harbin Institute of Technology, Heilongjiang,
Harbin 150090, China

*Corresponding author: yanghua@hit.edu.cn

Abstract

Ultra-high performance concrete (UHPC), despite its superior mechanical and durability
performances, has a high CO2 footprint owing to its high Portland cement content. This
drawback can be offset to a notable extent if a high volume of supplementary cementitious
materials can be utilized to produce UHPC while maintaining mechanical and durability
properties that are comparable to those of conventional UHPC. In this study, the effects of
high-volume cement replacement by Class-F fly ash (up to 70% by mass) on the mechanical,
durability, and microstructure properties of UHPC are investigated, with the aim of
encouraging moderate- to low-volume cement use in UHPC (from 600 down to 300 kg/m3).
Environmentally friendly-UHPC (ECO-UHPC) mixes, which have CO2 footprint intensities
of less than 5 kg/m3/MPa, have been synthesised with different replacements of cement by
Class-F fly ash. Results suggest that concretes with ultra-high strength (>150 MPa) and a
strength greater than 100 MPa can be prepared with up to 40% and 70% replacements,
respectively, without employing any special curing or fibres. A model has been developed
to predict the compressive strength of fly ash based UHPC from the contents and chemical
compositions of its binders. UHPCs made with up to 60% cement replacement by Class-F

Journal of Materials in Civil Engineering, 33(7), 0003726


(https://doi.org/10.1061/(asce)mt.1943-5533.0003726)
Chapter 4

fly ash exhibit durability properties comparable to that of UHPC without fly ash, in terms
of water absorption, initial rate of water absorption, and corrosion risk. The depth of
carbonation remains below the detection limit of 0.5 mm, up to 70% replacement.

Keywords: CO2 emission, half-cell potential, hydration modulus, reactive powder concrete,
ultra-high performance concrete

4.1 Introduction

Ultra-high performance concrete (UHPC) is a relatively new generation concrete.


Conventionally, Portland cement, silica fume, and very fine quartz aggregate (fine quartz
sand and powder) are used for its production. High-range water reducer is used to obtain
workable mix with a very low water to binder ratio (≤0.2). Since a large amount of binder
or reactive powder is used in UHPC production (e.g., 800–1100 kg/m3 of cement, 100–300
kg/m3 of silica fume), and coarse aggregate is normally avoided, the concrete is sometimes
referred to as reactive powder concrete. Non-fibre reinforced UHPC is brittle at material
level, but reinforced concrete (RC) members constructed with UHPC may exhibit ample
ductility (Yoo and Yoon 2015). To enhance the ductility of UHPC at the material level, steel
fibres are usually incorporated into it (e.g., commercially available UHPCs typically include
fibres). In this study, “UHPC” was used for non-fibre reinforced systems. For fibre-
reinforced system, the term ultra-high performance fibre-reinforced concrete (UHPFRC)
was used (Naaman 2011). The main feature of UHPC is its ultra-high strength in
compression. However, there is a lack of consensus on the definition of the term ultra-high
strength. A number of researchers consider ultra-high strength to be greater than or equal to
120 MPa (Huang et al. 2019; Wu et al. 2016b), while several researchers consider it to be
greater than 150 MPa (Courtial et al. 2013; Naaman 2011). The latter definition was adopted
in this study. The flexural strength of conventional UHPC can be greater than 15 MPa
(Soliman and Tagnit-Hamou 2017). UHPC also exhibits outstanding durability. The water
absorption of UHPC was found to be less than 1.2%, which is only 20%–35% of that of
normal-strength concrete (NSC) (Alsayed and Amjad 1996; Ghafari et al. 2012; Scheydt and
Müller 2012). Jaafar et al. (2018) found the corrosion risk of steel rebars embedded in UHPC
specimens to be substantially lower than those of rebars embedded in NSC and high-strength
concrete (HSC) specimens. At 60 days, the half-cell potential (HCP) of rebars embedded in
UHPC specimens was approximately 150 mV and 140 mV less negative than the HCPs of
rebars in NSC and HSC specimens, respectively. Conventional UHPC was observed to
exhibit zero or negligible (<0.5 mm) depth of carbonation after being exposed to accelerated
carbonation environment (Long et al. 2005; Scheydt and Müller 2012).

81
Chapter 4

Though UHPC shows extraordinary mechanical and durability properties, its CO2 footprint
is usually very high because of its large cement content. Several researchers sought to reduce
the cement content of UHPC by incorporating Class-F fly ash (a by-product of coal power
plant). Table 4.1 summarizes the mechanical properties reported in the literature for Class-
F fly ash based UHPCs cured in standard curing condition (28 days of curing in water or in
fog room at a relative humidity 95% ± 5%, at a temperature of 22 °C ± 3 °C). Very few of
these studies systematically investigated the influence of substituting Portland cement by
different contents of Class-F fly ash on the properties of UHPC (Chen et al. 2018; Peng et
al. 2010). Mainly compressive strength and in some cases flexural strength behaviours were
investigated in these studies up to 47% replacement of cement by Class-F fly ash. For 30%–
31% replacement, a maximum compressive strength of 135.7 MPa (Mueller et al. 2016) and
a maximum flexural strength of 18.4 MPa (Chen et al. 2018) were reported by previous
studies. For 47% replacement, Peng et al. (2010) reported a compressive strength of 88.3
MPa. Detailed study addressing the influence of high-volume Portland cement substitution
by Class-F fly ash, as high as 70%, on the mechanical properties of UHPC is not available.
In addition, no study in the literature explored the influence of high-volume Class-F fly ash
incorporation on the durability of UHPC, for example, water absorption, initial rate of water
absorption, corrosion risk of embedded rebar, and carbonation resistance.

Table 4.1. Review of mechanical properties of standard-cured UHPCs and UHPC pastes with
cement – silica fume – Class-F fly ash based binder system

Reference UHPC/ UHPC w/ba Fly ash 28-day mechanical properties


paste contenta,b (MPa)
fc fr fsp Ec
Chen et al. (2018) UHPC 0.2 0% 114 16.3 - -
UHPC 0.2 10% 115.8 16.7 - -
UHPC 0.2 20% 126 20 - -
UHPC 0.2 30% 121.3 18.4 - -
Liu et al. (2018) UHPC 0.18 21% 108.5 10.4 - -
Song et al. (2018) UHPC 0.2 18% 105 9.4 - -
Ferdosian et al. UHPC paste 0.18 0% 152 - - -
(2017) UHPC paste 0.18 17% 150 - - -
UHPC paste 0.18 26% 137 - - -
Mueller et al. UHPC 0.16 25% 147.2 - - 49700
(2016) UHPC 0.17 31% 135.7 - - 49900
Wang et al. (2016) UHPC 0.18 29% 110 - - -
Li et al. (2015) UHPC 0.16 9% 125.6 22.2 - -
UHPC 0.17 9% 120.8 18.5 - -
Peng et al. (2010) UHPC 0.18 0% 113.1 - - -
UHPC 0.18 24% 91.5 - - -
UHPC 0.18 47% 88.3 - - -
w/b = water to binder ratio, fc = compressive strength, fr = flexural strength, fsp = splitting tensile strength,
Ec = elastic modulus
a
By mass
b
With respect to the total amount of cement and fly ash

82
Chapter 4

Although the incorporation of a high volume of fly ash may reduce somewhat the
mechanical properties of UHPC (Elchalakani et al. 2017), the use of fly ash based UHPC
can be advantageous in several aspects. For instance, the use of fly ash may enhance the
workability of UHPC (Kwan and Li 2013); fly ash incorporation up to a certain extent may
improve particle packing of UHPC and thus result in comparable or better durability
(Chindaprasirt and Rukzon 2008); high-volume substitution of cement by fly ash may reduce
the formation of thermal cracks in large structures (e.g., raft foundation of high-rise buildings
or skyscrapers), especially those constructed in hot and arid areas (Aldred 2010); high-
volume fly ash use in UHPC may reduce the cost and hazards associated with fly ash
disposal; and the CO2 footprint of UHPC can be reduced by fly ash incorporation.

This study carried out a comprehensive experimental investigation to understand the effect
of high-volume replacement of Portland cement by Class-F fly ash on the properties of
UHPC. Maximum replacements, up to which ultra-high strength (i.e., strength ≥ 150 MPa)
and a very high strength of 100 MPa are achievable without employing fibres or special
curing, were aimed to be identified. In addition to compressive strength, mechanical
properties, such as elastic modulus, splitting tensile strength, and flexural strength, and
durability properties such as water absorption, initial rate of absorption, half-cell potential,
and carbonation resistance were investigated for different Class-F fly ash replacements.
Class-F fly ash was specifically chosen because the majority of fly ash produced in Australia
is Class-F (Temuujin et al. 2013). Lack of research investigating the relationships between
the mechanical properties of standard-cured UHPC in general (irrespective of types of binder
present in UHPC) motivated the authors to study the responses of elastic modulus, splitting
tensile strength, and flexural strength of standard-cured UHPC as functions of compressive
strength. Data points from both previous studies and this study were used to develop the
relationships. Microstructure analyses, such as X-ray diffraction (XRD) analysis and
scanning electron microscopy (SEM), were performed for selected environmentally friendly-
UHPC (ECO-UHPC) samples. To assess the eco-friendly nature of the ECO-UHPC mixes
with fly ash, their CO2 footprints were evaluated and compared with that of the ECO-UHPC
without fly ash.

4.2 Experimental program

4.2.1 Materials

Three different types of binders were used to prepare the UHPC mixes investigated in this
study: ASTM C150 / C150M-19a Type I Portland cement (ASTM C150 / C150M-19a
2019), silica fume conforming to AS/NZS 3582.3 (2016), and ASTM C618-19 Class-F fly

83
Chapter 4

ash (ASTM C618-19 2019) [AS/NZS 3582.1 Grade 1 fly ash (AS/NZS 3582.1 2016)]. The
silica fume used in this study was moderately densified (with bulk density 360–400 kg/m3).
The XRD patterns of the raw binders are shown in Figure 4.1. Crystalline phases present in
the as-received cement were mainly alite (Ca3SiO5 or, C3S), belite (Ca2SiO4 or, C2S),
aluminate (Ca3Al2O6 or, C3A) ferrite [Ca4(Al,Fe)2O7 or, C4AF], and gypsum (CaSO4.2H2O).
The broad hump in the 2θ range of 18°–28° of the as-received Class-F fly ash indicates that
the fly ash was mostly in an amorphous state. In addition, some crystalline peaks
corresponding to quartz (SiO2), mullite (Al6Si2O13), and hatrurite (Ca3SiO5) were identified
in the XRD pattern of raw fly ash. Few crystalline peaks corresponding to quartz and silicon
carbide (SiC) were observed in the XRD pattern of silica fume, while the majority of it was
found to be amorphous.

Quartz powder with particle size in the range of 0.5–300 μm (median diameter, d50 = 41.5
µm), and quartz sand with particle size in the range of 60–475 μm (d50 = 237.3 µm) were
used as aggregates. The particle size distributions (PSDs) of the binders and aggregates are
presented in Figure 4.2. Table 4.2 shows the physicochemical properties of all the binders
and aggregates. A polycarboxylate based admixture (with specific gravity 1.09 and 36% solid
content by mass), complying with the requirements of AS1478.1 (2000), was used as high-
range water reducer (HRWR). To reduce the amount of entrapped air, a defoaming agent
(with specific gravity 1 and 5.5% solid content by mass) was also used in all the mixes. Grade
500N deformed reinforcement bars (10 mm diameter), conforming to the requirements of
AS/NZS 4671 (2001), were used to prepare specimens for HCP test.

Figure 4.1. XRD patterns of raw binders (G = gypsum, A = alite, B = belite, F = ferrite, M =
mullite, H = hatrurite, Q = quartz, S = silicon carbide)

84
Chapter 4

Table 4.2. Physicochemical properties of binders and aggregates

Material Oxides (mass %) d50 (µm) Specific gravitya


Al2O3 CaO Fe2O3 K2O MgO MnO Na2O P2O5 SiO2 SO3 TiO2 Loss on ignition
Portland Cement 3.18 64.48 4.34 0.35 1.46 0.12 0.14 0.228 21.4 2.93 0.192 1.15 17.3 3.18
Silica fume 0.13 0.33 0.07 0.54 0.71 0.008 0.29 0.129 94.6 0.07 0.005 3.1 1.7 2.15
Class-F fly ash 23.93 6.98 7.94 1.02 1.27 0.104 0.38 0.535 55.9 0.3 1.312 1.71 9.8 2.38
Quartz powder 0.34 0.41 0.1 0.085 0.027 0.001 0.027 0.003 95.5 0.0 0.013 3.45 41.5 2.65
Sand 0.017 0.002 0.008 0.003 0.003 0.001 0.003 0.0 99.88 0.0 0.025 0.05 237.3 2.65
d50 = median diameter
a
As per manufacturer

Figure 4.2. Particle size distributions of the binders and aggregates

85
Chapter 4

4.2.2 Mix proportions

The mix proportions of the UHPC mixtures studied are summarized in Table 4.3. Mix F0
contained no fly ash. The ratios of Class-F fly ash to Portland cement (f/c) of F20, F40, F60,
and F70 were 20/80, 40/60, 60/40, and 70/30, respectively (by mass). The water to binder
ratio (w/b) of each of the mixes from the F-series UHPCs (F-UHPCs) was 0.164. Two mixes,
W0 and W60, were prepared with an increased w/b ratio of 0.193 in order to assess the effect
of w/b on the mechanical properties of UHPC. The f/c ratio of W0 and W60 were kept
similar to those of F0 and F60, respectively (i.e., 0/100 and 60/40, respectively). The binder
to aggregate ratio (b/a), and the ratio of silica fume content to the combined amount of
Portland cement and fly ash [sf/(c+f)] were kept constant in all the mixes investigated.

Figure 4.3 presents the combined PSDs of the F-UHPC mixes. For comparison, the PSD
based on the modified Andreasen and Andersen (A&A) model, as expressed in Equation
4.1, is also presented in Figure 4.3. The value of q in Equation 4.1 was chosen to be 0.23 as
suggested by previous studies (Liu et al. 2018; Song et al. 2018). If compared with the
modified A&A model, the compactness of UHPC mix increased with the increase of cement
replacement by fly ash.

𝐷 𝑞 − 𝐷𝑚𝑖𝑛 𝑞
𝑝(𝐷) = (4.1)
𝐷𝑚𝑎𝑥 𝑞 − 𝐷𝑚𝑖𝑛 𝑞

where D = particle size (µm), p(D) = cumulative percent finer than size D, Dmin and Dmax =
respectively minimum and maximum particle sizes (µm).

4.2.3 Mixing, specimen preparation, and curing

All the dry components (cement, silica fume, fly ash, quartz powder, and sand) of each mix
were mixed for 5 min at a low speed of 60 rpm in a mixer. Then 100% of the total mixing
water, 100% of the defoamer, and 80% of the superplasticizer were added to the dry mix.
The mixing was continued at low speed for another 3 min. After that, the remaining 20% of
the superplasticizer was added to the mix. When the mix was observed to have acquired a
moist appearance, the mixing speed was increased to 180 rpm. Mixing was continued until
the mix appeared to become flowable. At this stage, the mixer was geared up to 300 rpm
and was continued to rotate for 3 min. The prepared UHPC mix was then taken to a
workbench for slump flow testing and casting into moulds.

86
Chapter 4

Table 4.3. Mix proportions

ECO-UHPCs Constituents (kg/m3) Mix design parameters


Series Mix Cement Silica Fly Quartz Sand Water HRWRa Defoaming f/c w/b b/a sf/(c+f)
fume ash powder agenta
F-series F0 935 251 0 205 824 194 14.7 1.3 0/100 0.164 1.153 0.268
(F-UHPCs) F20 748 251 187 205 824 194 14.7 1.3 20/80 0.164 1.153 0.268
F40 561 251 374 205 824 194 14.7 1.3 40/60 0.164 1.153 0.268
F60 374 251 561 205 824 194 14.7 1.3 60/40 0.164 1.153 0.268
F70 281 251 655 205 824 194 14.7 1.3 70/30 0.164 1.153 0.268
W-series W0 922 248 0 202 812 226 14.5 1.3 0/100 0.193 1.153 0.268
(W-UHPCs) W60 369 248 553 202 812 226 14.5 1.3 60/40 0.193 1.153 0.268
c = cement content, sf = silica fume content, f = fly ash content, w = water content, b = binder content, a = aggregate content

Figure 4.3. Combined particle size distributions of F-UHPC mixes

87
Chapter 4

Figure 4.4. Preparation of a typical half-cell potential test specimen

Each mix was cast into 50 × 50 × 50 mm cube moulds, 50 × 100 mm cylindrical moulds,
and 40 × 40 × 160 mm prisms, to prepare specimens for unconfined compressive strength
(UCS), splitting tensile strength, and four-point bending tests, respectively. Cube moulds
with dimensions of 100 × 100 × 100 mm were used to prepare the durability test specimens.
To prepare specimens for HCP testing, an electrode of activated titanium (Ti) and a steel
rebar of 10 mm in diameter and 90 mm long were embedded inside each 100 × 100 × 100
mm mould (as shown in Figure 4.4) prior to casting the concrete mix. On the rebar, 20 mm
of each end was covered by sealant heat-shrink tape. The freshly moulded UHPC samples
were vibrated on a vibration table for 1 min and then moved to a standard curing room
(temperature 21 °C, 95% relative humidity). The specimens were demoulded after 24 h. The
mechanical test specimens were then kept under water (21 °C) until test time, while the
durability specimens were kept in the curing room (temperature 21 °C, 95% relative
humidity). After 28 days, half of the specimens prepared for measuring carbonation
resistance were moved to a carbonation chamber with a CO2 concentration of 10,000 ppm
(temperature 21 °C, 60% relative humidity) and kept in the chamber for another 28 days.
The curing regimes of the carbonation test specimens are illustrated in Figure 4.5.

4.2.4 Test methods

4.2.4.1 Slump flow test

The slump flow diameters of the UHPC samples were measured in accordance with AS
1012.3.5 (2015), using a custom-made mini slump cone having a height of 116 mm and top

88
Chapter 4

and bottom diameters of 38 mm and 76 mm respectively. The slump flow diameter of each
fresh UHPC was measured when the concrete stopped flowing following removal of the
slump cone. The aspect ratio of the mini cone was similar to that of the standard cone
suggested in AS 1012.3.5 (2015).

4.2.4.2 Mechanical tests

The mechanical tests were performed using a compression testing machine with a capacity
of 600 kN. The UCS tests were performed on 50 × 50 × 50 mm UHPC cubes in accordance
with ASTM C109 / C109M-16a (2016). The specimens prepared from the W-UHPC mixes
were tested for measuring compressive strengths at 28 days, while the F-UHPC specimens
were tested at 3, 7, 14, 28, and 56 days. Strain gauges were attached to the 28-day specimens
from the F-series to measure their vertical strains under compression. The static chord elastic
moduli of the specimens were calculated based on AS 1012.17 (1997). The splitting tensile
strength tests and four-point bending tests were performed at 28 days as per the guidelines
of AS 1012.10 (2000) and AS 1012.11 (2000) using 50 × 100 mm cylindrical specimens and
40 × 40 × 160 mm prisms, respectively. At a particular age, three specimens from a mix
were tested for measuring each of the mechanical properties.

Figure 4.6 shows the setup used for the four-point bending tests. Since the specimens used
for the four-point bending tests were small and concrete is a brittle material, the deflection
of specimen during the test was small. Therefore, bearing friction occurring during the test
was assumed to be negligible.

12000
CO2 concentration (ppm)

10000
Environment B
8000 [elevated CO2
(10000 ppm),
6000 21 °C, 60% RH]

4000 Environment A
[atmospheric CO2
2000 (300 ppm),
21 °C, 95% RH]
0
0 7 14 21 28 35 42 49 56
Curing time (days)

Figure 4.5. Curing environments of carbonation test specimens

89
Chapter 4

Figure 4.6. Four-point bending test setup

4.2.4.3 Durability tests

Water absorption (WA) and initial rate of absorption (IRA) tests were performed using 100
× 100 × 100 mm plain cubes prepared from the F-UHPC mixes, as per the guidelines of IS
3495 (Part 2) (1992) and AS/NZS 4456.17 (2003), respectively. Equations 4.2 and 4.3 were
used to calculate respectively the WA and the IRA of each specimen. Three specimens from
each F-UHPC mix were tested for both WA and IRA.

100(𝑚2 − 𝑚1 )
𝑊𝐴 = (%) (4.2)
𝑚1

1000(𝑚2,60 − 𝑚1 )
𝐼𝑅𝐴 = (kg/m2/min) (4.3)
𝐴

where, m1 = mass of oven-dried specimen (g), m2 = mass of specimen after being immersed
in water for 24 h (g), m2, 60 = mass of specimen after submergence of bottom 3 mm under
water for 60 s (g) (Figure 4.7), and A = area of the bottom surface of specimen (mm2). The
excess water on the surface of each specimen was wiped off using damp cloth before
measuring m2 or m2, 60.

90
Chapter 4

100 × 100 × 100 mm


UHPC cube

Submersion of
Metal bottom 3 mm for 60 s
rods
3 mm

Figure 4.7. Initial rate of absorption test setup

Voltmeter
Saturated
calomel
half-cell

Ti electrode

Rebar

Heat-shrink tape

100 × 100 × 100 mm


UHPC cube

(Elevation view)

Figure 4.8. Half-cell potential test setup

HCP tests of the UHPC specimens were conducted using a slightly modified version of the
approach specified in ASTM C876-15 (2015). A schematic illustration of the HCP test setup
used in this study is presented in Figure 4.8. The HCP of each specimen was evaluated based
on the potential difference between the embedded rebar and a reference saturated calomel
(SC) half-cell placed on top of the specimen. As suggested by ASTM C876-15 (2015), the

91
Chapter 4

HCP values obtained using SC half-cell were converted to copper-copper sulphate


[Cu/Cu(SO4)] half-cell equivalent HCP values according to the guidelines mentioned in
(Elchalakani et al. 2018; Meek et al. 2018). The HCP of each F-UHPC, at a particular age,
was calculated based on the average of the HCPs of three representative specimens. ASTM
C876-15 (2015) does not include the use of a Ti electrode in its procedure. However, Ti
electrodes were incorporated into the HCP specimens of this study to assess the accuracy of
the measured HCP readings. The accuracy of the HCP value of each specimen was cross-
checked by Equation 4.4.

Erebar/SC = ETi/rebar + ETi/SC (mV) (4.4)

where Erebar/SC, ETi/rebar, and ETi/SC = potential differences between rebar and SC half-cell, Ti
electrode and rebar, and Ti electrode and SC half-cell, respectively.

Carbonation tests were performed in accordance with EN 13295 (2004) using 100 × 100 ×
100 mm plain specimens prepared from mixes F0, F40, and F70. Half of the specimens were
tested after 28 days of curing in Environment A (Figure 4.5). The other half of the specimens
were tested after being cured in Environment A for 28 days and in Environment B (Figure
4.5) for another 28 days (in total 56 days). During testing, each specimen was split in the
middle and phenolphthalein indicator (1 g of phenolphthalein in a solution of 70 mL ethanol
and 30 mL water) was sprayed over the split section. The colour change of the split face,
upon spraying phenolphthalein solution, was observed to assess the carbonation resistance
of the corresponding UHPC mix. For each curing condition, three specimens from each
aforementioned mix were tested for carbonation.

4.2.4.4 Microstructure analyses

XRD analyses were conducted for dry powdered samples prepared from F0, F40, and F70
specimens cured in water for 28 days. For each sample, the diffraction pattern in the 2θ
range of 5–70° was recorded using an X-ray diffractometer with Cu K radiation (λ = 1.5404
Å), 40 kV, and 40 mA at a rate of 2°/min.

The F0, F40, and F70 specimens were crushed into small pieces, and then thin layers of
platinum coating were applied on the crushed pieces for SEM and EDS analyses. SEM
imaging was performed using a field emission scanning electron microscope equipped with
an energy dispersive spectrum (EDS) detector.

92
Chapter 4

4.3 Results and discussion

4.3.1 Workability

The slump flow results of the ECO-UHPC mixes are shown in Figure 4.9. To ensure good
workability, the control ECO-UHPC (F0) in this study was designed to have a slump flow
more than twice the diameter of the bottom of the slump cone (i.e., more than 2×76 = 152
mm). The slump flow diameter of F0 was 157.5 mm. The slump flow diameter of fresh
UHPC increased with increases in the f/c ratio. This was expected because the spherical
shapes of fly ash particles cause a ball bearing effect and reduce the friction between the
angular particles present in fresh UHPC (Kwan and Li 2013). The slump flow diameters of
F20, F40, F60, and F70 were respectively 2.34, 2.47, 2.53, and 2.68 times the bottom
diameter of the slump cone. With respect to the slump flow diameter of F0, the flow dimeters
of F40 and F70 were 19% and 29.2% higher, respectively. Again, for given f/c, the slump
flow diameter of fresh UHPC was observed to increase when w/b ratio was increased. The
flow diameters of W0 and W60 were 2.83 and 2.95 times the bottom diameter of the slump
cone, respectively. With respect to F0 and F60, the flow diameters of W0 and W60 were
36.5% and 16.4% higher, respectively.

Figure 4.9. (a) Slump flow results of the ECO-UHPC mixes, (b) typical slump flow of F0, and
(c) typical slump flow of F60

93
Chapter 4

Table 4.4. 28-day mechanical test results of ECO-UHPCs

ECO-UHPC 28-day mechanical properties, MPa (COV, %)


mix Compressive strength Elastic modulus Splitting tensile Flexural
strength strength
F0 198.9 (2.2) 53145 (2.3) 11.7 (3.6) 27.9 (4.6)
F20 179.2 (2.9) 49672 (3.1) 10.1 (4.0) 24.7 (4.8)
F40 152.6 (2.6) 44001 (2.9) 8.6 (4.9) 23.5 (3.7)
F60 123.5 (2.7) 38267 (1.9) 7.9 (2.6) 18.3 (4.8)
F70 106.9 (3.1) 34379 (3.6) 6.4 (4.9) 15.6 (5.7)
W0 172.1 (2.2) - 9.8 (4.4) 23.6 (4.9)
W60 99.1 (3.4) - 4.8 (4.5) 13.6 (3.2)

4.3.2 Mechanical properties

The average 28-day mechanical test results of the ECO-UHPCs are presented in Table 4.4
and Figure 4.10. The top and bottom error bars in Figure 4.10 represent the maximum and
minimum test results, respectively. The COVs of the 28-day mechanical test results are also
presented in Table 4.4. To examine the significance of variation between the mechanical test
results of two mixes, one-way ANOVA was conducted at a significance level (α) of 0.05.
ANOVA outputs for 28-day mechanical test results are presented in Table 4.5.

Figure 4.10. 28-day mechanical test results of the ECO-UHPCs: (a) compressive strength, (b)
elastic modulus, (c) splitting tensile strength, and (d) flexural strength

94
Chapter 4

Table 4.5. ANOVA outputs for 28-day mechanical test results

Mechanical property Mixes Sum of squares F P


Compressive strength F0, F20 582.73 24.87 0.008
F0, F40 3211.92 182.74 0.000
F0, F60 8511.01 552.36 0.000
F0, F70 12675.58 828.46 0.000
F0, W0 1071.74 63.21 0.001
F60, W60 896.26 79.55 0.001
Elastic modulus F0, F20 18100234.91 9.68 0.036
F0, F40 125417275.21 81.90 0.001
F0, F60 332065058.41 336.61 0.000
F0, F70 528266653.44 354.41 0.000
Splitting tensile strength F0, F20 3.65 21.70 0.010
F0, F40 13.83 78.76 0.001
F0, F60 20.83 192.68 0.000
F0, F70 41.03 301.28 0.000
F0, W0 5.36 29.80 0.005
F60, W60 14.63 323.26 0.000
Flexural strength F0, F20 15.33 9.97 0.034
F0, F40 29.13 23.96 0.008
F0, F60 140.07 115.08 0.000
F0, F70 229.28 187.30 0.000
F0, W0 27.95 18.45 0.013
F60, W60 32.62 68.08 0.001

220
Compressive strength (MPa)

180

140

100

F0 F20
60
F40 F60
F70
20
0 10 20 30 40 50 60
Age (day)

Figure 4.11. Compressive strength gains of F-UHPCs with time

4.3.2.1 Compressive strength

The 28-day compressive strengths of the ECO-UHPCs from the F- and W-series are
presented in Figure 4.10(a). Figure 4.11 shows the strength developments of the F-UHPCs
with time. The 28-day and 56-day compressive strengths of F0 were 198.9 MPa and 205.2
MPa, respectively. At the age of 3 days, F0 attained a strength of 127.5 MPa, which was
almost 64% of its 28-day strength. There were minor differences between the rates of strength

95
Chapter 4

gain of the UHPCs with time. Nevertheless, the strength gain rates of the UHPCs with fly
ash were generally low before 7 days compared with that F0. After 7 days, the strength gain
rates of the UHPCs with fly ash appeared to be comparable to or slightly higher than that of
F0. The 7-day strengths of F0, F20, F40, F60, and F70 were respectively about 80%, 75.6%,
72.4%, 65.6%, and 59.4% of their respective 28-day strengths. Irrespective of age, the
compressive strength of UHPC reduced with the increase of cement replacement by Class-
F fly ash. The decrements in the 28-day strengths of F20, F40, F60, and F70 were
respectively 9.9%, 23.3%, 37.9%, and 46.2% with respect to the 28-day strength of F0. Table
4.5 shows that the variations of the 28-day compressive strengths of the F-UHPCs with fly
ash from the 28-day strength of F0 were significant at the 0.05 level (i.e., P-values were less
than 0.05). The 28-day strengths of the W-UHPCs were lower than those of their
counterparts in F-series, which was expected because of the higher w/b ratio of the W-
UHPCs. The 28-day compressive strengths of W0 and W60 were 16.1% and 24.6% lower
with respect to the 28-day strengths of F0 and F60, respectively. The variation of the
compressive strength of W0 from that of F0 and the variation of the compressive strength of
W60 from that of F60 were significant at the 0.05 level (Table 4.5). The COVs of the 28-day
strengths of all the F- and W-UHPC mixes, mentioned in Table 4.4, were within the limit
(3.4%) specified in ASTM C109 / C109M-16a (2016) for blended cement based composite.

Results in Figure 4.10(a) also show that it is possible to produce UHPC with ultra-high
strength (> 150 MPa) up to 40% replacement of cement by fly ash, and a very high strength
(> 100 MPa) can be achieved up to 70% replacement. The performances of all the F-UHPCs
appeared to be better than conventional HSC, in terms of strength. Conventional HSC has
a strength greater than 41 MPa but it is seldom greater than 100 MPa. Therefore, slender
RC members can be constructed with the F-UHPCs compared to those built with HSC. It is
also worth noting that F70 had a cement content of 281 kg/m3, whereas the cement content
of conventional HSC is usually in the range of 400–600 kg/m3 (Yasin et al. 2017). For
massive structures (e.g., raft foundations of skyscrapers), concrete with low cement content
but with high strength is desirable to avoid thermal cracking (Aldred 2010). With such
structures, the use of high-volume fly ash based UHPC like F70 can be advantageous.
Further research at the structural scale is still required to explore the true potential of high-
volume fly ash based UHPC mixes.

The reduction of compressive strength of UHPC by increasing the f/c ratio agrees with the
results reported by previous studies (Ferdosian et al. 2017; Peng et al. 2010). However, some
researchers, like Chen et al. (2018), reported a slight increase in the strength of UHPC for a
low replacement of cement by fly ash. A higher sf/(c+f) ratio (= 0.268) was used to design

96
Chapter 4

the mixes used in the present study than that used in the mixes of (Chen et al. 2018) [sf/(c+f)
= 0.125 in (Chen et al. 2018)]. Possibly the presence of a higher amount of silica fume
positively influenced the strength of the control mix of this study, since silica fume in UHPC
helps transform calcium hydroxide (CH) crystals into denser calcium silicate hydrate (C-S-
H) gel. If there is an optimum amount of silica fume in a UHPC mix to react with the CH
crystals produced from a given amount of Portland cement, replacing cement by other
pozzolanic materials (e.g., Class-F fly ash) may not benefit the compressive strength of
UHPC but may rather impair the strength owing to the dilution effect (reduction of cement
content) (Bonavetti et al. 2000). On the other hand, if the silica fume content is less than the
optimum, replacing certain amount of Portland cement by a pozzolanic material may
increase the strength facilitating further conversion of CH to C-S-H [and calcium
aluminosilicate hydrate (C-A-S-H), which is more common in concrete with Class-F fly ash
(Berry et al. 1994)]. Hence, the strength increment upon low substitution of cement by fly
ash, reported in (Chen et al. 2018), could be a result of using a low sf/(c+f) ratio in the mixes.
In addition to the contents and chemical compositions of the binders (e.g., Portland cement,
silica fume, and fly ash), the PSDs of the materials used in UHPC may influence its
compressive strength. In the case of the present study, though the mixes with higher fly ash
content had higher compactness (Figure 4.3), the compressive strength of UHPC was found
to reduce with increasing fly ash content. This is most likely because the dilution effect was
dominant in influencing the compressive strengths of the UHPCs investigated in this study.

A number of researchers attempted to relate the compressive strength of NSC and HSC to
the chemical compositions of their binders (Kuder et al. 2012; Xie and Visintin 2018; Xie et
al. 2018). Xie and Visintin (2018) proposed a model to predict compressive strength of NSC
and HSC from the reactivity moduli (RMs) of binders [namely, hydration modulus (HM);
silica modulus (SM); and alumina modulus (AM)], w/b ratio (on mass basis), and b/a ratio
(on volume basis) in the form of Equation 4.5.

𝛼1 𝐻𝑀 + 𝛼2 𝑆𝑀 + 𝛼3 𝐴𝑀
𝑓𝑐 = (𝑏/𝑎)𝑛 (4.5)
𝑤/𝑏

where the values of α1, α2, α3, and n are 11.7, 0.62, 0.94, and 0.24 respectively, for NSC and
HSC.

The RMs of a binder are expressed in terms of its oxide contents, as shown in Equations 4.6–
4.8.

97
Chapter 4

𝐶𝑎𝑂 + 𝑀𝑔𝑂 + 𝐴𝑙2 𝑂3


𝐻𝑀 = (4.6)
𝑆𝑖𝑂2

𝑆𝑖𝑂2
𝑆𝑀 = (4.7)
𝐴𝑙2 𝑂3 + 𝐹𝑒2 𝑂3

𝐴𝑙2 𝑂3
𝐴𝑀 = (4.8)
𝐹𝑒2 𝑂3

where HM is representative of the hydraulic reactivity of the binder, and SM and AM are
representative of the pozzolanic reactivity of the binder.

In case of a blended binder system, RMs can be evaluated using Equations 4.9–4.11.

𝐻𝑀 = ∑𝑛𝑖=1 𝐻𝑀𝑖 ∙ 𝑚𝑖 (4.9)

𝑆𝑀 = ∑𝑛𝑖=1 𝑆𝑀𝑖 ∙ 𝑚𝑖 (4.10)

𝐴𝑀 = ∑𝑛𝑖=1 𝐴𝑀𝑖 ∙ 𝑚𝑖 (4.11)

where HMi, SMi, and AMi = hydration modulus, silica modulus, and alumina modulus of
each individual binder present in blended binder system, respectively; and mi = mass fraction
of each binder. Since RMs take into account both the chemical compositions and the
contents of all the binders present in a blended binder system, establishing a relationship in
the form of Equation 4.5 for a particular type of concrete provides a means to predict the
compressive strength of that type of concrete from the chemical compositions and the
contents of its binders.

The composition of UHPC is different from that of NSC or HSC, in that the former usually
contains a w/b ratio less than 0.2, while the latter generally has w/b ratio much higher than
0.2. Moreover, the b/a ratio of UHPC is 3–5 times the b/a of NSC or HSC. The model
proposed by Xie and Visintin (2018) was therefore slightly modified for UHPC incorporating
fly ash, as shown in Equation 4.12, based on a non-linear regression analysis of the
experimental data of this study and the dataset found in the literature. The data-points are
presented in Figure 4.12. The dataset includes data from all previous studies on standard-
cured cement – silica fume – fly ash based UHPC, reporting: the chemical compositions of
fly ash and other binders used and detailed mix proportion. The w/b and b/a in Equation
4.12 are on mass basis. Because of a lack of data in the literature, the PSDs of the materials
were not taken into consideration in developing the relationship.

7.72𝐻𝑀 + 0.12𝑆𝑀 + 1.81𝐴𝑀


𝑓𝑐 = (𝑏/𝑎)0.34 (4.12)
𝑤/𝑏

98
Chapter 4

Figure 4.12. Relationship between 28-day compressive strength and reactivity moduli for
standard-cured UHPC with fly ash

4.3.2.2 Elastic modulus

Figure 4.10(b) shows the 28-day elastic moduli of the F-UHPCs. The average elastic
modulus of F0 was 53145 MPa at 28 days. The average elastic moduli of F20, F40, F60,
and F70 reduced respectively by 6.5%, 17.2%, 28%, and 35.3% with respect to the elastic
modulus of F0. The variations of the 28-day elastic moduli of the F-UHPCs with fly ash
from that of F0 were significant at the 0.05 level (Table 4.5).

The elastic moduli of the F-UHPCs are plotted against their respective compressive strengths
in Figure 4.13, which makes it possible to evaluate the response of elastic modulus of UHPC
with fly ash as a function of compressive strength. To understand the relationship between
the elastic modulus and the compressive strength of standard-cured UHPC in general, elastic
modulus versus compressive strength data points collected from previous studies
investigating standard-cured UHPC are also presented in Figure 4.13. Data from UHPCs
with coarse aggregate are not included in Figure 4.13. For comparison, models suggested by
different standards to estimate elastic moduli of NSC and HSC (ACI 318 2014; ACI 363R
1992; AS 3600 2009), and models proposed by previous studies to predict elastic moduli of
UHPC and UHPFRC (Alsalman et al. 2017; Graybeal 2007; Ma et al. 2004) are presented
in Figure 4.13 as well. Table 4.6 summarizes the equations representing the previously
proposed models.

99
Chapter 4

It is obvious from Figure 4.13 that the ACI 318 (2014) and AS 3600 (2009) models
overestimate the elastic modulus of UHPC. This may be because these models were mainly
proposed for concrete with coarse aggregate (which is usually stiffer than cement matrix),
while coarse aggregate is not generally used in UHPC. Ma et al. (2004) observed that, for a
given compressive strength, UHPC with basalt coarse aggregate exhibited higher elastic
modulus than UHPC without any coarse aggregate. The researchers attributed this
phenomenon to the high stiffness of basalt aggregate.

Figure 4.13. Relationship between elastic modulus and compressive strength of standard-cured
UHPC (points in black circles represent data from cement – silica fume – fly ash based
UHPCs)

Table 4.6. Elastic modulus – compressive strength models proposed by standards and previous
studies

Standard/ reference Equation Concrete type


Alsalman et al. (2017) 𝐸𝑐 = 8010 (𝑓𝑐 )0.36 UHPC and UHPFRC
Graybeal (2007) 𝐸𝑐 = 3840 √𝑓𝑐 UHPC and UHPFRC
Ma et al. (2004) 1
UHPC
𝑓𝑐 3
𝐸𝑐 = 19000 ( )
10
ACI 318 (2014) 𝐸𝑐 = 4700 √𝑓𝑐 NSC
ACI 363R (1992) 𝐸𝑐 = 3320√𝑓𝑐 + 6900 HSC
AS 3600 (2009) 𝐸𝑐 = 5050 √𝑓𝑐 NSC and HSC
Ec = elastic modulus (MPa), fc = compressive strength (MPa)

100
Chapter 4

The ACI 363R (1992) model, and the models proposed by Alsalman et al. (2017), Graybeal
(2007) and Ma et al. (2004) for UHPC or UHPFRC, tend to give better estimations of the
elastic modulus of standard-cured UHPC than ACI 318 (2014) and AS 3600 (2009) models.
Nevertheless, the models still overestimate the elastic moduli of UHPCs with compressive
strengths below 135 MPa. This may be due to a number of reasons. First, the ACI 363R
(1992) model was proposed for HSC, which also comprises coarse aggregate similar to NSC,
although the paste volume of HSC is usually slightly higher and coarse aggregate content is
slightly lower than those of NSC (Sbia et al. 2015). Second, the models proposed by
Alsalman et al. (2017), Graybeal (2007), and Ma et al. (2004) for UHPC are based on the
test data from UHPCs or UHPFRCs cured in both standard and special curing regimes (e.g.,
autoclaving, steam curing, heat curing). Shen et al. (2019) applied nanoindentation
technique to study the elastic moduli of the hydration products of UHPCs cured in different
regimes. Their findings suggest that steam curing and autoclaving impart high elastic moduli
to the hydration products of UHPC, which in turn may cause high elastic modulus of steam-
cured or autoclaved UHPC compared with that cured at ambient temperature. Third, the
models proposed by Graybeal (2007) and Ma et al. (2004) were developed based on the
results from UHPCs with only quartz powder and/or refined quartz sand as aggregates. An
experimental study conducted by Alsalman et al. (2017) suggests that untreated natural sand
may contain soft particles, and its use as aggregate may impair the elastic modulus of UHPC.
This may be a reason behind the overestimation of the data taken from (Alsalman et al.
2017) by the models proposed by Graybeal (2007) and Ma et al. (2004). These models
overestimate the experimental data of the F-UHPC with 40% or more replacement of
cement by Class-F fly ash. Previous studies on NSC and HSC (Gholampour and
Ozbakkaloglu 2017; Huang et al. 2013; Kayali and Sharfuddin Ahmed 2013) confirmed that
high-volume substitution of Portland cement by fly ash causes pronounced reduction in the
elastic modulus of concrete, which also seems to be true in the case of UHPC. It is worth
noting that the ratios of elastic modulus to compressive strength of the fly ash based UHPCs
reported in (Mueller et al. 2016) were higher than those of this study. Mueller et al. (2016)
used higher quartz aggregate contents to design UHPC mixes than that used in the current
study, which possibly induced higher elastic moduli in the mixes reported in their study
(Zhao and Sun 2014). The b/a ratios of the mixes in (Mueller et al. 2016) were in the range
of 0.70–0.74 by volume, whereas a b/a ratio of 1.06 by volume was used in this study.

Based on the data points presented in Figure 4.13, a more conservative representation of the
relationship between the elastic modulus and the compressive strength of standard-cured
UHPC would be Equation 4.13.

101
Chapter 4

𝐸𝑐 = 2661 (𝑓𝑐 )0.56 [80 MPa ≤ fc ≤ 200 MPa] (4.13)

4.3.2.3 Splitting tensile strength and flexural strength

The 28-day splitting tensile strengths and flexural strengths of the ECO-UHPCs are shown
in Figures 4.10(c) and 4.10(d), respectively. Without the incorporation of fibres, the average
28-day splitting tensile strengths of F0, F20, F40, F60, and F70 were 11.7 MPa, 10.1 MPa,
8.6 MPa, 7.9 MPa, and 6.4 MPa, respectively, while the average flexural strengths of the
UHPCs were respectively 27.9 MPa, 24.7 MPa, 23.5 MPa, 18.3 MPa, and 15.6 MPa at 28
days. Table 4.5 shows that the variations of the splitting tensile strengths or the flexural
strengths of the F-UHPCs incorporating fly ash from that of F0 were significant at the 0.05
level. The splitting tensile strength and the flexural strength of UHPC reduced when w/b
was increased from 0.164 to 0.193. The splitting tensile strengths of W0 and W60 were
respectively 16.2% and 39.3% lower with respect to those of F0 and F60, and their flexural
strengths were 15.1% and 25.1% lower, respectively. The variation between the splitting
tensile strengths or flexural strengths of F0 and W0 or those of F60 and W60 was significant
(Table 4.5).

Figure 4.14 presents splitting tensile strength as a function of compressive strength based on
the results obtained in this study and from previously conducted studies on standard-cured
UHPC. The splitting tensile strength – compressive strength relationships proposed by ACI
318 (2014), ACI 363R (1992), and AS 3600 (2009) (as listed in Table 4.7) seem to be
conservative, especially for UHPCs having compressive strengths greater than 100 MPa.
This is because the relationships were primarily developed for NSC and HSC, both of which
contain coarse aggregate. In NSC and HSC, cracks generate around the interface between
paste and coarse aggregate under external load. Exclusion of coarse aggregate from UHPC
reduces such crack generation to a large extent and thus enhances the splitting tensile
strength of the concrete (Richard and Cheyrezy 1995). Moreover, the paste – aggregate
interface in UHPC is usually much denser than NSC or HSC because of the use of very fine
sand as aggregate and the pozzolanic reactivity of silica fume (Ghafor et al. 2020; Zhu et al.
2015). A more reasonable relationship between the splitting tensile strength and the
compressive strength of standard-cured UHPC would be Equation 4.14.

𝑓𝑠𝑝 = 0.095 𝑓𝑐 0.92 [70 MPa ≤ fc ≤ 200 MPa] (4.14)

102
Chapter 4

Table 4.7. Splitting tensile strength – compressive strength and flexural strength – compressive
strength models proposed by standards and previous studies

Relationship Standard/ reference Equation Concrete type


Splitting tensile strength – ACI 318 (2014) 𝑓𝑠𝑝 = 0.56 √𝑓𝑐 NSC
compressive strength ACI 363R (1992) 𝑓𝑠𝑝 = 0.59 √𝑓𝑐 HSC
AS 3600 (2009) 𝑓𝑠𝑝 = 0.4√𝑓𝑐 NSC and HSC
Flexural strength – Ahmad et al. (2015) 𝑓𝑟 = 3.08 √𝑓𝑐 UHPFRC (with 6.2%
compressive strength steel fibre by mass)
ACI 318 (2014) 𝑓𝑟 = 0.62 √𝑓𝑐 NSC
ACI 363R (1992) 𝑓𝑟 = 0.94 √𝑓𝑐 HSC
AS 3600 (2009) 𝑓𝑟 = 0.6√𝑓𝑐 NSC and HSC
fsp = splitting tensile strength (MPa), fr = flexural strength (MPa), fc = compressive strength (MPa)

Figure 4.14. Relationship between splitting tensile strength and compressive strength of
standard-cured UHPC (points in black circles represent data from cement – silica fume – fly
ash based UHPCs)

The flexural strengths of the F- and W-UHPCs are plotted against their respective
compressive strengths in Figure 4.15. Also presented in Figure 4.15 are 197 datapoints
collected from previous studies investigating standard-cured UHPC. In total, 204 data points
are presented (7 from this study + 197 from the literature). The use of different types and
contents of materials as binders and aggregates for producing the UHPCs in different studies
accounts for the scatter of the presented data. For instance, the use of nano-silica or nano-
calcium carbonate as supplementary binder in UHPC tended to give high ratios of flexural
strength to compressive strength compared with UHPC without a nano-binder (Li et al.

103
Chapter 4

2015; Wu et al. 2018). Again, the use of river sand (Liu et al. 2018; Song et al. 2018) or
recycled glass cullet (Soutsos et al. 2005), for example, as aggregate instead of conventional
quartz sand tended to impart low flexural strength to UHPC. Nevertheless, a reasonable fit
for the data presented in Figure 4.15 can be expressed by Equation 4.15. It is also worth
mentioning that Equation 4.15 provides relatively satisfactory estimations, with an R2 value
of 0.76, for the flexural strength of cement – silica fume – fly ash based UHPC containing
quartz sand and quartz powder as aggregates.

𝑓𝑟 = 0.4 𝑓𝑐 0.79 [70 MPa ≤ fc ≤ 200 MPa] (4.15)

The equations corresponding to the previously proposed flexural strength versus


compressive strength relationships are presented in Table 4.7. Lines (in Figure 4.15)
representing the models proposed by ACI 318 (2014), ACI 363R (1992), and AS 3600 (2009)
to estimate flexural strengths of HSC and NSC suggest that these models tend to
underestimate the flexural strength of standard-cured UHPC. Again, the model proposed by
Ahmad et al. (2015) for UHPFRC overestimates all the data points shown in Figure 4.15.
This was expected because fibre incorporation enhances the performance of UHPC in
flexure.

Figure 4.15. Relationship between flexural strength and compressive strength of standard-
cured UHPC

104
Chapter 4

4.3.3 Durability properties

4.3.3.1 Water absorption (WA) and initial rate of absorption (IRA)

The 7-day and 28-day WA and IRA results of the F-UHPCs are shown in Figures 4.16(a)
and 4.16(b), respectively. The 7-day WAs and IRAs of F0, F20, F40, F60, and F70 are
0.56%, 1.05%, 0.84%, 0.78%, and 1.28% and 0.032 kg/m2/min, 0.062 kg/m2/min, 0.049
kg/m2/min, 0.055 kg/m2/min, and 0.072 kg/m2/min, respectively. At 28 days, the WAs
and IRAs of the F-UHPCs were observed to be 0.37%, 0.42%, 0.57%, 0.43%, and 1.01%
and 0.029 kg/m2/min, 0.027 kg/m2/min, 0.039 kg/m2/min, 0.032 kg/m2/min, and 0.069
kg/m2/min, respectively. Two opposing factors may be responsible for influencing the pore-
structure of UHPC when cement is partially substituted by fly ash. On the one hand, fly ash
particles, owing to their high fineness compared with cement particles (from Table 4.2, d50
of cement = 17.3 µm, d50 of fly ash = 9.8 µm), tend to fill pores of UHPC mix (filler effect)
(Chindaprasirt and Rukzon 2008). On the other hand, in UHPC with optimized silica fume
content, Class-F fly ash inclusion may reduce the overall formation of C-S-H gel; moreover,
high fly ash content may also slightly slow down the formation of C-S-H (as indicated by
the compressive strength test results). As a result, the reduction of water-permeable pores
due to the formation of C-S-H may become less pronounced in UHPC with high Class-F fly
ash content, especially at an early age.

At 7 days, the F-UHPCs incorporating fly ash, regardless of fly ash content, exhibited higher
WAs and IRAs than F0. However, the differences between the WAs or IRAs of F0 and
UHPCs incorporating fly ash (except F70) became lower at 28 days. This may be because,
in F0, the formation of a large portion of its total C-S-H at 28 days took place at 7 days.
Again, the 7-day WAs and IRAs of F40 and F60 were lower than those of F20. Better
packing of particles in F40 and F60 than that in F20 (owing to the filler effect) most likely
offset the effect of lower C-S-H formation in the former at 7 days. However, with the passage
of time, as more C-S-H formed in F20, the differences between the WA or IRA of F20 and
those of F40 and F60 became lower. Consequently, the 28-day WA and IRA of F20 were
slightly lower than those of F40 and comparable to those of F60. The WA and IRA of F70
were not observed to be benefitted much from the filler effect at 7 or 28 days, possibly
because of a substantially low formation of C-S-H in this concrete. Nevertheless, the WAs
of all the F-UHPCs, both at 7 and 28 days, were far less than 2%, which is the maximum
limit for the concrete being considered as adequately durable for construction works with
design life of 50 years (Elchalakani et al. 2017).

105
Chapter 4

Figure 4.16. (a) Water absorption, and (b) initial rate of absorption test results of F-UHPCs at
7 and 28 days

Age (day)
0 5 10 15 20 25 30
-100

-150
Half-cell potential (mV)

-200

-250

-300 F0 F20 F40

F60 F70
-350

Figure 4.17. Half-cell potential test results of F-UHPCs

4.3.3.2 Half-cell potential (HCP)

Figure 4.17 presents the HCP test results of the F-UHPCs at different ages. The 3-day HCPs
of all the UHPCs were higher than −350 mV. In all cases, HCP tended to become less
negative with curing time, implying a reduced risk of corrosion of the embedded rebars at a
later age (up to 28 days). The results comply with previous studies investigating the
development of HCP of NSC and HSC with time in humid environment (Elchalakani et al.
2017; Kayali and Zhu 2005). The 7-day HCP of F0 was less negative than the HCPs of the
UHPCs with fly ash. At 7 days, the HCPs of F0, F20, F40, F60, and F70 were –183.4 mV,

106
Chapter 4

–215.9 mV, –208.5 mV, –201.5 mV, and –229.4 mV, respectively. The HCP increase rate
(rate of becoming less negative) with time became less pronounced for F0 after 7 days
compared with the UHPCs incorporating fly ash. For F0, the HCP increase rate between 7
and 28 days was 2.6 mV/day; whereas for the UHPCs with fly ash, the values were greater
than 3.5 mV/day. The HCP increase rate with time was found to be most pronounced for
F20 (4.2 mV/day) between 7 and 28 days. The 28-day HCPs of F20, F40, and F60 were
comparable to that of F0. The HCPs of F0, F20, F40, F60, and F70 were –129.3 mV, –127.1
mV, –133.7 mV, –123.5 mV, and –154.1 mV, respectively, at 28 days. Reduced porosities
(based on the WA and IRA results) of F20, F40, and F60 at 28 days, compared to their
porosities at 7 days, possibly helped reduce the corrosion risk of rebars in these UHPCs and
attain comparable behaviour to that of F0 at 28 days (Yodsudjai and Pattarakittam 2017).
Similar phenomena were observed by Sujjavanich et al. (2005) and Zou et al. (2016) for
NSCs with up to 50%–65% fly ash content.

4.3.3.3 Carbonation resistance

Upon application of phenolphthalein indicator, a purple or pink colour was observed all
over the cross-sections of both F0 and F40 specimens after 28 days of exposure to
Environment A or after 28 days of exposure to Environment A and followed by 28 days
exposure to Environment B (Figure 4.18). This implies that, irrespective of curing
environment, the depth of carbonation was below the detection limit (i.e., <0.5 mm) in F0
and F40 specimens as per EN 13295 (2004). These observations are consistent with the
results obtained by Long et al. (2005), and Scheydt and Müller (2012) from their
investigations on conventional UHPC without fly ash. The colour of the cross-sections of
the F70 samples, regardless of exposure condition, appeared to be light upon the application
of phenolphthalein indicator in comparison to F0 and F40 specimens. This could be a result
of low CaO content (because of low cement content) in F70, compared to F0 or F40, which
led to the formation of a low amount of CH, and/or C-S-H and C-A-S-H, and hence
imparted low alkalinity to F70. However, neither a carbonation front nor any totally
uncoloured region was observed in any of the F70 specimens (i.e., carbonation depth was
below the detection limit regardless of curing condition). This phenomenon was interesting
to observe since such high-volume incorporation of fly ash did not induce any increase in
the depth of carbonation of F70 specimens after exposure to Environment B. The dense
microstructure of UHPC is most likely the key factor that imparts such behaviour. For HSC
with high-volume fly ash content, Shi et al. (2009) concluded that its carbonation is greatly
influenced by its w/b ratio. In their study, the carbonation depth of HSC with 60% fly ash
was observed to reduce by about 65% when the w/b ratio was reduced from 0.3 to 0.25.

107
Chapter 4

Similar behaviour was observed by Wang and Park (2016). It can therefore be inferred that
the very low w/b (= 0.166) of the F-UHPCs investigated in this study substantially reduced
the diffusion of CO2 into them and resulted in carbonation depth below the detection limit.

4.3.4 Microstructure analyses

Figure 4.19 shows the XRD patterns of F0, F40, and F70 powders. The crystalline phases
of F0 were ettringite (AFt), portlandite [Ca(OH)2 or CH], alite, belite, and quartz. The
intensity of portlandite was found to be less pronounced in the XRD pattern of F0 compared
with those reported for normal-strength mortars (Ashraf et al. 2009; Korpa et al. 2009). This
was expected, because the dense microstructure of UHPC, owing to its very low w/b ratio,
does not provide much room for the development of CH crystals (French 1991; Larbi and
Bijen 1990). Furthermore, silica fume in UHPC reduces portlandite intensity by converting
a portion of it into C-S-H. The C-S-H phase could not be identified in the XRD pattern of
any of the powders, owing to its nano-crystalline structure (Horszczaruk et al. 2015; Tabet
et al. 2018). However, the low intensities of portlandite and ettringite indicate that the main
hydration product was C-S-H (Korpa et al. 2009). Slightly lower intensities of portlandite
and clinker minerals in the XRD pattern of F40 than those of F0 were the results of lower
cement content in the former than that in the latter. The portlandite peak almost disappeared
in the XRD pattern of F70. This is because the F70 mix had a very low amount of cement
but a bulk quantity of pozzolanic material. Higher intensity of quartz at 27° and slightly
more pronounced mullite peaks at 16.5° and 26.5° in the XRD pattern of F70 than F40
indicate the presence of higher amount of unreacted fly ash in F70.

Figure 4.18. Illustration of carbonation resistances of specimens (a) F0, (b) F40, and (c) F70
after 28 days of exposure to Environment A (300 ppm CO2), followed by 28 days of exposure
to Environment B (10,000 ppm CO2)

108
Chapter 4

Figure 4.19. XRD patterns of F0, F40 and F70 (magnified intensity – left) (E = ettringite, P =
portlandite)

The SEM images of F0, F40, and F70 are shown in Figure 4.20. Atomic percentages and
ratios of the major elements at several points and areas of F0, F40, and F70, derived from
EDS analysis, are also shown in Figure 4.20. F40 was observed to contain unreacted fly ash
particles (marked F) and partially reacted hollow fly ash spheres (marked H). More
unreacted fly ash particles were observed in the SEM image of F70, which confirms the
XRD results. This was expected because lower lime content in F70 restricted the formation
of portlandite and, consequently, caused more fly ash particles to remain unreacted. This
again indicates the formation of a lower amount of C-S-H gel in F70, the effect of which
could be observed on its mechanical properties (i.e., F70 exhibiting lower strength).
Anhydrous cement particles (marked C) were observed in F0.

4.3.5 CO2 emission and cost analysis

Figure 4.21(a) presents the approximate CO2 emissions from the unit production of F-
UHPCs and other UHPCs reported in the literature having different fly ash contents. For
comparison, the CO2 emission from HSC, as reported in (Elchalakani et al. 2017), is also
shown in Figure 4.21(a). The CO2 emissions from conventional UHPCs (i.e., UHPCs
without fly ash) in general exceed 850 kg/m3. The CO2 footprint of UHPC reduces with
increasing replacement of cement by fly ash. The CO2 footprints of F20, F40, F60, and F70
are respectively 16.7%, 33.3%, 50%, and 58.3% lower, with respect to the CO2 footprint of
F0.

109
Chapter 4

Figure 4.20. SEM images and EDS results of F0, F40 and F70 (C = unreacted cement, F =
unreacted fly ash, H = partially reacted fly ash)

110
Chapter 4

Figure 4.21. (a) CO2 footprints, and (b) CO2 intensities of different types of concrete

The CO2 footprints of UHPCs are generally higher than that of conventional HSC.
However, it should be noted that the high compressive strength of UHPC will reduce the
size of structural members and thus reduce the volume of material required for construction
compared with HSC. Hence, the CO2 intensities (ci) [CO2 footprint per unit compressive
strength (Damineli et al. 2010)] of different types of concretes are presented in Figure 4.21(b)
in order to get a better understanding of the ecological influence of these concretes. The cis
of the F-UHPCs are lower than that of HSC. The ci of the UHPC with f/c > 20/80, reported
in (Chen et al. 2018), is comparable to that of HSC. It can also be seen from Figure 4.21(b)
that the cis of all the F-UHPCs lie below the international future benchmark of 5
kg/m3/MPa (Damineli et al. 2010; Elchalakani et al. 2017).

Figure 4.22 presents the material costs of different types of concrete [in Australian dollars,
A$]. The material cost of HSC is generally less than half the cost of UHPC because of the
inclusion of coarse aggregate. Use of non-conventional low-cost material, such as river sand,
may reduce the cost of UHPC, as can be seen from Figure 4.22 for the UHPC reported in
(Song et al. 2018). Substitution of cement by fly ash reduces the material cost of UHPC. The
costs per cubic meter of F0, F20, F40, F60, and F70 are A$1692, A$1648, A$1603, A$1558,
and A$1535, respectively. Nevertheless, the costs of UHPCs incorporating fly ash are still
high compared to HSC. This is because commercially available fly ash is not inexpensive.
The use of unprocessed fly ash may be economically more attractive. A study can be
undertaken in the future to investigate the performance of UHPC incorporating unprocessed
fly ash. Table 4.8 summarizes the data used for the calculation of the CO2 footprints and the
costs of different concretes presented in Figures 4.21 and 4.22, respectively.

111
Chapter 4

Figure 4.22. Costs of different types of concrete (A$ = Australian dollars)

Table 4.8. CO2 footprints and costs of different materials

Material Cement Silica Fly Crushed Natural Water Admixturea


fume ash aggregate aggregate
CO2 emission 959 82 93 21 5 1 508
per tonne of
material (kg)
(Elchalakani
et al. 2017)
Cost per 0.75 0.52 0.51 0.56 (quartz 0.12 (river 0.0015 4.4 (HRWR)
kilogram of sand) sand) 3.23 (defoaming
material 0.7 (quartz agent)
(A$) powder)
0.11
(crushed
stone)
a
Including liquid content

4.4 Conclusions

Based on the results presented in the preceding section, the following conclusions can be
drawn:

1. With proper adjustments to the mix design parameters, such as silica fume to cement
ratio, w/b ratio, b/a ratio, etc., it is possible to produce UHPC having a 28-day
compressive strength of 198.9 MPa, a splitting tensile strength of 11.7 MPa and a
flexural strength of 27.9 MPa without any special curing or incorporation of fibres.
2. The strength of UHPC generally reduces with increasing replacement of cement by
Class-F fly ash. Nevertheless, ultra-high strength (>150 MPa) can still be achieved up to
40% replacement of Portland cement by fly ash. The splitting tensile strength and

112
Chapter 4

flexural strength obtained for F-UHPC with 40% fly ash are respectively 8.6 MPa and
23.5 MPa. It is interesting to note that the compressive strength and flexural strength
attained in this study for F-UHPC with 40% Class-F fly ash are higher than previously
reported strengths for standard-cured UHPC with lower Class-F fly ash content (e.g.,
30%). F-UHPC with 70% fly ash, despite having lower cement content (281 kg/m3) than
conventional HSC (400–600 kg/m3), exhibits a very high 28-day compressive strength
of 106.9 MPa. Such low-cement concrete with high strength is desirable in the
construction of massive structures (e.g., raft foundation of skyscraper) to avoid thermal
cracking. F-UHPC with 70% fly ash exhibits a splitting tensile strength of 6.4 MPa and
a flexural strength of 15.6 MPa. The F-UHPCs exhibit better performance than
conventional HSC since conventional HSC seldom exhibits strength greater than 100
MPa. Therefore, slenderer structural members can be constructed with the F-UHPCs.
3. The workability of fresh UHPC increases with increasing substitution of cement by
Class-F fly ash. F-UHPCs with 40% and 70% fly ash respectively exhibit 19% and 29.2%
higher slump flow diameters with respect to F-UHPC without fly ash.
4. A model has been presented to predict the 28-day compressive strength of UHPC
incorporating fly ash from the reactivity moduli of its binders (related to the contents
and the chemical compositions of the binders), w/b ratio, and b/a ratio. Though quality-
controlled Class-F fly ash can nowadays be sourced from several commercial
companies, the chemical composition of this material may vary from source to source.
Trial mixes are therefore necessary before using a new material in construction. In such
a case the predictive model presented in this study may help reduce the number of trials
because the model considers the chemical compositions of all binders to be used in
UHPC production. The model currently does not consider the PSDs of materials.
Further study can be conducted to extend the model taking into account the PSDs of
materials.
5. Relationships have been developed to estimate the elastic modulus, splitting tensile
strength, and flexural strength of standard-cured UHPC from its compressive strength,
based on the data of this study and the data collected from the literature.
6. With up to 60% replacement of cement by Class-F fly ash, UHPC exhibits a durability
performance similar to that of the UHPC without fly ash, in terms of WA, IRA, and
HCP. For UHPCs made with up to 70% replacement of cement by fly ash, the depths of
carbonation remain below the detection limit (<0.5 mm).
7. The CO2 footprint of UHPC can be reduced by 33.3% and 58.3% substituting 40% and
70% of cement by fly ash, respectively. All the ECO-UHPCs, with w/b = 0.164,
developed in this study have CO2 intensities less than 5 kg/m3/MPa.

113
Chapter 4

Acknowledgements

The authors acknowledge the facilities provided by the Centre for Microscopy,
Characterization & Analysis, The University of Western Australia, for conducting the
microstructure analyses. Special thanks are given to Adelaide Brighton Cement, Australia;
Mapei, Australia; and Sika, Australia for providing respectively the cement,
superplasticizer, and defoamer used in this study. The authors thank Ms. Alexandra Meek
for providing guidance to conduct the durability tests. The authors also thank Mr. Allan Xu,
Mr. Samuel Bouffler, Mr. Shane Walshe, and Mr. Kai Zhao for their contributions during
the laboratory work. The research was carried out while the first author was in receipt of a
University Postgraduate Award and the Australian Government Research Training
Program Scholarship at The University of Western Australia.
(Chen et a l. 2 018; Hua ng et a l. 20 19; Li e t al . 201 5; Liu et al . 20 18; M ouna nga et al. 2012 ; Peng et al . 20 10; Song e t al. 2018 ; Wang et al . 20 16; Wu et al . 20 16b ; Zhu e t al . 201 6)
(Abbas et al. 2015 ; ACI 318 2 014 ; ACI 363R 1 992; A hmad e t al . 201 4; Alsalma n et al. 2017 ; A S 3 600 2 009; Cwirze n et a l. 2 008; Ges ogl u et a l. 2 016b ; Graybeal 200 7; Hannaw i et a l. 2 016; Holsc hemacher et a l. 2 004; Ma et a l. 2 004; Máca et al . 20 12; M ueller e t al. 201 6; Sovj ák et a l. 2 013; Wu e t al . 201 6a)
(Abbas et al. 2015 ; Ab dulkareem et a l. 2 018; ACI 3 18 2 014; ACI 3 63R 19 92; AS 360 0 200 9; D hunda si a nd Kha dira naikar 201 9; Ges ogl u et al . 20 16b ; a; G oaiz et al. 2018 ; Han e t al. 201 7; K arihaloo a nd Vriese 1999 ; Rua n e t al. 2018)
(Abdulkareem et a l. 2 018; ACI 3 18 20 14; ACI 3 63R 19 92; Ahm ad et al. 2015 ; A S 3 600 2 009; Che n et a l. 20 18; C ourtia l et a l. 2 013; Cwirze n et a l. 20 08; D hundasi a nd K hadirana ikar 2 019; G upta 2 016; Ha n et al. 2017 ; Hua ng et a l. 2 017 , 201 9; Li et al. 2015 ; Liu e t al. 201 8; Máca et al. 2012 ; Mao e t al. 201 9; M ounanga et a l. 2 012; Ng e t al . 201 4; R ong et al . 20 10; R ua n et al . 20 18; Schwartze ntr uber et a l. 2 004; Song e t al . 201 8; Soutsos e t al . 200 5; Sov ják e t al. 2013 ; Ta fraoui e t al . 200 9; Wu et al . 20 16b; Yu et al . 20 14, 2 015b ; a; Yuns heng e t al . 200 8; Zhu et a l. 2 016)

References
Abbas, S., Soliman, A. M., and Nehdi, M. L. (2015). “Exploring mechanical and durability
properties of ultra-high performance concrete incorporating various steel fiber lengths
and dosages.” Construction and Building Materials, 75, 429–441.
Abdulkareem, O. M., Ben Fraj, A., Bouasker, M., and Khelidj, A. (2018). “Effect of
chemical and thermal activation on the microstructural and mechanical properties of
more sustainable UHPC.” Construction and Building Materials, 169, 567–577.
ACI 318. (2014). Building Code Requirements for Structural Concrete. Farmington Hills, MI.
ACI 363R. (1992). State-of-the-Art Report on High-Strength Concrete.
Ahmad, S., Hakeem, I., and Azad, A. K. (2014). “Effect of curing, fibre content and
exposures on compressive strength and elasticity of UHPC.” Advances in Cement Research,
27(4), 233–239.
Ahmad, S., Zubair, A., and Maslehuddin, M. (2015). “Effect of key mixture parameters on
flow and mechanical properties of reactive powder concrete.” Construction and Building
Materials, 99, 73–81.
Aldred, J. (2010). “Burj Khalifa - A new high for high-performance concrete.” Proceedings of
the Institution of Civil Engineers: Civil Engineering, 163(2), 66–73.
Alsalman, A., Dang, C. N., Prinz, G. S., and Hale, W. M. (2017). “Evaluation of modulus
of elasticity of ultra-high performance concrete.” Construction and Building Materials, 153,
918–928.
Alsayed, S. H., and Amjad, M. A. (1996). “Strength, Water Absorption and Porosity of
Concrete Incorporating Natural and Crushed Aggregate.” Journal of King Saud University
- Engineering Sciences, 8(1), 109–119.
AS/NZS 3582.1. (2016). Supplementary cementitious materials for use with portland and blended
cement - Fly ash.
AS/NZS 3582.3. (2016). Supplementary cementitious materials for use with portland and blended
cement - amorphous silica.
AS/NZS 4456.17. (2003). Masonry units and segmental pavers and flags - methods of test:
determining initial rate of absorption (suction).

114
Chapter 4

AS/NZS 4671. (2001). Steel reinforcing materials.


AS 1012.10. (2000). Methods of testing concrete: determination of indirect tensile strength of concrete
cylinders (“Brazil” or splitting test).
AS 1012.11. (2000). Methods of testing concrete: determination of the modulus of rupture.
AS 1012.17. (1997). Methods of testing concrete: determination of the static chord modulus of
elasticity and Poisson’s ratio of concrete specimens.
AS 1012.3.5. (2015). Methods of testing concrete: determination of properties related to the
consistency of concrete - slump flow, T500 and J-ring test.
AS 1478.1. (2000). Chemical admixtures for concrete, mortar and grout, Part 1: Admixtures for
concrete.
AS 3600. (2009). Concrete structures.
Ashraf, M., Naeem Khan, A., Ali, Q., Mirza, J., Goyal, A., and Anwar, A. M. (2009).
“Physico-chemical, morphological and thermal analysis for the combined pozzolanic
activities of minerals additives.” Construction and Building Materials, 23(6), 2207–2213.
ASTM C150 / C150M-19a. (2019). Standard specification for portland cement. West
Conshohocken, PA.
ASTM C618-19. (2019). Standard specification for coal fly ash and raw or calcined natural pozzolan
for use in concrete. West Conshohocken, PA.
ASTM C876-15. (2015). Standard Test Method for Corrosion Potentials of Uncoated Reinforcing
Steel in Concrete. West Conshohocken, PA.
ASTMC109 / C109M-16a. (2016). Standard test method for compressive strength of hydraulic
cement mortars (using 2-in. or [50-mm] cube specimens). West Conshohocken, PA.
Berry, E. E., Hemmings, R. T., and Zhang, M. H. (1994). “Hydration in high-volume fly
ash concrete binders.” ACI Materials Journal, 91(4), 382–389.
Bonavetti, V., Donza, H., Rahhal, V., and Irassar, E. (2000). “Influence of initial curing on
the properties of concrete containing limestone blended cement.” Cement and Concrete
Research, 30(5), 703–708.
Chen, T., Gao, X., and Ren, M. (2018). “Effects of autoclave curing and fly ash on
mechanical properties of ultra-high performance concrete.” Construction and Building
Materials, 158, 864–872.
Chindaprasirt, P., and Rukzon, S. (2008). “Strength, porosity and corrosion resistance of
ternary blend Portland cement, rice husk ash and fly ash mortar.” Construction and
Building Materials, 22(8), 1601–1606.
Courtial, M., De Noirfontaine, M. N., Dunstetter, F., Signes-Frehel, M., Mounanga, P.,
Cherkaoui, K., and Khelidj, A. (2013). “Effect of polycarboxylate and crushed quartz in
UHPC: microstructural investigation.” Construction and Building Materials, 44, 699–705.
Cwirzen, A., Penttala, V., and Vornanen, C. (2008). “Reactive powder based concretes:
Mechanical properties, durability and hybrid use with OPC.” Cement and Concrete
Research, 38(10), 1217–1226.
Damineli, B. L., Kemeid, F. M., Aguiar, P. S., and John, V. M. (2010). “Measuring the eco-
efficiency of cement use.” Cement and Concrete Composites, 32(8), 555–562.
Dhundasi, A. A., and Khadiranaikar, R. B. (2019). Effect of curing conditions on mechanical
properties of reactive powder concrete with different dosage of quartz powder. Lecture Notes in Civil
Engineering.
Elchalakani, M., Basarir, H., and Karrech, A. (2017). “Green concrete with high-volume fly
ash and slag with recycled aggregate and recycled water to build future sustainable

115
Chapter 4

cities.” Journal of Materials in Civil Engineering, 29(2), 04016219.


Elchalakani, M., Dong, M., Karrech, A., Li, G., Mohamed Ali, M. S., Xie, T., and Yang,
B. (2018). “Development of Fly Ash- and Slag-Based Geopolymer Concrete with
Calcium Carbonate or Microsilica.” Journal of Materials in Civil Engineering, 30(12),
04018325.
EN 13295. (2004). Products and systems for the protection and repair of concrete structures - test
methods - determination of resistance to carbonation.
Ferdosian, I., Camões, A., and Ribeiro, M. (2017). “High-volume fly ash paste for
developing ultra-high performance concrete (UHPC).” Ciencia e Tecnologia dos Materiais,
29(1), e157–e161.
French, W. J. (1991). “Concrete petrography: A review.” Quarterly Journal of Engineering
Geology, 24(1), 17–48.
Gesoglu, M., Güneyisi, E., Asaad, D. S., and Muhyaddin, G. F. (2016a). “Properties of low
binder ultra-high performance cementitious composites: Comparison of nanosilica and
microsilica.” Construction and Building Materials, 102, 706–713.
Gesoglu, M., Güneyisi, E., Muhyaddin, G. F., and Asaad, D. S. (2016b). “Strain hardening
ultra-high performance fiber reinforced cementitious composites: Effect of fiber type and
concentration.” Composites Part B: Engineering, 103, 74–83.
Ghafari, E., Costa, H., Júlio, E., Portugal, A., and Durães, L. (2012). “Optimization of
UHPC by adding nanomaterials.” 3rd International Symposium on UHPC and
Nanotechnology for High Performance Construction Materials, 71–78.
Ghafor, K., Qadir, S., Mahmood, W., and Mohammed, A. (2020). “Statistical variations
and new correlation models to predict the mechanical behaviour of the cement mortar
modified with silica fume.” Geomechanics and Geoengineering.
Gholampour, A., and Ozbakkaloglu, T. (2017). “Performance of sustainable concretes
containing very high volume Class-F fly ash and ground granulated blast furnace slag.”
Journal of Cleaner Production, 162, 1407–1417.
Goaiz, H. A., Yu, T., and Hadi, M. N. S. (2018). “Quality Evaluation Tests for Tensile
Strength of Reactive Powder Concrete.” Journal of Materials in Civil Engineering, 30(5),
04018070.
Graybeal, B. A. (2007). “Compressive behavior of ultra-high-performance fiber-reinforced
concrete.” ACI Materials Journal, 104(2), 146–152.
Gupta, S. (2016). “Effect of content and fineness of slag as high volume cement replacement
on strength and durability of ultra-high performance mortar.” J. Build. Mater. Struct., 3,
43–54.
Han, B., Wang, Z., Zeng, S., Zhou, D., Yu, X., Cui, X., and Ou, J. (2017). “Properties and
modification mechanisms of nano-zirconia filled reactive powder concrete.” Construction
and Building Materials, 141, 426–434.
Hannawi, K., Bian, H., Prince-Agbodjan, W., and Raghavan, B. (2016). “Effect of different
types of fibers on the microstructure and the mechanical behavior of Ultra-High
Performance Fiber-Reinforced Concretes.” Composites Part B: Engineering, 86, 214–220.
Holschemacher, K., Weisse, D., and Klotz, S. (2004). “Bond of Reinforcement in Ultra
High Strength Concrete.” Proceedings of the International Symposium on Ultra High
Performance Concrete, 375–387.
Horszczaruk, E., Mijowska, E., Kalenczuk, R. J., Aleksandrzak, M., and Mijowska, S.
(2015). “Nanocomposite of cement/graphene oxide - Impact on hydration kinetics and
Young’s modulus.” Construction and Building Materials, 78, 234–242.

116
Chapter 4

Huang, C. H., Lin, S. K., Chang, C. S., and Chen, H. J. (2013). “Mix proportions and
mechanical properties of concrete containing very high-volume of Class F fly ash.”
Construction and Building Materials, 46, 71–78.
Huang, H., Gao, X., and Jia, D. (2019). “Effects of rheological performance, antifoaming
admixture, and mixing procedure on air bubbles and strength of UHPC.” Journal of
Materials in Civil Engineering, 31(4), 04019016.
Huang, H., Gao, X., Wang, H., and Ye, H. (2017). “Influence of rice husk ash on strength
and permeability of ultra-high performance concrete.” Construction and Building
Materials, 149, 621–628.
IS 3495 (Part 2). (1992). Methods of tests of burnt clay building bricks, Part 2: determination of water
absorption. New Delhi.
Jaafar, M. F. M., Saman, H. M., Ariffin, N. F., Muthusamy, K., Ahmad, S. W., and Ismail,
N. (2018). “Corrosion monitoring on steel reinforced nano metaclayed-UHPC towards
strain modulation using fiber Bragg grating sensor.” IOP Conference Series: Materials
Science and Engineering, 431, 122006.
Karihaloo, B. L., and Vriese, K. M. B. D. (1999). “Short-Fibre Reinforced Reactive Powder
Concrete.” Third International RILEM Workshop on High Performance Fiber Reinforced
Cement Composites, RILEM Publications SARL, 53–63.
Kayali, O., and Sharfuddin Ahmed, M. (2013). “Assessment of high volume replacement
fly ash concrete-concept of performance index.” Construction and Building Materials, 39,
71–76.
Kayali, O., and Zhu, B. (2005). “Corrosion performance of medium-strength and silica fume
high-strength reinforced concrete in a chloride solution.” Cement and Concrete Composites,
27(1), 117–124.
Korpa, A., Kowald, T., and Trettin, R. (2009). “Phase development in normal and ultra
high performance cementitious systems by quantitative X-ray analysis and
thermoanalytical methods.” Cement and Concrete Research, 39(2), 69–76.
Kuder, K., Lehman, D., Berman, J., Hannesson, G., and Shogren, R. (2012). “Mechanical
properties of self consolidating concrete blended with high volumes of fly ash and slag.”
Construction and Building Materials, 34, 285–295.
Kwan, A. K. H., and Li, Y. (2013). “Effects of fly ash microsphere on rheology,
adhesiveness and strength of mortar.” Construction and Building Materials, 42, 137–145.
Larbi, J. A., and Bijen, J. M. J. M. (1990). “Effects of water-cement ratio, quantity and
fineness of sand on the evolution of lime in set portland cement systems.” Cement and
Concrete Research, 20(5), 783–794.
Li, W., Huang, Z., Cao, F., Sun, Z., and Shah, S. P. (2015). “Effects of nano-silica and
nano-limestone on flowability and mechanical properties of ultra-high-performance
concrete matrix.” Construction and Building Materials, 95, 366–374.
Liu, K., Yu, R., Shui, Z., Li, X., Ling, X., He, W., Yi, S., and Wu, S. (2018). “Effects of
Pumice-Based Porous Material on Hydration Characteristics and Persistent Shrinkage
of Ultra-High Performance Concrete (UHPC).” Materials, 12(1), 11.
Long, G. C., Xie, Y. J., Wang, P. M., and Jiang, Z. W. (2005). “Properties and
micro/mecrostructure of reactive powder concrete.” J. Chin. Ceram. Soc., 4, 456–461.
Ma, J., Orgass, M., Dehn, F., Schmidt, D., and Tue, N. V. (2004). “Comparative
Investigations on Ultra-High Performance Concrete with or without Coarse
Aggregates.” Proceedings of the international symposium on ultra high performance concrete,
205–212.
Máca, P., Zatloukal, J., and Konvalinka, P. (2012). “Development of Ultra High

117
Chapter 4

Performance Fiber Reinforced Concrete mixture.” ISBEIA 2012 - IEEE Symposium on


Business, Engineering and Industrial Applications, 861–866.
Mao, X., Qu, W., and Zhu, P. (2019). “Mixture Optimization of Green Reactive Powder
Concrete with Recycled Powder.” Journal of Materials in Civil Engineering, 31(5),
04019033.
Meek, A. H., Beckett, C. T. S., Carsana, M., and Ciancio, D. (2018). “Corrosion protection
of steel embedded in cement-stabilised rammed earth.” Construction and Building
Materials, 187, 942–953.
Mounanga, P., Cherkaoui, K., Khelidj, A., Courtial, M., De Noirfontaine, M. N., and
Dunstetter, F. (2012). “Extrudable reactive powder concretes: Hydration, shrinkage and
transfer properties.” European Journal of Environmental and Civil Engineering, 16(SUPPL.
1).
Mueller, U., Williams Portal, N., Chozas, V., Flansbjer, M., Larazza, I., da Silva, N., and
Malaga, K. (2016). “Reactive powder concrete for façade elements – A sustainable
approach.” Journal of Facade Design and Engineering, 4(1–2), 53–66.
Naaman, A. E. (2011). “Half a century of progress leading to ultra-high performance fiber
reinforced concrete: Part 1-overall review.” 2nd Int. RILEM Conference on Strain Hardening
Cementitious Composites, Rio de Janeiro, Brazil, 12–14.
Ng, S., Sandberg, L. I. C., and Jelle, B. P. (2014). “Insulating and Strength Properties of an
Aerogel-Incorporated Mortar Based an UHPC Formulations.” Key Engineering Materials,
629–630, 43–48.
Peng, Y., Hu, S., and Ding, Q. (2010). “Preparation of reactive powder concrete using fly
ash and steel slag powder.” Journal Wuhan University of Technology, Materials Science
Edition, 25(2), 349–354.
Richard, P., and Cheyrezy, M. (1995). “Composition of reactive powder concretes.” Cement
and Concrete Research, 25(7), 1501–1511.
Rong, Z., Sun, W., and Zhang, Y. (2010). “Dynamic compression behavior of ultra-high
performance cement based composites.” International Journal of Impact Engineering, 37(5),
515–520.
Ruan, Y., Han, B., Yu, X., Li, Z., Wang, J., Dong, S., and Ou, J. (2018). “Mechanical
behaviors of nano-zirconia reinforced reactive powder concrete under compression and
flexure.” Construction and Building Materials, 162, 663–673.
Sbia, L. A., Peyvandi, A., Soroushian, P., Balachandra, A. M., and Sobolev, K. (2015).
“Evaluation of modified-graphite nanomaterials in concrete nanocomposite based on
packing density principles.” Construction and Building Materials, 76(1), 413–422.
Scheydt, J., and Müller, H. (2012). “Microstructure of ultra high performance concrete
(UHPC) and its impact on durability.” 3rd International Symposium on on Ultra High
Performance Concrete, 349–356.
Schwartzentruber, A., Philippe, M., and Marchese, G. (2004). “Effect of PVA, glass and
metallic fibers, and of an expansive admixture on the cracking tendency of ultrahigh
strength mortar.” Cement and Concrete Composites, 26(5), 573–580.
Shen, P., Lu, L., He, Y., Wang, F., and Hu, S. (2019). “The effect of curing regimes on the
mechanical properties, nano-mechanical properties and microstructure of ultra-high
performance concrete.” Cement and Concrete Research, 118, 1–13.
Shi, H. sheng, Xu, B. wan, and Zhou, X. chen. (2009). “Influence of mineral admixtures on
compressive strength, gas permeability and carbonation of high performance concrete.”
Construction and Building Materials, 23(5), 1980–1985.
Soliman, N. A., and Tagnit-Hamou, A. (2017). “Using glass sand as an alternative for quartz

118
Chapter 4

sand in UHPC.” Construction and Building Materials, 145, 243–252.


Song, Q., Yu, R., Shui, Z., Wang, X., Rao, S., and Lin, Z. (2018). “Optimization of fibre
orientation and distribution for a sustainable Ultra-High Performance Fibre Reinforced
Concrete (UHPFRC): Experiments and mechanism analysis.” Construction and Building
Materials, 169, 8–19.
Soutsos, M. N., Millard, S. G., and Karaiskos, K. (2005). “Mix design, mechanical
properties, and impact resistance of reactive powder concrete (RPC).” International
RILEM Workshop on High Performance Fiber Reinforced Cementitious Composites (HPFRCC)
in Structural Applications, 549–560.
Sovják, R., Vavřiník, T., Máca, P., Zatloukal, J., Konvalinka, P., and Song, Y. (2013).
“Experimental investigation of ultra-high performance fiber reinforced concrete slabs
subjected to deformable projectile impact.” Procedia Engineering, 65, 120–125.
Sujjavanich, S., Sida, V., and Suwanvitaya, P. (2005). “Chloride permeability and corrosion
risk of high-volume fly ash concrete with mid-range water reducer.” ACI Materials
Journal, 102(3), 177–182.
Tabet, W. E., Cerato, A. B., Madden, A. S. E., and Jentoft, R. E. (2018). “Characterization
of hydration products’ formation and strength development in cement-stabilized
kaolinite using TG and XRD.” Journal of Materials in Civil Engineering, 30(10).
Tafraoui, A., Escadeillas, G., Lebaili, S., and Vidal, T. (2009). “Metakaolin in the
formulation of UHPC.” Construction and Building Materials, 23(2), 669–674.
Temuujin, J., Rickard, W., and Van Riessen, A. (2013). “Characterization of various fly
ashes for preparation of geopolymers with advanced applications.” Advanced Powder
Technology, 24(2), 495–498.
Wang, D., Shi, C., Wu, Z., Wu, L., Xiang, S., and Pan, X. (2016). “Effects of nanomaterials
on hardening of cement–silica fume–fly ash-based ultra-high-strength concrete.”
Advances in Cement Research, 28(9), 555–566.
Wang, X. Y., and Park, K. B. (2016). “Analysis of compressive strength development and
carbonation depth of high-volume fly ash cement pastes.” ACI Materials Journal, 113(2),
151–161.
Wu, Z., Shi, C., He, W., and Wang, D. (2016a). “Uniaxial Compression Behavior of Ultra-
High Performance Concrete with Hybrid Steel Fiber.” Journal of Materials in Civil
Engineering, 28(12), 06016017.
Wu, Z., Shi, C., and Khayat, K. H. (2016b). “Influence of silica fume content on
microstructure development and bond to steel fiber in ultra-high strength cement-based
materials (UHSC).” Cement and Concrete Composites, 71, 97–109.
Wu, Z., Shi, C., Khayat, K. H., and Xie, L. (2018). “Effect of SCM and nano-particles on
static and dynamic mechanical properties of UHPC.” Construction and Building Materials,
182, 118–125.
Xie, T., and Visintin, P. (2018). “A unified approach for mix design of concrete containing
supplementary cementitious materials based on reactivity moduli.” Journal of Cleaner
Production, 203, 68–82.
Xie, T. Y., Elchalakani, M., Mohamed Ali, M. S., Dong, M. H., Karrech, A., and Li, G.
(2018). “Mechanical and durability properties of self-compacting concrete with blended
binders.” Computers and Concrete, 22(4), 407–417.
Yasin, A. K., Bayuaji, R., and Susanto, T. E. (2017). “A review in high early strength
concrete and local materials potential.” IOP Conference Series: Materials Science and
Engineering, 267(1), 1–9.
Yodsudjai, W., and Pattarakittam, T. (2017). “Factors influencing half-cell potential

119
Chapter 4

measurement and its relationship with corrosion level.” Measurement: Journal of the
International Measurement Confederation, 104, 159–168.
Yoo, D. Y., and Yoon, Y. S. (2015). “Structural performance of ultra-high-performance
concrete beams with different steel fibers.” Engineering Structures, 102, 409–423.
Yu, R., Spiesz, P., and Brouwers, H. J. H. (2014). “Effect of nano-silica on the hydration
and microstructure development of Ultra-High Performance Concrete (UHPC) with a
low binder amount.” Construction and Building Materials, 65, 140–150.
Yu, R., Spiesz, P., and Brouwers, H. J. H. (2015a). “Development of an eco-friendly Ultra-
High Performance Concrete (UHPC) with efficient cement and mineral admixtures
uses.” Cement and Concrete Composites, 55, 383–394.
Yu, R., Spiesz, P., and Brouwers, H. J. H. (2015b). “Development of Ultra-High
Performance Fibre Reinforced Concrete (UHPFRC): Towards an efficient utilization of
binders and fibres.” Construction and Building Materials, 79, 273–282.
Yunsheng, Z., Wei, S., Sifeng, L., Chujie, J., and Jianzhong, L. (2008). “Preparation of
C200 green reactive powder concrete and its static-dynamic behaviors.” Cement and
Concrete Composites, 30(9), 831–838.
Zhao, S., and Sun, W. (2014). “Nano-mechanical behavior of a green ultra-high
performance concrete.” Construction and Building Materials, 63, 150–160.
Zhu, P., Mao, X., Qu, W., Li, Z., and Ma, Z. J. (2016). “Investigation of using recycled
powder from waste of clay bricks and cement solids in reactive powder concrete.”
Construction and Building Materials, 113, 246–254.
Zhu, Z., Li, B., and Zhou, M. (2015). “The influences of iron ore tailings as fine aggregate
on the strength of ultra-high performance concrete.” Advances in Materials Science and
Engineering, 2015, 412878.
Zou, Z. H., Wu, J., Wang, Z., and Wang, Z. (2016). “Relationship between half-cell
potential and corrosion level of rebar in concrete.” Corrosion Engineering Science and
Technology, 51(8), 588–595.

120
CHAPTER
5
Development of ECO-UHPC utilizing gold
mine tailings as quartz sand alternative

Tanvir Ahmeda, Mohamed Elchalakania, Hakan Basarirb, Ali Karrecha, Ehsan Sadrossadata, Bo
Yangc,d,*
a
Department of Civil, Environmental and Mining Engineering, Faculty of Engineering, Computing and Mathematical
Sciences, The University of Western Australia, 35 Stirling Highway, Crawley, WA 6009, Australia
b
Department of Geoscience and Petroleum, Norwegian University of Science and Technology, NO-7491 Trondheim,
Norway
c
Key Laboratory of New Technology for Construction of Cities in Mountain Area, Chongqing University, Chongqing
400045, China
d
School of Civil Engineering, Chongqing University, Chongqing 400045, China

*Corresponding author: yang0206@cqu.edu.cn

Graphical abstract

Cleaner Engineering and Technology, 4, 100176


(https://doi.org/10.1016/j.clet.2021.100176)
Chapter 5

Abstract

To ensure ultra-high performance, in terms strength and durability, coarse aggregate is


typically avoided in UHPC (ultra-high performance concrete). Instead, very fine quartz sand
is usually used as the only aggregate. However, excessive extraction of sand from natural
resources and its grinding and refining processes to prepare very fine quartz-rich sand are
not economically or environmentally lucrative. Limited number of studies sought to address
this concern, and very few of these studies investigated the use of mine tailings (quartz based
tailings and iron ore tailings) in UHPC as sand alternatives. In the present study, the
possibility of utilizing gold mine tailings, sourced from a gold mine in Western Australia
(WA), as conventional quartz sand substitute in UHPC has been investigated. Results
suggest that UHPCs, made with up to 80% replacement of quartz sand by the tailings,
exhibit compressive strengths comparable to or higher than that of the UHPC with 100%
quartz sand. 28-day strength greater than 120 MPa is achievable up to 100% replacement.
The water absorption and the initial rate of absorption values of UHPCs with tailings are
generally lower than those of the UHPC without tailings. The leachability of toxic metals
from UHPC with up to 100% tailings content lies well below the regulatory thresholds. The
combined material and transportation cost of UHPC can be reduced by up to 33.1%
replacing quartz sand by the tailings, for construction near the mine site. The CO2 emission
can be reduced by up to 12.1%. In the area near the mine site, utilization of the tailings can
economically and environmentally, as well as in terms of durability, be a better option than
quartz sand for construction works that require UHPC with 28-strength in excess of 120
MPa.

Keywords: CO2, cost, durability, mine waste, sustainability, ultra-high performance


concrete

5.1 Introduction

Compared to normal-strength concrete (NSC) or high-strength concrete (HSC), ultra-high


performance concrete (UHPC) exhibits extraordinary mechanical and durability
performances. NSC exhibits compressive strength in the range of 20–40 MPa. HSC can
achieve strength greater than 40 MPa but seldom greater than 100 MPa. Whereas UHPC,
according to a group of researchers, must exhibit a compressive strength greater than 120
MPa (Huang et al. 2019), or greater than 150 MPa according to others (Courtial et al. 2013).
The water absorption of NSC is much higher than 3.5% (Neville 2011), and the absorption
of HSC is usually in the range of 1.5%–3% (Nematollahi et al. 2011). The water absorption
of UHPC, on the other hand, is typically less than 1.5% (Scheydt and Müller 2012).

122
Chapter 5

Although UHPC exhibits outstanding mechanical and durability performances, the material
cost of UHPC is usually high compared with NSC or HSC. Total exclusion of coarse
aggregate and use of only very fine high-quality quartz sand (maximum size ≤ 600µm, quartz
content ≥ 95%) is one of the factors contributing to its high cost. This is because the
extraction, grinding, and refining processes of such fine and high-quality aggregate are
energy-intensive and expensive. Extraction and grinding of quartz sand are also associated
with several negative environmental consequences. For instance, CO2 emission,
deforestation, loss of biodiversity, soil degradation, lowering of water table, particulate
matter air pollution, etc. (Gavriletea 2017). According to the United Nations Environment
Programme (UNEP), the global sand extraction rate is currently substantially higher than
its natural replenishment rate, as sand and gravel are the second most consumed natural
resources after water (UNEP 2014). The global consumption of sand and gravel is
approximately 32–50 billion tonnes per year (Bendixen et al. 2019). A number of studies
were carried out to investigate the influences of different by-product or waste based
alternatives to conventional quartz sand on the properties of UHPC. Jiao et al. (2020) used
waste-glass sand to produce UHPC. The 28-day compressive strength of the UHPC with
waste-glass sand was observed to increase by around 25% with respect that of the UHPC
with quartz sand. Yang et al. (2020) fully substituted quartz sand by recycled rock dust to
produce UHPC and observed an increase of 9.5% in the compressive strength. Zhang et al.
(2018) used recycled fine aggregate from demolished concrete in UHPC and the strength
was found to decrease by 13.3% with respect to that of the control.

In addition to the foregoing types of alternative aggregate, mine tailings (by-products of


mineral extraction processes) can potentially be used as aggregates for UHPC production.
The mining industry was estimated to produce around 14 billion tonnes of tailings annually,
and the amount is increasing because of the increasing use of low-grade ores to meet the
global demand for valuable metals (Kinnunen et al. 2018). Mine tailings are usually very
fine (especially the ones from mines of precious minerals like gold), due to the grinding of
ore into fine particles to extract the target metal (Dong et al. 2019). Because of this, quartz
based mine tailings have the potential to be utilized as aggregates and in some cases as
supplementary binders in UHPC without the requirement of further grinding. Mine tailings
often contain toxic heavy metals. The dense micro-structure of UHPC is, however, supposed
to restrict the heavy metals from being leached out of the concrete (Pyo et al. 2018). The use
of tailings in UHPC may also address a number of concerns. For instance, the hazards
associated with the mismanagement of tailings can be reduced (e.g., workers’ safety risk,
chemical and toxic metal pollution, etc.), CO2 emission associated with the extraction and

123
Chapter 5

grinding of quartz sand can be decreased, natural sources of quartz sand can be saved from
being depleted to some extent, etc. (Ince 2019).

As can be seen from Table 5.1, a number of studies were conducted to investigate the
influence of utilizing different types of tailings, as sand substitutes, on the properties of NSC
and HSC. Few studies can be found in the literature which investigated the possibility of
utilizing mine tailings to substitute sand in UHPC. Zhao et al. (2014) and Zhu et al. (2015)
investigated the influence of replacing sand by iron ore tailings on the properties of UHPC.
The researchers concluded that full substitution of sand by iron ore tailings reduces the
strength of UHPC. However, Zhao et al. (2014) observed that up to 20% replacement, the
28-day strength of UHPC remained comparable to that of the UHPC without iron ore
tailings. Pyo et al. (2018) used a type of quartz-based mine tailings to partially replace (30%
by mass) quartz sand in UHPC, and obtained strength comparable to that of the UHPC with
100% quartz sand. Research addressing the influence of utilizing gold mine tailings, as
quartz sand alternative, on the properties of UHPC is scarce in the literature.

In the present study, the possibility of using a type of gold mine tailings as aggregate in
UHPC was investigated. UHPC mixes were prepared, replacing quartz sand by the tailings
at different percentages (up to 100% by mass). Mechanical properties such as compressive
strength, elastic modulus, splitting tensile strength, and durability properties such as water
absorption, initial rate of water absorption, carbonation resistance were investigated for the
prepared mixes. Leaching toxicity test was conducted for each UHPC mix, to assess the
leachability of the heavy metals present in the tailings from the UHPC. Scanning electron
microscopy (SEM) was performed for selected UHPC mixes to understand their macro-
behaviours. Economic and ecological influences of using the tailings in UHPC, as
replacement of conventional quartz sand, were also examined.

5.2 Materials and methods

5.2.1 Materials

The gold mine tailings used in this study were sourced from a gold mine site in Western
Australia (WA). Due to a confidentiality agreement, the exact location of the mine is not
mentioned in this chapter. The mine will be termed as Mine 1 in the following discussion.

124
Chapter 5

Table 5.1. A review of mechanical properties of standard-cured† structural concretes/ mortars incorporating different types of tailings as sand alternatives
Concrete type Reference Tailings Material replaced Replacement, % w/b Replacement at fc,28-d, MPa
Location Type maximum fc,28-d, % Maximum At 0% replacement (fc0,28-d) At maximum replacement
UHPC Pyo et al. (2018) Sang-dong mine, South Quartz based tailings Quartz sand 0, 15, 30 0.17 15, 30 133* 132* 133*
(fc0,28-d ≥ 120 MPa) Korea
Zhu et al. (2015) China Iron ore tailings Quartz sand 0, 100 0.16 0 189Γ 189Γ 183Γ
River sand 0, 100 100 183Γ 172Γ 183Γ
Zhao et al. (2014) China Iron ore tailings River sand 0, 10, 20, 30, 40, 50, 100 0.16- 20 139.2 138 97.2
0.20
VHSC/ VHSM Zhang et al. (2020a) Liaoning, China Iron ore tailings Manufactured 0, 20, 40, 60, 80, 100 0.18 40 120 105 95
(100 MPa ≤ fc0,28-d < 120 sand
MPa)
HSC/ HSM Gao et al. (2020) Shangluo, Shaanxi, China Molybdenum tailings Sand 0, 25, 50, 100 0.42 0 55 55 48
(40 MPa ≤ fc0,28-d < 100 0.31 0 61.5 61.5 55.5
MPa) Li et al. (2020) Yaogou pond, Shaanxi, Iron ore tailings River sand 0, 10, 20, 30, 40, 50, 70, 100 0.40 30 52.2 42 37.5
China
Zhang et al. (2020b) Daye, Hubei, China Copper tailings Manufactured 0, 20 0.32 0 59 59 57
sand
Zhang et al. (2020c) Shangluo, Shaanxi, China Iron ore tailings River sand 0, 100 0.35 100 62.4 60.4 62.4
Xu et al. (2018) China Kaolin tailings River sand 0, 20, 40, 60, 80, 100 0.50 60 58.6 57.5 52.8
NSC/ NSM Protasio et al. (2021) Germano dam, Minas Iron ore tailings Quartz sand 0, 10, 20, 30 0.55 10 40.9 34.5 39.8
(20 MPa ≤ fc0,28-d < 40 Gerais, Brazil
MPa) Liu et al. (2020a) China Graphite tailings River sand 0, 10, 20, 30, 40, 50, 60, 70, 0.52 10 42.2 36.9 28
80, 90, 100
Fisonga (2019) Mindolo dam, Copperbelt, Copper tailings Sand 0, 30, 50, 70, 100 0.48- 30 35.6 33.8 25
Zambia 0.53
Ince (2019) Lefke-Xeros, Cyprus Gold mine tailings Sand 0, 10, 20, 30 0.62 30 42 32.5 41.5
Kathirvel et al. (2018) Tamil Nadu, India Graphite ore tailings River sand 0, 10, 20, 30, 40, 50, 60, 70, 0.50 10 33.2 32.5 14.5
80, 90, 100 0.55 10 27 26.3 8
Treated graphite ore River sand 0, 10, 20, 30, 40, 50, 60, 70, 0.50 10 34.5 32.5 20.5
tailings 80, 90, 100 0.55 10 28 26.3 14.2
Gupta et al. (2017) Khetri, Rajasthan, India Copper tailings River sand 0, 10, 20, 30, 40, 50, 60, 70 0.48- 20 31.9 30.9 23.6
0.50
Shettima et al. (2016) Johor, Malaysia Iron ore tailings River sand 0, 25, 50, 75, 100 0.50 25 42.9 38.0 38.5
Tian et al. (2016) Zibo, China Iron ore tailings River sand 0, 25, 35, 45 0.55 35 39.5 38.7 36.8
Thomas et al. (2013) Khetri, Rajasthan, India Copper tailings River sand 0, 10, 20, 30, 40, 50, 60 0.40 10 41.5 35 36.5
0.45 20 39 37.5 33
0.50 30 37 31.5 32.5
UHPC = ultra-high performance concrete/ composite, VHSC/ VHSM = very-high-strength concrete/ mortar, HSC/ HSM = high-strength concrete/ mortar, NSC/ NSM = normal-strength concrete/ mortar, fc0,28-d = 28-day compressive strength of control mix (mix with 0% tailings), fc,28-d =
28-day compressive strength, w/b = water to binder ratio

Cured in water or in fog room at a relative humidity of 95% ± 5%, and at a temperature of 24 °C ± 6 °C
*Air cured
ΓCured in hot water at 100 °C on second and third days

125
Chapter 5

Mining operation in Mine 1 consists of open pit and underground mining. The ore
processing plant includes two process trains. Process routes include crushing, milling,
flotation, and cyanide leaching. Flotation tailings are thickened. The cyanide in the tailings
is recovered and recycled, and a WAD (weak acid dissociable) cyanide level of less than 50
ppm is maintained before discharging the tailings into the tailings storage facility. Further
degradation of the remaining cyanide is facilitated through sunlight exposure (Logsdon et
al. 1999). For this study, two tonnes of tailings exposed to sun-drying for more than three
years were harvested from an old storage area. In the following discussion, the tailings will
be termed as WAGT1 (Western Australian Gold Mine Tailings sourced from Mine 1). The
composition of major elements in WAGT1, in terms of oxides, are presented in Table 5.2.
Table 5.2 also shows the concentrations of several heavy elements (e.g., As, Pb, Zn, Cd, Cr,
and Cu) in WAGT1. The chemical properties presented in Table 5.2 are based on elemental
concentrations measured using inductively coupled plasma - optical emission spectrometry
(ICP-OES).

Table 5.2. Chemical and physical properties of solid materials

Material WAGT1 Sand Cement Silica fume Fly ash


Major elements Al2O3 9.36 0.017 3.18 0.13 23.93
(mass% of CaO 2.48 0.002 64.48 0.33 6.98
oxide)
Fe2O3 4.94 0.008 4.34 0.07 7.94
K2O 1.62 0.003 0.35 0.54 1.02
MgO 1.23 0.003 1.46 0.71 1.27
MnO 0.079 0.001 0.120 0.008 0.104
Na2O 2.22 0.003 0.14 0.29 0.38
P2O5 0.058 0.00 0.228 0.129 0.535
SiO2 71.2 99.88 21.4 94.6 55.9
a
SO3 1.89 0.00 2.93 0.07 0.30
TiO2 0.494 0.025 0.192 0.005 1.312
Loss on 4.06 0.05 1.15 3.1 1.71
ignition
Heavy elements Zn 151
in WAGT1 Cd 0.9
(mg/kg)
As 232
Pb 12
Cu 397
d50 (µm) 141.4 237.3 17.3 1.7 9.8
b
Specific gravity 2.7 2.65 3.18 2.15 2.38
d50 = median diameter
a
Sulphur (S) content in WAGT1 = 0.76%
b
As per manufacturer

126
Chapter 5

The cyanide concentration in the as-received WAGT1 was not measured in the current
study. However, the mining company, in charge of WAGT1 management in Mine 1, is a
signatory to the International Cyanide Management Code; and it was confirmed by the
company that the cyanide concentration in the as-received WAGT1 was well below the
hazard level. Detailed investigation on the possibility of acid generation from the sulphide
minerals in WAGT1 was not conducted in the present study. However, pH was measured
for a mixture of 1 L of deionized water and 60 g of as-received WAGT1. The mixture was
stirred for half an hour before measuring the pH. The pH of the mixture was found to be
9.37. The pH value indicates that the as-received WAGT1 were not acidic.

Figure 5.1 shows the scanning electron micrograph of WAGT1. Figure 5.1 also presents the
X-ray diffraction (XRD) pattern of WAGT1 in the 2θ range of 5–70°, recorded using a
diffractometer with Cu K radiation (λ = 1.5404 Å), 40 kV, and 40 mA at a rate of 2°/min.
Quartz was found to be the main crystalline phase from the XRD pattern of WAGT1. Peaks
corresponding to several other minerals like muscovite, albite, halite, and polyhalite were
also observed in their XRD pattern. It is worth noting that these minerals are softer than
quartz. On Mohs scale, quartz’s hardness is 7. Mohs hardness is in the range of 2–2.5 for
muscovite, 6–6.5 for albite, 2–2.5 for halite, and 2.5–3.5 for polyhalite.

Figure 5.1. XRD pattern and scanning electron micrograph of WAGT1; [M = muscovite
(KAl3Si3O10(F,OH)2), A = albite (NaAlSi3O8), P = polyhalite (Ca2H4K2MgO18S4), H = halite
(NaCl), Q = quartz (SiO2)]

127
Chapter 5

Figure 5.2. Particle size distributions of solid materials

The cement used in this study conforms to the requirements of ASTM C150 / C150M-16
Type I Portland cement (ASTM C150 / C150M-16 2016), silica fume conforms to AS/NZS
3582.3 amorphous silica (AS/NZS 3582.3 2016), and fly ash conforms to ASTM C618-19
Class-F fly ash (ASTM C618-19 2019). The quartz sand had a median diameter (d50) of 237.3
µm and an absorption capacity of 0.38%. The chemical and physical properties of the
cement, silica fume, fly ash, and quartz sand are presented in Table 5.2. It can be seen from
Table 5.2 that the SiO2 content of the quartz sand was 99.8% and the SiO2 content of
WAGT1 was 71.2%. This indicates that the quartz sand had higher quartz content than that
in WAGT1. Figure 5.2 shows the particle size distributions (PSD) of all the solid materials.
A polycarboxylate based high-range water reducer (HRWR) with 36% solid content and a
defoaming agent with 5.5% solid content were used as chemical admixtures.

5.2.2 Mix proportions

The mix proportions of the prepared UHPC mixes are summarized in Table 5.3. Quartz
sand was used as the aggregate in the control mix T0. 20% of the quartz sand was replaced
by WAGT1 (by mass) to prepare T20, 40% was replaced to prepare T40, 60% was replaced
to prepare T60, and 100% was replaced to prepare T100. The water to binder ratio (w/b) of
the mixes was kept constant at 0.182.

The combined PSD of the solid constituents of each mix is shown in Figure 5.3. The target
combined PSD [modified Andreasen and Andersen (A&A) model (Equation 5.1)] is also
shown in Figure 5.3. The value of q in Equation 5.1 was selected to be 0.22, as suggested by
previous studies for mixes with high powder content (<250 µm) (Kim et al. 2016).

128
Chapter 5

𝐷 𝑞 − 𝐷𝑚𝑖𝑛 𝑞
𝑝(𝐷) = (5.1)
𝐷𝑚𝑎𝑥 𝑞 − 𝐷𝑚𝑖𝑛 𝑞

D represents particle size (µm), p(D) is cumulative percent finer than size D, Dmin is the
minimum size of particle (µm), and Dmax is the maximum size of particle (µm).

The root-mean-square error (RMSE) and determination coefficient (R2) between the target
PSD curve and each combined PSD curve are presented in Figure 5.3. It can be seen that
the RMSEs between the combined PSDs of the mixes with WAGT1 and the target PSD
were lower than that between the combined PSD of T0 and the target PSD. This indicates
better particle packing was achieved when WAGT1 were incorporated. The lowest RMSE
was observed for T60.

Table 5.3. Mix proportions of UHPC mixes

Mix Constituents (kg/m3) w/b


ID. Cement
Fly Silica Sand WAGT1 Water HRWR* Defoaming
ash fume agent*
T0 794 140 233 969 - 212 14.7 1.3 0.182
T20 794 140 233 775 194 212 14.7 1.3 0.182
T40 794 140 233 581 388 212 14.7 1.3 0.182
T60 794 140 233 388 581 212 14.7 1.3 0.182
T80 794 140 233 194 775 212 14.7 1.3 0.182
T100 794 140 233 - 969 212 14.7 1.3 0.182
w/b = water to binder ratio
*Solid content

100
T0 RMSE = 5.433
90 R2 = 0.975
T20
80 T40 RMSE = 4.340
T60 R2 = 0.984
70
T80 RMSE = 3.808
60 T100 R2 = 0.987
% Finer

50 Target (modified A&A) RMSE = 4.063


R2 = 0.985
40 RMSE = 4.985
R2 = 0.977
30
20
RMSE = 6.288
10 R2 = 0.963
0
0.1 1 10 100 1000
Particle size (μm)

Figure 5.3. Combined particle size distributions of UHPC mixes (RMSE = root-mean-square
error, R2 = determination coefficient)

129
Chapter 5

5.2.3 Specimen preparation

For the preparation of each UHPC mix, dry constituents such as cement, silica fume, fly
ash, quartz sand, and/or WAGT1 were mixed using a mixer for 3 minutes. After that, water,
HRWR, and defoaming agent were added to the dry mixture and mixed until the mixture
became flowable. 50 × 50 × 50 mm cube moulds were used to prepare the specimens for
compressive strength tests, 50 × 100 mm cylindrical moulds for splitting tensile strength
tests, and 100 × 100 × 100 mm cube moulds for durability tests. The moulded specimens
were vibrated for 1 minute and then moved to a humidity- and temperature-controlled curing
room (95% relative humidity, 21 °C temperature). The specimens were demoulded after 24
h. The mechanical test specimens were fully submerged in water (21 °C) until the time of
testing, while the durability specimens were left inside the humidity-controlled curing room.
After 28 days, the carbonation test specimens were kept in a CO2-controlled chamber (60%
relative humidity, 21 °C temperature) at an elevated CO2 concentration of 10,000 ppm for
56 days.

5.2.4 Test methods

5.2.4.1 Fresh properties

Slump flow tests were conducted for the fresh UHPC mixes in accordance with AS 1012.3.5
(2015). A mini slump cone, having a height of 116 mm, a top diameter of 38 mm, and a
bottom diameter of 76 mm, was used for the slump flow tests. The aspect ratio of the cone
was identical to that of the standard cone specified by AS 1012.3.5 (2015).

Wet packing density was measured for each fresh UHPC mix based on the calculation
method presented in (Wong and Kwan 2008). Wet packing density (φ) can be expressed by
Equation 5.2.

𝑉𝑠
𝜑= (5.2)
𝑉

V is the volume of mould (mm3), and Vs is the total volume of solid constituents (mm3). Vs
can be determined using Equation 5.3.

𝑚
𝑀 ∑𝑛𝑘=1 𝑘
𝐺𝑘
𝑉𝑠 = ( 𝑛 ) (5.3)
𝜌𝑤 𝑚𝑤 + ∑𝑘=1 𝑚𝑘

M is the mass of fresh UHPC occupying a volume of V after being subjected to 1 min of
vibration (g), ρw is the density of water (g/mm3), mw is the mass of water per unit volume (g),

130
Chapter 5

mk is the mass of each solid material per unit volume (g), Gk is the specific gravity of each
dry material.

5.2.4.2 Mechanical tests

A compression testing machine with a capacity of 600 kN was used to perform the
mechanical tests. The compressive strength tests were conducted at 7, 28, and 56 days
according to guidelines of ASTM C109 / C109M-16a (2016). Strain gauges were attached
to the 28-day specimens for measuring their strains under compression. The elastic moduli
of the specimens were calculated according to AS 1012.17 (1997). The splitting tensile
strength tests were conducted at 28 days according to the guidelines of AS 1012.10 (2000).
For each mechanical property, three specimens from each mix were tested.

5.2.4.3 Durability tests

The initial rate of absorption (IRA) tests were conducted using 100 × 100 × 100 mm plain
cubes in accordance with the guidelines of IS 3495 (Part 2) (1992), and the water absorption
(WA) tests were conducted using the same specimens as per the guidelines of AS/NZS
4456.17 (2003). To calculate the WA of each specimen Equation 5.4 was used, and Equation
5.5 was used to calculate the IRA. Three specimens from each mix were tested for both WA
and IRA.

100(𝑚𝑖 − 𝑚𝑓 )
𝑊𝐴 = (%) (5.4)
𝑚𝑖

1000(𝑚𝑓,60 − 𝑚𝑖 )
𝐼𝑅𝐴 = (kg/m2/min) (5.5)
𝐴

mi is the mass of oven-dried specimen (g), mf is the mass of specimen after 24 hours of
immersion in water (g), mf,60 is the mass of specimen after immersion of bottom 3 mm in
water for 60 s (g), A is the bottom surface area of specimen (mm2). Excess water on the
specimen surface was wiped off using a damp cloth prior to measuring mf or mf,60.

Carbonation tests were carried out according to EN 13295 (2004) using 100 × 100 × 100
mm specimens. Each specimen was split into two halves. Phenolphthalein indicator
(mixture of 1 g phenolphthalein powder, 70 ml ethanol, and 30 ml water) was then sprayed
over the split section of each half. The colour change of each of the sections was observed
to assess the carbonation resistance of the respective mix.

131
Chapter 5

Figure 5.4. Leachate sample preparation for leaching toxicity test – (a) horizontal shaking of
UHPC pieces and leaching liquid, and (b) vacuum filtration of leachate

Table 5.4. Fresh properties of UHPC mixes

Mix T0 T20 T40 T60 T80 T100


Slum flow diameter (mm) 210 170 166.5 155 145 105
Wet packing density 0.771 0.793 0.795 0.808 0.789 0.792

5.2.4.4 Leaching toxicity test

Leaching toxicity test was conducted for each UHPC mix using a slightly modified version
of the approach mentioned in HJ 557 (2010). HJ 557 (2010) specifies the use of deionized
water as leaching liquid. However, in this study, acetic acid (CH3COOH) solution having a
pH of 2.65 was chosen as the leaching liquid, complying with (Wang et al. 2018). 100 g of
crushed UHPC pieces with a maximum size of 3 mm were mixed with 1 L of leaching liquid.
The mixture was subjected to 8 h of horizontal agitation and then left at a steady position
for 16 h. After that, the supernatant was filtered through vacuum-filtration process (Figure
5.4). The concentrations of heavy metal in the supernatant were measured using ICP-OES.

5.2.4.5 SEM analysis

SEM analysis was conducted using small pieces of T0 and T100. Backscattered scanning
electron micrographs were captured using a field emission scanning electron microscope
(FESEM).

5.3 Results and discussion

5.3.1 Fresh behaviour

The slump flow diameters of the fresh UHPC mixes are presented in Table 5.4.

132
Chapter 5

The slump flow diameters of the UHPC mixes decreased with the increasing replacement of
sand by WAGT1. The reduction of the slump flow due to the incorporation of WAGT1 can
be attributed to the non-uniform surface texture (Figure 5.1) and fine size of WAGT1
particles (from Table 5.2, the median diameter of the sand was 237.3 µm, and the median
diameter of WAGT1 was 141.4 µm). WAGT1, owing to their higher fineness, had higher
surface-area per unit volume than that of the sand. Consequently, the inclusion of WAGT1,
substituting certain percentage of the sand, increased the water demand to lubricate the
particles present in fresh UHPC and reduced its slump flow (Shettima et al. 2016). Table 5.4
also presents the wet packing densities of the UHPC mixes. Although the UHPCs with
WAGT1 exhibited lower slump than that of T0, their packing densities were, in general,
higher than that of T0. This observation is in agreement with the RMSE and R2 values
presented in Figure 5.3.

5.3.2 Compressive strength and splitting tensile strength

The 7-day, 28-day, and 56-day compressive strengths of the UHPCs with different replacements
of quartz sand by WAGT1 are presented in Figure 5.5(a). Up to 80% substitution of quartz sand
by WAGT1, the 28-day compressive strength of UHPC remained either comparable to or slightly
higher than that of T0. Similar phenomena were observed at 7 and 56 days. Substitution of
sand by WAGT1, beyond 80%, impaired the strength of UHPC if compared with the
strength of T0. The 28-day compressive strength of T0 was 142.6 MPa, and the 56-day
strength was 151.9 MPa. The 28-day compressive strength changed by +4.6% for T20,
−1.6% for T40 +4.4% for T60, −1.3% for T80, and −7.6% for T100, with respect to the 28-
day strength of T0. At 56 days, the compressive strength changed by +4.8% for T20, −0.1%
for T40, +2.2% for T60, −2.0% for T80, and −6.6% for T100, with respect to that of T0. The
compressive strength of UHPC incorporating WAGT1 can be influenced by different
factors. High fineness of WAGT1 particles is likely to result in better particle packing,
consequently higher packing density as can be seen from Table 5.4, and reduce porosity of
hardened UHPC (Zhang et al. 2020a). Such feature of WAGT1 is expected to benefit the
strength of UHPC. On the other hand, low crystalline quartz content (Table 5.2) and
presence of softer minerals (e.g., muscovite, halite, polyhalite, etc.) in WAGT1 (Figure 5.1)
tend to impart low elastic modulus to UHPC (Perkins 1999). Low quartz content of
WAGT1, as a result, is likely to cause reduced strength compared to UHPC with quartz
sand (Zhao et al. 2014). Improved packing density of UHPC attained through WAGT1
incorporation most likely helped the UHPCs with up to 80% WAGT1 achieve strength
comparable to or slightly higher than that of T0. The reduction in strength of T100 can be
attributed to the low quartz content in WAGT1. The 28-day compressive strengths of the

133
Chapter 5

UHPCs made with up to 100% substitution of quartz sand by WAGT1 were greater than
120 MPa. T100 was observed to achieve 28-day strength (131.8 MPa) comparable to that
presented in (Pyo et al. 2018) for UHPC with 30% quartz based tailings as aggregate, and
higher than that presented in (Zhao et al. 2014) for UHPC with 100% iron ore tailings (Table
5.1). UHPCs made with up to 60% substitution of quartz sand by WAGT1 attained
compressive strengths greater than 150 MPa at 56 days, without employing any special
curing (e.g., autoclaving, heat-curing, steam-curing, etc.). It is also worth noting that the
ECO-UHPCs developed in this study exhibited substantially higher compressive strengths
than conventional NSC (exhibiting strength usually less than 40 MPa). The ECO-UHPC
mixes can be used to construct slenderer structural members (Fehling et al. 2015), and
thinner roads or pavements (Larrard and Sedran 2011).

The 28-day splitting tensile strengths of the UHPCs are presented in Figure 5.5(b). The
splitting tensile strength of T0 was 7.51 MPa. The splitting tensile strength changed by
+8.9% for T20, +13.2% for T40, +5.5% for T60, −9.2% for T80, and −2.8% for T100, with
respect to that of T0. Dense particle packing facilitated by high fineness of WAGT1
positively influenced the splitting tensile strengths of the UHPCs with up to 60% WAGT1.
However, presence of soft minerals on the surfaces of WAGT1 particles perhaps resulted in
weak bonds between the concrete matrix and WAGT1 particles (Alsalman et al. 2017).
Formation of such weak WAGT1 particle – matrix bonds was possibly the governing factor
to impart reduced splitting tensile strengths to the UHPCs with WAGT1 content of more
than 60%.

Figure 5.5. (a) Compressive strength results, and (b) 28-day splitting tensile strength results of
UHPCs

134
Chapter 5

Figure 5.6. (a) 28-day elastic moduli of UHPCs, (b) Ec as a function of √(fc), and (c) Ec as a
function of √(fc)/φ; (Ec = elastic modulus, fc = compressive strength, φ = wet packing density)

5.3.3 Elastic modulus

The 28-day elastic moduli of the UHPC mixes are presented in Figure 5.6(a). The results
show that the elastic modulus of UHPC generally decreased with the increase of WAGT1
content. This is most likely because WAGT1 were softer than the quartz sand owing to the
lower quartz content (Table 5.2) and presence of softer minerals (e.g., muscovite, halite,
polyhalite, etc.) in WAGT1 (Figure 5.1). The 28-day elastic modulus of T0 was 42.6 GPa.
The elastic modulus decreased by 1.4% for T20, 4.9% for T40, 5.8% for T60, 7.2% for T80,
and 10.9% for T100, with respect to the elastic modulus of T0.

The elastic moduli of the UHPC mixes are plotted against the square roots of their respective
compressive strengths [√(fc)] in Figure 5.6(b). The R2 obtained for the line of best fit was
found to be 0.47, and the RMSE was found to be 1.12. In Figure 5.6(c), the elastic moduli
of the UHPC mixes are plotted against their respective √(fc) to φ (wet packing density) ratios.
Figure 5.6(c) shows a low scatter of the presented data compared to that of the data presented

135
Chapter 5

in Figure 5.6(b). The R2 between the line of best fit and the data presented in Figure 5.6(c)
was found to be 0.90, and the RMSE was found to be 0.50. Lower scatter of the data in
Figure 5.6(c) implies that the elastic modulus of UHPC incorporating WAGT1 can be better
represented as a function of √(fc)/φ than as a function of √(fc). This is possibly because, in
comparison to the compressive strength, the elastic modulus of UHPC incorporating
WAGT1 is less sensitive to the variation of the packing density. Due to limited datapoints,
the present study did not propose any equation. The study can be extended in future
compiling more elastic modulus, compressive strength, and wet packing density results to
develop a relationship for estimating the elastic modulus of UHPC incorporating WAGT1.

5.3.4 Durability

Figure 5.7 shows the water absorption (WA) and the initial rate of absorption (IRA) test
results of the UHPCs at 28 days. The WAs and IRAs of the UHPCs incorporating WAGT1
were generally lower than those of T0. Zhang et al. (2020a) observed similar behaviour in
their investigation when manufactured sand was replaced by finer iron ore tailings. Fine
particles of WAGT1 most likely helped fill the micro-voids of UHPC mix, and helped
UHPC to attain lower 28-day WA and IRA (Gupta and Vyas 2018). The 28-day WA was
1.123% for T0, 1.121% for T20, 1.028% for T40, 1.012% for T60, 1.011% for T80, and
0.866% for T100. The IRA was 0.082 kg/m2/min for T0, 0.067 kg/m2/min for T20, 0.078
kg/m2/min for T40, 0.055 kg/m2/min for T60, 0.046 kg/m2/min for T80, and 0.055
kg/m2/min for T100, at 28 days. It was interesting to observe that the WAs of all the UHPCs
were less than 1.5%. Concrete exhibiting WA less than 2% is considered durable, for
construction works with fifty-year design life in coastal regions (Ahmed et al. 2021). The
WA of NSC is typically much higher than 3.5% (Neville 2011). The ECO-UHPCs developed
in this study can be used to prepare more durable structures, pavements, or roads than those
built with NSC.

Figure 5.7. Water absorption and initial rate of absorption results of UHPCs

136
Chapter 5

Figure 5.8. Illustration of carbonation resistances of UHPCs

Figure 5.8 shows the colour changes of the cut sections of elevated CO2 cured UHPC
specimens upon the application of phenolphthalein solution. Dark pink colour was observed
throughout the cross-sections of all the UHPC specimens. This implies no visible
carbonation front was present in any of the specimens. In other words, the depth of
carbonation was below the detection limit (<0.5 mm) for all the UHPCs. These observations
on the UHPCs incorporating different contents of WAGT1 agree with the results obtained
by Scheydt and Müller (2012) for conventional UHPC. Promising carbonation resistances
of the UHPCs, irrespective of their aggregate types, can mainly be attributed to their very
dense matrices owing to the use of very low w/b ratio (=0.182) and the use of silica fume.
The dense microstructures of the UHPCs substantially restricted the diffusion of CO 2 into
the concretes and resulted in almost zero carbonation depths.

5.3.5 Leaching toxicity

The concentrations of the heavy metals in the leachates extracted from different UHPC –
leaching liquid mixtures are presented in Table 5.5. For all the UHPCs, the concentrations
of the heavy metals were well below their respective regulatory thresholds set by different
environment protection guidelines and regulations. Such low concentrations of the toxic
metals in the leachates, compared to their respective regulatory thresholds, indicate safe use
of WAGT1 in UHPC. Long-term leaching test can be conducted in future for gaining deeper
insight into the leaching behaviour of heavy metals in UHPCs incorporating WAGT1.
Three factors possibly influenced the leaching toxicity test results of the current study. First,
concentrations of some of the heavy metals (e.g., Cd, Pb) were very low in the as-received

137
Chapter 5

WAGT1. Second, certain heavy metals in WAGT1 were perhaps present in oxidation states
which show low leachability or low mobility (e.g., possible presence of arsenic as arsenate)
(Shrivastava et al. 2015). Third, the very dense microstructure of UHPC highly restricted
the leaching behaviour of toxic metals present in WAGT1 (Wang et al. 2018). The leaching
behaviour of leachable heavy metals from a particular type of tailings in UHPC, compared
to their behaviours in low-strength and normal-strength mortars, can be illustrated by the
schematic diagram shown in Figure 5.9. In comparison to low-strength and normal-strength
mortars, UHPC is very dense and less porous because of its low w/b ratio and the presence
of silica fume in it. The binder to aggregate ratio of UHPC is also much higher than those
of low-strength and normal-strength mortars. Because of this, the aggregate particles in
UHPC are covered by thick layers of paste. Foregoing features of UHPC have the potential
to provide substantial hindrance to the leaching of heavy metals from the tailings present in
it (Liu et al. 2020b).

Table 5.5. Concentrations (mg/L) of heavy metals in leachate samples

Heavy Sample Regulatory threshold


metal T0 T20 T40 T60 T80 T100 NSW GB GB 40
EPA 5085.3 16889 CFR
(2014) (2007) (2008) 261.24
Zn 0.181 0.192 0.215 0.159 0.149 0.162 - 100 100 -
Cd <0.001 <0.001 <0.001 <0.001 <0.001 <0.001 1 1 0.15 1
As <0.005 <0.005 <0.005 <0.005 0.006 <0.005 5 5 0.3 5
Pb <0.005 <0.005 <0.005 <0.005 <0.005 <0.005 5 5 0.25 5
Cu 0.029 0.046 0.067 0.070 0.076 0.139 - 100 40 -

Figure 5.9. Leaching behaviour of leachable heavy metals from tailings in – (a) low-strength
mortar, (b) normal-strength mortar, and (c) UHPC

138
Chapter 5

5.3.6 Microstructure analysis

Figure 5.10(a) shows the backscattered electron micrograph of T0, and Figure 5.10(b) shows
that of T100. Because of the low w/b ratio, a lot of cement particles (white features in the
matrix) were seen to remain unreacted in both T0 and T100 samples. Larger pores were
observed in the matrix of T0. This indicates denser packing was achieved for T100 than that
of T0. The observation is consistent with the higher wet packing density of T100 than that
of T0 (Table 5.4). The WAGT1 particles exhibited bright colour in the backscattered
electron microscopy, compared to the sand particles. This is because the sand mainly
contained Si-based quartz whereas WAGT1, apart from Si-based quartz, had minerals
incorporating heavier elements (such as, Ca, Fe, etc.) (Table 5.2). WAGT1 particles were
generally observed to be surrounded by wider cracks than sand particles. This may be
attributed to the weak bonding between WAGT1 particles and matrix, occurring because of
the presence of soft minerals in WAGT1 (Alsalman et al. 2017). Such property of WAGT1
imparted low strength to T100 if compared with the strength of T0 (Perkins 1999).

5.3.7 Economic and environmental evaluation

Figure 5.11(a) presents the estimated material costs and combined material and
transportation costs of UHPCs with different WAGT1. The estimated CO2 footprints of the
UHPC mixes, based on both the CO2 emission from materials and the combined emission
from materials and transportation, are presented in Figure 5.11(b).

Table 5.6 presents the data used for the calculation of the material costs and CO2 footprints
of different UHPC mixes. The material costs and CO2 footprints of the cement, silica fume,
fly ash, and quartz sand reported in Chapter 4 (Ahmed et al. 2021) were used for the
calculation of the material cost and CO2 footprint of T0. For calculating the material costs
and CO2 footprints of the mixes with WAGT1, approximate cost (A$53/t) and CO2
emission (2.1 kg/t) for screening out coarse rock fraction from raw WAGT1 were
considered, in addition to the costs and CO2 footprints of other materials reported in Chapter
4 (Ahmed et al. 2021).

In order to determine the costs and CO2 emissions from transportation, the cement, silica
fume, fly ash, quartz sand, and chemical admixtures (HRWR and defoamer) were assumed
to be transported from local sources in WA. The cement source was assumed to be located
at a distance of 20.1 km from the Perth CBD (central business district), silica fume source at
a distance of 152 km, fly ash source at a distance of 209 km, quartz sand source at a distance
of 28.9 km, and the source of chemical admixtures at a distance of 23 km. Due to the

139
Chapter 5

confidentiality agreement the exact distance of Mine 1 from the CBD is not mentioned here.
The cost of transport (articulated truck freight) was considered to be A$0.09/t/km (Shaikh
et al. 2019), and CO2 emission was considered to be 0.071 kg/t/km (O’Brien et al. 2009).

Figure 5.10. Backscattered electron micrographs of (a) T0, and (b) T100; (P = pore, S = quartz
sand, T = WAGT1)

140
Chapter 5

Figure 5.11. (a) Costs, and (b) CO2 footprints of UHPC mixes

Table 5.6. CO2 footprints and costs of materials

Material Cement Silica Fly Quartz WAGT1 Water Admixturea


fume ash sand
Material CO2 959b 82b 93b 21b 2.1 1b 508b
footprint (kg/t)
Material cost 0.75b 0.52b 0.51b 0.56b 0.053 0.0015b 4.4b (HRWR)
(A$/kg) 3.23b (defoaming
agent)
a
Including liquid content
b
Adopted from (Ahmed et al. 2021)

The estimated material cost of T0 is A$1586/m3. The material cost can be reduced by 6.2%
for T20, 12.4% for T40, 18.6% for T60, 24.8% for T80, and 31% for T100, with respect to
the cost of T0. Very low processing cost of WAGT1, in comparison to the cost related to the
energy-intensive processes of extraction, grinding, and refining of quartz sand, imparts
reduced material costs to the mixes with WAGT1. The combined material and
transportation costs of UHPC, for constructions near both Perth CBD and Mine 1, also
reduce when quartz sand is replaced by WAGT1. The combined material and transportation
cost reduces by 4.3% for T20, 8.6% for T40, 12.9% for T60, 17.2% for T80, and 21.6% for
T100, with respect to the cost of T0, for construction near the CBD. For construction in the
area near Mine 1, the combined cost can be reduced by 6.6% for T20, 13.2% for T40, 19.9%
for T60, 26.5% for T80, and 33.1% for T100.

The material CO2 footprint of UHPC reduces slightly when quartz sand is substituted by
WAGT1. This can mainly be attributed to the reduction/ omission of the extraction,
grinding, and refining phases of quartz sand manufacturing process when WAGT1 are used.
However, the reduction in material CO2 footprint of UHPC, for certain replacement of sand
by WAGT1, is less pronounced compared to the reduction in material cost. This is because

141
Chapter 5

the use of high-volume cement for UHPC production is the major contributor (90%–92%)
to the material CO2 footprint of UHPC. For construction near the CBD, the combined CO2
emission from materials and transportation increases by up to 11.6% when quartz sand is
replaced by WAGT1. This can be attributed to the large distance of Mine 1 from the CBD,
compared to the distance between the quartz sand source and the CBD. Studies can be
carried out in future investigating the influences of using gold tailings from mines nearer to
the CBD on the properties of UHPC. However, for construction near Mine 1, the combined
CO2 emission from materials and transportation can be decreased by up 12.1% if WAGT1
are used instead of quartz sand.

Along with the CO2 emission, extraction and grinding processes of quartz sand have a
number of other environmental consequences. Some of the most important ones, as
indicated in (Gavriletea 2017), are – deforestation, loss of biodiversity, soil degradation,
lowering of water table, particulate matter air pollution during grinding, etc. The natural
sources of quartz sand are depleting so drastically that the demand for sand may surpass the
supply by mid-century (Bendixen et al. 2019). Long-term continuous disposal of tailings is
also associated with several concerns. Fine fraction of the disposed tailings can pollute the
surrounding air (Dong et al. 2019). To facilitate long-term disposal, construction of new
tailings storage facilities (TSFs) or gradual enlargement of the existing tailings dam structure
is often required. Construction of new TSFs requires the use of additional landmasses.
Again, gradual enlargement of the tailings dam structure increases the risk of dam failure as
well as the difficulty of tailings management (Edraki et al. 2014). Dam failure may pose a
severe threat to the safety of the workers and the people residing in the locality, and can even
lead to the pollution of surrounding surface and ground waters, soil, crops, etc. (Ince 2019).
Although WAGT1 at Mine 1 are currently managed as per the guidelines of ANCOLD
(Australian National Committee on Large Dams) and ICMM (International Council on
Mining and Metals), continuous disposal of WAGT1 may make the management more
challenging in future years. The problems, related both to the manufacturing process of fine
quartz sand and to the disposal of WAGT1, can be addressed to some extent if quartz sand
in UHPC is partially or fully substituted by WAGT1.

5.4 Conclusion

Based on the results and discussion presented in the preceding section, the following
conclusions can be drawn:

1. UHPCs, incorporating up to 80% WAGT1, exhibit compressive strength comparable to


or slightly higher than the compressive strength of UHPC with 100% quartz sand. The

142
Chapter 5

28-day compressive strength of UHPC with 20%–80% WAGT1 changes by −1.3% to


+4.58%, with respect to the 28-day strength of UHPC with 100% quartz sand. UHPC
with 100% WAGT1 exhibits 7.6% lower compressive strength with respect to that of
UHPC with 100% quartz sand. However, strength greater than 120 MPa is achievable
at 28 days for UHPC with 100% WAGT1. 28-day strength achieved in this study for
UHPC incorporating 100% WAGT1 (131.8 MPa) is also comparable to or higher than
those presented in the previous studies for ambient temperature cured UHPCs with
quartz based tailings and iron ore tailings. UHPCs made with up to 60% substitution of
quartz sand by WAGT1 can attain compressive strengths greater than 150 MPa at 56
days without employing any special curing.
2. UHPCs with up to 60% WAGT1 generally exhibit higher (5.5% to 13.2%) splitting
tensile strengths than that of UHPC without WAGT1. Beyond 60% replacement,
splitting tensile strength generally decreases (2.8% to 9.2%). The elastic modulus of
UHPC generally reduces (up to 10.9%) as quartz sand is replaced by WAGT1.
3. The flowability of fresh UHPC reduces with the substitution of quartz sand by WAGT1.
However, better wet packing density can be attained for UHPC with WAGT1 than that
of UHPC with 100% quartz sand.
4. UHPCs incorporating WAGT1 generally exhibit lower water absorption and initial rate
of absorption values than those of UHPC with 100% quartz sand. UHPC, irrespective
of WAGT1 content, shows substantial resistance to carbonation.
5. The leachability of toxic metals from UHPC, irrespective of its WAGT1 content, lies
well below the regulatory thresholds. Long-term leaching test can be conducted in future
for gaining further insight into the leaching behaviour of heavy metals in UHPCs
incorporating WAGT1.
6. The combined material and transportation cost of UHPC can be reduced by up to 33.1%,
for construction in the area near Mine 1. For construction near the Perth CBD, the
combined cost reduces by up to 21.6% when quartz sand is replaced by WAGT1.
7. For construction near Mine 1, the combined CO2 emission from materials and transport
can be reduced by up to 12.1% when quartz sand is replaced by WAGT1. However, the
combined CO2 emission can increase by up to 11.6%, for construction near the Perth
CBD. Studies can be carried out in future to investigate the properties of UHPCs
incorporating gold tailings from mines nearer to the CBD, which may be economically
and environmentally more attractive for construction near the CBD.
8. Utilization of WAGT1 as quartz sand substitute for UHPC production, especially near
Mine 1, can reduce the demand for quartz sand to some extent. In the long run, negative
environmental impacts of quartz sand extraction can be reduced (e.g., deforestation, loss

143
Chapter 5

of biodiversity, soil degradation, lowering of water table, etc.). Cleaner UHPC


production in such a way can also help prevent concerns/ problems, related to the long-
term disposal of WAGT1, from arising in the future (e.g., need for new landmasses, risk
of tailings dam failure, etc.).

In summary, UHPC exhibiting 28-day compressive strength greater than 120 MPa, water
tightness better than UHPC incorporating quartz sand, and carbonation resistance similar
to conventional UHPC can be produced using WAGT1 as aggregate. The cost and CO2
emission can also be reduced for construction near Mine 1 if WAGT1 is used instead of
quartz sand. The results imply, in the area near Mine 1, the use of WAGT1 can
economically and environmentally, as well as in terms of durability, be a better option than
quartz sand for construction works that require UHPC with 28-strength in excess of 120
MPa. Moreover, the ECO-UHPCs developed in this study, owing to their substantially
higher strength and better durability than NSC, can be used to construct slenderer and/ or
more durable structural members, pavements, or roads than those built with NSC.

Acknowledgements

The authors acknowledge the facilities provided by the Centre for Microscopy,
Characterization & Analysis, The University of Western Australia for performing the XRD
and SEM analyses. Special thanks are given to Mr. Andrew van de Ven for his support to
conduct the leaching toxicity test. The research was carried out while the first author was a
recipient of a University Postgraduate Award and the Australian Government Research
Training Program Scholarship at The University of Western Australia.

References
40 CFR 261.24. Toxicity characteristic.
Ahmed, T., Elchalakani, M., Karrech, A., Dong, M., Mohamed Ali, M. S., and Yang, H.
(2021). “ECO-UHPC with High-Volume Class-F Fly Ash: New Insight into Mechanical
and Durability Properties.” Journal of Materials in Civil Engineering, 33(7), 0003726.
Alsalman, A., Dang, C. N., Prinz, G. S., and Hale, W. M. (2017). “Evaluation of modulus
of elasticity of ultra-high performance concrete.” Construction and Building Materials, 153,
918–928.
AS/NZS 3582.3. (2016). Supplementary cementitious materials for use with portland and blended
cement - amorphous silica.
AS/NZS 4456.17. (2003). Masonry units and segmental pavers and flags - methods of test:
determining initial rate of absorption (suction).
AS 1012.10. (2000). Methods of testing concrete: determination of indirect tensile strength of concrete
cylinders (“Brazil” or splitting test).
AS 1012.17. (1997). Methods of testing concrete: determination of the static chord modulus of

144
Chapter 5

elasticity and Poisson’s ratio of concrete specimens.


AS 1012.3.5. (2015). Methods of testing concrete: determination of properties related to the
consistency of concrete - slump flow, T500 and J-ring test.
ASTM C150 / C150M-16. (2016). Standard specification for portland cement. West
Conshohocken, PA.
ASTM C618-19. (2019). Standard specification for coal fly ash and raw or calcined natural pozzolan
for use in concrete. West Conshohocken, PA.
ASTMC109 / C109M-16a. (2016). Standard test method for compressive strength of hydraulic
cement mortars (using 2-in. or [50-mm] cube specimens). West Conshohocken, PA.
Bendixen, M., Best, J., Hackney, C., and Iversen, L. L. (2019). “Time is running out for
sand.” Nature, 571(7763), 29–31.
Courtial, M., De Noirfontaine, M. N., Dunstetter, F., Signes-Frehel, M., Mounanga, P.,
Cherkaoui, K., and Khelidj, A. (2013). “Effect of polycarboxylate and crushed quartz in
UHPC: microstructural investigation.” Construction and Building Materials, 44, 699–705.
Dong, L., Tong, X., Li, X., Zhou, J., Wang, S., and Liu, B. (2019). “Some developments
and new insights of environmental problems and deep mining strategy for cleaner
production in mines.” Journal of Cleaner Production, 210, 1562–1578.
Edraki, M., Baumgartl, T., Manlapig, E., Bradshaw, D., Franks, D. M., and Moran, C. J.
(2014). “Designing mine tailings for better environmental, social and economic
outcomes: a review of alternative approaches.” Journal of Cleaner Production, 84(1), 411–
420.
EN 13295. (2004). Products and systems for the protection and repair of concrete structures - test
methods - determination of resistance to carbonation.
Fehling, E., Schmidt, M., Walraven, J., Leutbecher, T., and Fröhlich, S. (2015). Ultra-high
performance concrete UHPC: fundamentals, design, examples. Ultra-High Performance Concrete
UHPC: Fundamentals, Design, Examples.
Fisonga, M., Wang, F., and Mutambo, V. (2019). “Sustainable utilization of copper tailings
and tyre-derived aggregates in highway concrete traffic barriers.” Construction and
Building Materials, 216, 29–39.
Gao, S., Cui, X., Kang, S., and Ding, Y. (2020). “Sustainable applications for utilizing
molybdenum tailings in concrete.” Journal of Cleaner Production, 266, 122020.
Gavriletea, M. D. (2017). “Environmental impacts of sand exploitation: analysis of sand
market.” Sustainability (Switzerland), 9(7), 1118.
GB 16889. (2008). Standard for pollution control on the landfill site of municipal solid waste.
GB 5085.3. (2007). Identification standards for hazardous wastes - identification for extraction
toxicity.
Gupta, L. K., and Vyas, A. K. (2018). “Impact on mechanical properties of cement sand
mortar containing waste granite powder.” Construction and Building Materials, 191, 155–
164.
Gupta, R. C., Mehra, P., and Thomas, B. S. (2017). “Utilization of copper tailing in
developing sustainable and durable concrete.” Journal of Materials in Civil Engineering,
29(5), 04016274.
HJ 557. (2010). Solid waste─extraction procedure for leaching toxicity ─ horizontal vibration
method.
Huang, H., Gao, X., and Jia, D. (2019). “Effects of rheological performance, antifoaming
admixture, and mixing procedure on air bubbles and strength of UHPC.” Journal of
Materials in Civil Engineering, 31(4), 04019016.
Ince, C. (2019). “Reusing gold-mine tailings in cement mortars: mechanical properties and

145
Chapter 5

socio-economic developments for the Lefke-Xeros area of Cyprus.” Journal of Cleaner


Production, 238, 117871.
IS 3495 (Part 2). (1992). Methods of tests of burnt clay building bricks, Part 2: determination of water
absorption. New Delhi.
Jiao, Y., Zhang, Y., Guo, M., Zhang, L., Ning, H., and Liu, S. (2020). “Mechanical and
fracture properties of ultra-high performance concrete (UHPC) containing waste glass
sand as partial replacement material.” Journal of Cleaner Production, 277, 123501.
Kathirvel, P., Kwon, S. J., Lee, H. S., Karthick, S., and Saraswathy, V. (2018). “Graphite
ore tailings as partial replacement of sand in concrete.” ACI Materials Journal, 115(3),
481–492.
Kim, H., Koh, T., and Pyo, S. (2016). “Enhancing flowability and sustainability of ultra
high performance concrete incorporating high replacement levels of industrial slags.”
Construction and Building Materials, 123, 153–160.
Kinnunen, P., Ismailov, A., Solismaa, S., Sreenivasan, H., Räisänen, M. L., Levänen, E.,
and Illikainen, M. (2018). “Recycling mine tailings in chemically bonded ceramics – a
review.” Journal of Cleaner Production, 174, 634–649.
Larrard, F. de, and Sedran, T. (2011). “High and ultra-High performance concrete in
pavement: tools for the road eternity.” 9th International Symposium on High Performance
Concrete: Design, Verification and Utilization, Rotorua, New Zealand, hal-00877025.
Li, T., Wang, S., Xu, F., Meng, X., Li, B., and Zhan, M. (2020). “Study of the basic
mechanical properties and degradation mechanism of recycled concrete with tailings
before and after carbonation.” Journal of Cleaner Production, 259, 120923.
Liu, H., Li, B., Xue, J., Hu, J., and Zhang, J. (2020a). “Mechanical and electroconductivity
properties of graphite tailings concrete.” Advances in Materials Science and Engineering,
2020, 9385097.
Liu, T., Wei, H., Zou, D., Zhou, A., and Jian, H. (2020b). “Utilization of waste cathode
ray tube funnel glass for ultra-high performance concrete.” Journal of Cleaner Production,
249, 119333.
Logsdon, M. J., Hagelstein, K., and Mudder, T. I. (1999). The management of cyanide in gold
extraction. Ottawa.
Nematollahi, B., M.R., R. S., and Voo, Y. L. (2011). “Ultra high performance concrete
(UHPC) technology from material to structure : a review.” International building &
infrastructure technology conference, A. N. A. Ghani, M. A. O. Mydin, and N. F. Abas,
eds., Universiti Sains Malaysia, Penang, Malaysia, 359–366.
Neville, A. M. (2011). Properties of concrete. Pearson, London, UK.
NSW EPA. (2014). Waste classification guidelines – Part 1: classification of waste. Sydney.
O’Brien, K. R., Ménaché, J., and O’Moore, L. M. (2009). “Impact of fly ash content and fly
ash transportation distance on embodied greenhouse gas emissions and water
consumption in concrete.” International Journal of Life Cycle Assessment, 14(7), 621–629.
Perkins, W. G. (1999). “Polymer toughness and impact resistance.” Polymer Engineering and
Science, 39(12), 2445–2460.
Protasio, F. N. M., Avillez, R. R. de, Letichevsky, S., and Silva, F. de A. (2021). “The use
of iron ore tailings obtained from the Germano dam in the production of a sustainable
concrete.” Journal of Cleaner Production, 278, 123929.
Pyo, S., Tafesse, M., Kim, B. J., and Kim, H. K. (2018). “Effects of quartz-based mine
tailings on characteristics and leaching behavior of ultra-high performance concrete.”
Construction and Building Materials, 166, 110–117.
Scheydt, J., and Müller, H. (2012). “Microstructure of ultra high performance concrete

146
Chapter 5

(UHPC) and its impact on durability.” 3rd International Symposium on on Ultra High
Performance Concrete, 349–356.
Shaikh, F. U. A., Nath, P., Hosan, A., John, M., and Biswas, W. K. (2019). “Sustainability
assessment of recycled aggregates concrete mixes containing industrial by-products.”
Materials Today Sustainability, 5.
Shettima, A. U., Hussin, M. W., Ahmad, Y., and Mirza, J. (2016). “Evaluation of iron ore
tailings as replacement for fine aggregate in concrete.” Construction and Building Materials,
120, 72–79.
Shrivastava, A., Ghosh, D., Dash, A., and Bose, S. (2015). “Arsenic contamination in soil
and sediment in India: sources, effects, and remediation.” Current Pollution Reports, 1(1),
35–46.
Thomas, B. S., Damare, A., and Gupta, R. C. (2013). “Strength and durability
characteristics of copper tailing concrete.” Construction and Building Materials, 48, 894–
900.
Tian, Z. X., Zhao, Z. H., Dai, C. Q., and Liu, S. J. (2016). “Experimental study on the
properties of concrete mixed with iron ore tailings.” Advances in Materials Science and
Engineering, 2016, 8606505.
UNEP. (2014). Sand, rarer than one thinks.
Wang, X., Yu, R., Shui, Z., Zhao, Z., Song, Q., Yang, B., and Fan, D. (2018).
“Development of a novel cleaner construction product: ultra-high performance concrete
incorporating lead-zinc tailings.” Journal of Cleaner Production, 196, 172–182.
Wong, H. H. C., and Kwan, A. K. H. (2008). “Packing density of cementitious materials:
Part 1-measurement using a wet packing method.” Materials and Structures/Materiaux et
Constructions, 41(4), 689–701.
Xu, W., Wen, X., Wei, J., Xu, P., Zhang, B., Yu, Q., and Ma, H. (2018). “Feasibility of
kaolin tailing sand to be as an environmentally friendly alternative to river sand in
construction applications.” Journal of Cleaner Production, 205, 1114–1126.
Yang, R., Yu, R., Shui, Z., Gao, X., Xiao, X., Fan, D., Chen, Z., Cai, J., Li, X., and He,
Y. (2020). “Feasibility analysis of treating recycled rock dust as an environmentally
friendly alternative material in ultra-high performance concrete (UHPC).” Journal of
Cleaner Production, 258, 120673.
Zhang, H., Ji, T., Zeng, X., Yang, Z., Lin, X., and Liang, Y. (2018). “Mechanical behavior
of ultra-high performance concrete (UHPC) using recycled fine aggregate cured under
different conditions and the mechanism based on integrated microstructural
parameters.” Construction and Building Materials, 192, 489–507.
Zhang, W., Gu, X., Qiu, J., Liu, J., Zhao, Y., and Li, X. (2020a). “Effects of iron ore tailings
on the compressive strength and permeability of ultra-high performance concrete.”
Construction and Building Materials, 260, 119917.
Zhang, Y., Shen, W., Wu, M., Shen, B., Li, M., Xu, G., Zhang, B., Ding, Q., and Chen, X.
(2020b). “Experimental study on the utilization of copper tailing as micronized sand to
prepare high performance concrete.” Construction and Building Materials, 244, 118312.
Zhang, Z., Zhang, Z., Yin, S., and Yu, L. (2020c). “Utilization of iron tailings sand as an
environmentally friendly alternative to natural river sand in high-strength concrete:
shrinkage characterization and mitigation strategies.” Materials, 13(24), 1–15.
Zhao, S., Fan, J., and Sun, W. (2014). “Utilization of iron ore tailings as fine aggregate in
ultra-high performance concrete.” Construction and Building Materials, 50, 540–548.
Zhu, Z., Li, B., and Zhou, M. (2015). “The influences of iron ore tailings as fine aggregate
on the strength of ultra-high performance concrete.” Advances in Materials Science and
Engineering, 2015, 412878.

147
CHAPTER
6
Concluding remarks

6.1 Main findings

The main conclusions, that can be drawn from the scope of this thesis and that are related
to the research objectives mentioned in Chapter 1, are summarized below:

Chapter 2: The experimental work for Chapter 2 was planned to address Research objectives
1 and 2. At a given water to binder (w/b) ratio, very-low-C3A cement (VLAC) imparts better
flowability to UHPC than moderate-C3A cement (MAC). For a given slump flow, VLAC
based UHPC exhibits higher compressive strength than that of MAC based UHPC. Hence,
VLAC was selected as the primary binder for preparing UHPC mixes in the rest of the
project. Based on the scope of work planned for Chapter 2, the optimum silica fume to binder
ratio and binder to aggregate ratio that impart the best performance to the control UHPC,
in terms of compressive strength, are around 0.2 (=20/100) and 1.14, respectively. At a silica
fume to binder ratio of 0.2, the compressive strength, elastic modulus, splitting tensile
strength, and flexural strength of UHPC generally reduce with the replacement of cement
with GGBS. TG-DTG results support this conclusion. However, the 28-day compressive
strength of more than 150 MPa can be achieved up to 60% replacement, without employing
special curing or incorporating fibres. Comparison of the elastic modulus, splitting tensile
strength, and flexural strength test results with the predictions based on several standard
models for NSC and HSC reveals that separate models, especially for predicting splitting
tensile strength and flexural strength, are needed for UHPC. The water absorption and initial
rate of absorption of UHPC, and the corrosion risk of rebar embedded in UHPC reduce with
the increase of cement replacement with GGBS. Up to 60% replacement, UHPC exhibits
carbonation resistance like that of the UHPC without GGBS. In this study, visible shrinkage
cracks were not detected on the surface of any of the specimens made with GGBS.

Chapter 3: The scope of work for this chapter was planned to meet Research objective 3.
The inclusion of GGBS substituting 30% of the VLAC in UHPFRC reduces the compressive
strength. However, the 28-day flexural strength of UHPFRC can be increased by 65.6%
incorporating 30% GGBS. A more pronounced deflection hardening part in the load –
Chapter 6

deflection curve (under flexure) can be achieved for UHPFRC with GGBS than for the
composite without GGBS. The water absorption, and the initial and secondary sorptivities
of UHPFRC with GGBS are lower than those of the UHPFRC without GGBS. The surface
chloride concentration of UHPFRC with 30% GGBS was lower than that of the UHPFRC
without GGBS, after 216 d of exposure to cyclic drying and wetting with 10% NaCl solution.
Corrosion of fibres was observed only at the surface. For any of the mixes, the corrosion was
not observed to progress inside the specimens. The levels of chloride penetration resistance
of VLAC based UHPFRCs without and with GGBS can be ‘very high’ and ‘extremely high’,
respectively. X-ray diffractometry (XRD) confirms the formation of more Friedel’s salt in
the mix with GGBS. Visible shrinkage cracks were not found on the surface of any of the
UHPFRC specimens made with GGBS.

Chapter 4: The scope of work for Chapter 4 was outlined to address Research objectives 4
and 5. High-volume incorporation of Class-F fly ash reduces the compressive strength,
elastic modulus, splitting tensile strength, and flexural strength of UHPC. The presence of
more unreacted fly ash particles in UHPC with higher fly ash content, observed from SEM
images, is in agreement with the lower strength of such cementitious mix. However, 28-day
compressive strengths of more than 150 MPa and 100 MPa can be achieved with up to 40%
and 70% replacements, respectively, without employing any special curing or fibres. A
model has been developed to predict the compressive strength of UHPC incorporating fly
ash from the contents and chemical compositions of its binders. Relationships have also
been developed to predict elastic modulus, splitting tensile strength, and flexural strength of
standard-cured UHPC with/without fly ash from the compressive strength. UHPCs made
with up to 60% cement replacement with fly ash exhibit durability properties comparable to
that of UHPC without fly ash, in terms of water absorption, initial rate of water absorption,
and corrosion risk. The depth of carbonation remains below the detection limit of 0.5 mm,
up to 70% replacement. Visible shrinkage cracks were not found on the surface of any
specimen made with fly ash.

Chapter 5: The experimental plan for Chapter 5 was set to meet Research objective 6.
UHPCs, made with up to 80% replacement of quartz sand by gold mine tailings, sourced
from a mine in WA (exact name of the mine is not disclosed due to confidentiality reasons),
exhibit compressive strengths comparable to or higher than that of the UHPC with 100%
quartz sand. 28-day strength greater than 120 MPa is achievable up to 100% replacement.
The water absorption and the initial rate of absorption values of UHPCs with tailings are
generally lower than those of the UHPC without tailings, and the carbonation resistances of
UHPCs with tailings are similar to that of conventional UHPC. The leachability of toxic

149
Chapter 6

metals from UHPC with up to 100% tailings content lies well below the regulatory
thresholds. Visible shrinkage cracks were not detected in this study on the surface of any of
the specimens made gold mine tailings.

In general, UHPCs incorporating industrial by-products like GGBS, Class-F fly ash, and
gold-mine tailings have good potential as construction materials. Nevertheless, few
limitations of the by-products should be taken into consideration before their large-scale use.
For example, the properties of GGBS and fly ash sometimes vary between sources. Making
a sufficient number of trial mixes with these by-products to achieve desired UHPC properties
can partially solve this problem. The gold mine tailings used in this study do not necessarily
represent other gold tailings from different mines. Hence, detailed investigation is necessary
before using any other type of gold mine tailings. The possibility of toxic leaching, acid
generation, etc. from UHPC made with the gold mine tailings should also be examined.

6.2 Future research recommendations

Some investigations that can be undertaken in the future are highlighted below:

1. To reduce both the CO2 footprint and cost of UHPC, combined incorporation of high-
volume GGBS or Class-F fly ash and gold mine tailings, respectively as SCM and
aggregate, in UHPC can be considered. A study can therefore be carried out addressing
the mechanical, durability, and microstructure properties of UHPC incorporating both
GGBS/fly ash and tailings.
2. The effects of using unprocessed fly ash on the mechanical, durability, and
microstructure properties of UHPC can be explored.
3. Long-term half-cell potential test can be carried out to examine the corrosion behaviour
of rebar embedded in UHPC with a very high fly ash content like 70%.
4. To instil more confidence in the use of gold mine tailings as aggregate in UHPC, tailings
from several other mines in WA can be characterized, and the mechanical and durability
properties of UHPCs incorporating those tailings can be examined. The possibility of
toxic leaching, acid generation, etc. from the UHPCs made with these tailings should
also be assessed.
5. A study can be undertaken to investigate the possibility of using finer gold mine tailings,
with median particle sizes comparable to that of cement or silica fume, as supplementary
binders in UHPC.
6. Direct tensile properties and ductility properties of UHPFRCs incorporating GGBS, fly
ash, and tailings can be investigated.

150
Chapter 6

7. To further verify the self-compacting behaviour of the UHPFRC mixes presented in


Chapter 3, tests like T500, L-box, U-box, etc. can be carried out.
8. Durability properties such as sulphate attack resistance, frost damage resistance,
resistance to alkali-silica reaction, etc. can be evaluated for the ECO-UHPCs developed
in this study.
9. The structural performances of beams, columns, walls, and connections built with the
ECO-UHPCs can be evaluated. The possibility of using the ECO-UHPCs as repair
materials can be examined.
10. Detailed comparative life cycle assessment (LCA), such as cradle-to-gate or cradle-to-
grave LCA, can be carried out for structures built with conventional UHPC and UHPCs
incorporating GGBS, fly ash, and/or tailings.

151
APPENDIX A
(Conference paper related to Chapter 2)

Shrinkage of UHPC incorporating very-low-


C3A cement and ground granulated blast-
furnace slag

Tanvir Ahmeda, Mohamed Elchalakania, Ali Karrecha, Waleed Nawaza, Huiyuan Liua
a
Department of Civil, Environmental and Mining Engineering, Faculty of Engineering, Computing and Mathematical
Sciences, The University of Western Australia, 35 Stirling Highway, Crawley, WA 6009, Australia

Abstract

Ultra-high performance concrete (UHPC), owing to containing very high cement content,
usually exhibits high shrinkage compared to normal strength concrete (NSC). Shrinkage
may lead to the formation of cracks. Excessive shrinkage cracks may impair the mechanical
performance of concrete and its durability. In this study, the influence of high-volume
cement substitution by ground granulated blast-furnace slag (GGBS) on the shrinkage
behaviour of very-low-C3A cement based UHPC has been investigated. The results suggest
that the 7-day and 28-day shrinkages of UHPC can be reduced by up to 26.6% and 31.6%,
respectively, by substituting GGBS for 60% of very-low-C3A cement. The compressive
strength of UHPC has been observed to reduce with increasing replacement of very-low-
C3A cement by GGBS. Nevertheless, UHPC incorporating up to 60% GGBS has been
observed to attain 28-day strength greater than 150 MPa, without employing any special
curing or fibres.

Keywords: autogenous shrinkage, drying shrinkage, ettringite, supplementary cementitious


material

A.1 Introduction

Ultra-high performance concrete (UHPC) is characterized by its extraordinary mechanical


and durability properties which far surpass those of normal-strength concrete (NSC). In
order to achieve such high performance, a large amount of cement (800–1200 kg/m3) and a
very low water to binder ratio (w/b < 0.2) are used. When a cementitious mix is designed
with very low w/b ratio and when its binder system consists of high-volume of cement [i.e.,

Concrete 2021, The 30th Biennial National Conference of the Concrete Institute of Australia, 138
Appendix A

the water to cement ratio (w/c) is also very low], the internal moisture content of the
cementitious mix is not sufficient to facilitate complete hydration of all the cement particles.
Therefore, cementitious mix like UHPC is susceptible to substantial self-desiccation and
thus to autogenous shrinkage (Ghafari et al. 2016; Wu et al. 2017). Internal stresses
generated by high shrinkage strain may result in microcracks, which may eventually affect
the mechanical and durability performances of UHPC. Different approaches were adopted
by researchers to address this concern. Xie et al. (2018) found that the use of 3% shrinkage
reducing admixture and the replacement of 50% of the mixing water by crushed ice reduced
the 90-day shrinkage of UHPC by 65.7% and 20.4%, respectively. Reduction in the
autogenous shrinkage of ultra-high performance fibre reinforced concrete (UHPFRC)
through the use of shrinkage reducing admixture was also confirmed by different studies
(Yoo et al. 2014, 2015, 2016). Fang et al. (2020) observed that UHPFRC mixes
incorporating steel fibres exhibited lower shrinkage than plain UHPC. Moreover, shrinkage
of UHPFRC was observed to reduce with the increase of fibre volume fraction; and hooked
fibres were found to induce more pronounced reduction than the straight ones (Fang et al.
2020). Soliman and Nehdi (2013) found the use of partially hydrated cementitious materials
and super absorbent polymer to be effective in reducing the early age autogenous shrinkage
of UHPC. Liu et al. (2017) observed that inclusion of 0.24% super absorbent polymer (by
mass of binder) reduced the 7-day autogenous shrinkage of UHPFRC by around 50%.

Some researchers also investigated the influence of different supplementary cementitious


materials (SCMs) on the autogenous shrinkage of UHPFRC. Ghafari et al. (2016) found
that cement – fly ash and cement – GGBS binder systems imparted lower autogenous
shrinkage to UHPFRC than cement – silica fume binder system. Liu et al. (2017) observed
that the replacement of cement by fly ash or GGBS decreases the autogenous shrinkage of
UHPFRC. Staquet and Espion (2004) found the binder system incorporating very-low-C3A
cement (2.1% C3A), silica fume, and metakaolin to be more effective in reducing the
autogenous shrinkage of UHPFRC than the binder system containing moderate-C3A cement
(8.1% C3A) and silica fume.

In this study, the influence of replacing very-low-C3A cement (with 1.1% C3A content) by
high-volume GGBS (up to 60%) on the shrinkage of UHPC was investigated. The 7-day and
28-day compressive strengths of the prepared UHPC mixes were also measured to
understand the influence of high-volume inclusion of GGBS in very-low-C3A cement based
UHPC.

153
Appendix A

A.2 Experimental program

A.2.1 Materials

Binders used in this study include ASTM C150 / C150M-16 (2016) Type I Portland cement
with a very low C3A content (1.1%), silica fume conforming to AS/NZS 3582.3 (2016), and
GGBS conforming to AS/NZS 3582.2 (2016). Table A.1 presents the chemical
compositions of the binders. Quartz sand used in this study had a median diameter (d50) of
237.3 µm. A polycarboxylate based high-range water reducer (HRWR) and a defoamer with
36% and 5.5% solid contents, respectively, were used as chemical admixtures.

A.2.2 Mix proportions

The mix proportions of the UHPC mixes are presented in Table A.2. The w/b ratio of all
the mixes was kept constant at 0.166, and the ratio of silica fume to the total content of
cement and GGBS was kept constant at 0.25. The numbers next to ‘C’ and ‘S’ in the mix
title of a UHPC mix represent respectively the percentage contents of very-low-C3A cement
and GGBS with respect to the total content of cement and GGBS.

A.2.3 Mixing and specimen preparation

The mixing procedure adopted for the preparation of the UHPC mixes can be described by
the flow chart shown in Figure A.1. Fresh UHPCs were poured into 25 × 25 × 285 mm
prism moulds and 50 × 50 × 50 mm cube moulds to prepare specimens for shrinkage and
compressive strength tests, respectively. The freshly moulded specimens were vibrated for a
minute and then moved to a curing room (95% relative humidity, 21 °C temperature). After
24 hours, the specimens were demoulded. The shrinkage test specimens were then stored at
a temperature of 25 °C and a relative humidity < 50%. The compressive strength test
specimens, on the other hand, were fully submerged in water (21 °C). Both the shrinkage
and compressive strength test specimens were cured in the aforementioned conditions until
testing at 7 and 28 days.

A.2.4 Test methods

The total shrinkage strains [including both drying and autogenous shrinkage strains (Xie et
al. 2018)] of UHPCs were measured in accordance with the guidelines of AS 2350.13 (2016)
using 25 × 25 × 285 mm prisms. Three prisms were tested to measure the average shrinkage

154
Appendix A

strain of each mix at a certain age. A comparator was used for measuring the length changes
of the prisms (Figure A.2). Equation A.1 was used for strain calculation.

𝐿𝑖 − 𝐿𝑓
𝜀= × 106 (µm/m) (A.1)
𝐿

where, Li = comparator reading of specimen taken on demoulding minus comparator


reading of reference bar (mm), Lf = comparator reading of specimen at a certain age minus
comparator reading of reference bar at the same age (mm), and L = nominal gauge length
(mm).

The compressive strength tests were performed on 50 × 50 × 50 mm cubes in accordance


with the guidelines of ASTM C109 / C109M-16a (2016). Scanning electron microscopy
(SEM) was conducted on crushed pieces of C100 and C70-S30 cubes, subjected to 28-days
of curing in water, using a field emission scanning electron microscope. Energy dispersive
spectrum (EDS) analysis was also performed for C100 and C70-S30 to detect certain reaction
products.

Table A.1. Chemical compositions of binders


Material Oxides (mass %)

Al2O3 CaO Fe2O3 K2O MgO MnO Na2O P2O5 SiO2 SO3 TiO2 Loss on
ignition
Very-low- 3.18 64.48 4.34 0.35 1.46 0.12 0.14 0.228 21.4 2.93 0.192 1.15
C3A
cement
Silica 0.13 0.33 0.07 0.54 0.71 0.008 0.29 0.129 94.6 0.07 0.005 3.1
fume
GGBS 13.04 42.8 0.81 0.3 5.46 0.17 0.28 0.02 31.14 4.02 0.56 1.4
Very-low-C3A cement phases: C3S (alite) = 63.7%, C2S (belite) = 13.4%, C3A (tricalcium aluminate) = 1.1%, C4AF
(ferrite) = 13.2%

Table A.2. Mix proportions


Mix title Constituents w/b
Very-low-C3A cement Silica fume GGBS Sand Water HRWR Defoamer
C100 1 0.25 - 1.1 0.208 0.016 0.0014 0.166
C85-S15 0.85 0.25 0.15 1.1 0.208 0.016 0.0014 0.166
C70-S30 0.7 0.25 0.3 1.1 0.208 0.016 0.0014 0.166
C55-S45 0.55 0.25 0.45 1.1 0.208 0.016 0.0014 0.166
C40-S60 0.4 0.25 0.6 1.1 0.208 0.016 0.0014 0.166

155
Appendix A

Figure A.1. Mixing procedure

Figure A.2. Length change measurement using comparator

A.3 Results and discussion

A.3.1 Shrinkage

Figure A.3 shows the 7-day and 28-day total shrinkages of the UHPCs incorporating
different GGBS to very-low-C3A cement ratios. Irrespective of age, the total shrinkage
strains of the UHPCs containing GGBS were generally lower than that of C100. Lower total
shrinkages of the UHPCs containing GGBS can be attributed to the lower hydraulic
reactivity of GGBS than that of cement (Liu et al. 2017). Lower reactivity of a binder system
is likely to reduce the autogenous shrinkage of the cementitious mix incorporating the binder
system (Liu et al. 2017). It is important to note that the drying shrinkage of UHPC is usually
very low (Xie et al. 2018). Therefore, the autogenous shrinkage of UHPC can be considered
as the major contributor to its total shrinkage. Again, higher SO3 content in Table A.1 for
GGBS than for cement is indicative of higher gypsum content in the former (Aly and
Sanjayan 2008). GGBS, in presence of gypsum, facilitates the formation of early age

156
Appendix A

ettringite. On the other hand, very-low-C3A cement results in a substantially low amount of
ettringite at an early age (Mardani-Aghabaglou et al. 2017). Ettringite formation at an early
age can induce expansive stress (Matschei et al. 2005). High total shrinkage strain of each
UHPC at 7-day, shown in Figure A.3, indicates that its shrinkage stress was much higher
than its expansive stress at that age (if any). Nevertheless, it is possible that higher expansive
stresses, induced by higher ettringite formation in the UHPCs incorporating GGBS,
somewhat helped the UHPCs attain lower shrinkage strains than the UHPC without GGBS.
The total shrinkage strains of all the UHPCs were observed to increase with time. It is,
however, worth noticing that the increments in the 28-day shrinkage strains of C55-S45 and
C40-S60 with respect to their 7-day strains were less pronounced compared to the other
mixes. This may be because of the substantial reduction of the cement contents in C55-S45
and C40-S60 (dilution effect) (Elchalakani et al. 2017). The 7-day and 28-day shrinkage
strains of C100 were 874.3 μm/m and 1173.5 μm/m, respectively. The 7-day shrinkage
strains of C85-S15, C70-S30, C55-S45, and C40-S60 reduced by 16%, 12%, 28%, and 26.6%
with respect to the 7-day strain of C100. At 28 days, the shrinkage strains of C85-S15, C70-
S30, C55-S45, and C40-S60 were observed to decrease by 17%, 5.5%, 31.4%, and 31.6%
with respect to that of C100.

A.3.2 Compressive strength

Figure A.4 presents the 7-day and 28-day compressive strengths of the UHPC mixes
incorporating different GGBS to very-low-C3A cement ratios. The results indicate that the
compressive strengths of the UHPCs generally reduced with increasing replacement of very-
low-C3A cement by GGBS. Such results imply lower hydraulic reactivity of GGBS than
very-low-C3A cement. It is, furthermore, important to note that owing to the low w/b ratio
of UHPC and the presence of silica fume, portlandite content in UHPC is typically very low
compared to that in NSC (Abdulkareem et al. 2018; Ahmed et al. 2021). Low portlandite
content in UHPC is likely to limit the pozzolanic activity of GGBS. The 7-day and 28-day
strengths of C100 were 148.7 MPa and 181.2 MPa, respectively. The 7-day strengths of C85-
S15, C70-S30, C55-S45, and C40-S60 decreased by 2.6%, 8.4%, 9.3%, and 14.4% with
respect to the 7-day strength of C100. At 28 days, the strengths of C85-S15, C70-S30, C55-
S45, and C40-S60 were observed to reduce by 1.3%, 6.6%, 10.5%, and 16.1% with respect
to that of C100. Although GGBS inclusion reduced the strength of UHPC, all the mixes
prepared in this study were observed to achieve 28-day strengths greater than 150 MPa
without employing any special curing or fibres.

157
Appendix A

1400
7-day 28-day
1200

Shrinkage strain (μm/m)


1000

800

600

400

200

0
C100 C85-S15 C70-S30 C55-S45 C40-S60

Figure A.3. Total shrinkage strains of UHPCs at 7 and 28 days

200
7-day 28-day
Compressive strength (MPa)

160

120

80

40

0
C100 C85-S15 C70-S30 C55-S45 C40-S60

Figure A.4. Compressive strengths of UHPCs at 7 and 28 days

A.3.3 SEM analysis

The scanning electron micrographs of crushed pieces of C100 and C70-S30 are shown in
Figure A.5. The results obtained from EDS analysis, performed to detect certain reaction
products, are also presented in Figure A.5. Calcium silicate hydrate (C-S-H) was found to
be the main reaction product in the SEM images of both C100 and C70-S30. Traces of
portlandite (CH) crystals were also detected in C100. Ettringite (AFt) crystals were observed
in the vicinity of partially reacted GGBS particles in C70-S30. More pronounced micro-
cracking was observed in the matrix of C100 than in the C70-S30 matrix. This observation
is in agreement with the shrinkage results obtained for C100 and C70-S30. However, it is
worth mentioning that some of the micro-cracks possibly formed during the crushing of
UHPC cubes for sample preparation as well.

158
Appendix A

Ca% Si% Al% Si/Ca Al/Ca


Spot 1
42.4 3.19 0.21 0.08 0.00

(a) SEM images and EDS results of C100

Ca% Si% Al% Si/Ca Al/Ca


Spot 2
20.7 3.14 8.12 0.15 0.39

(b) SEM images and EDS results of C70-S30

Figure A.5. SEM images and EDS results of C100 and C70-S30 (M = C-S-H rich matrix, P =
portlandite, S = GGBS, R = reaction products, G = gypsum, E = ettringite)

A.4 Conclusions

Based on the results obtained in this study, the following conclusions can be drawn:

1. The total shrinkage strain of UHPC reduces when very-low-C3A cement is replaced by
GGBS. At 7 days, a reduction in the shrinkage strain of 26.6% is attainable for 60%
replacement. A reduction of 31.6% is attainable at 28 days.

159
Appendix A

2. At a w/b ratio of 0.166, replacement of very-low-C3A cement by GGBS reduces the


compressive strength of UHPC.
3. At 7 and 28 days, UHPC with 40/60 very-low-C3A cement to GGBS ratio exhibits
respectively 14.4% and 16.1% lower compressive strengths, with respect to those of the
UHPC without GGBS.

Acknowledgements

The authors acknowledge the facilities and support provided by the Centre for Microscopy,
Characterization & Analysis (CMCA), The University of Western Australia to perform the
SEM and EDS analyses. The study was conducted while the first author was in receipt of a
University Postgraduate Award and the Australian Government Research Training
Program Scholarship at The University of Western Australia.

References
Abdulkareem, O. M., Ben Fraj, A., Bouasker, M., and Khelidj, A. (2018). “Effect of
chemical and thermal activation on the microstructural and mechanical properties of
more sustainable UHPC.” Construction and Building Materials, 169, 567–577.
Ahmed, T., Elchalakani, M., Karrech, A., Dong, M., Mohamed Ali, M. S., and Yang, H.
(2021). “ECO-UHPC with High-Volume Class-F Fly Ash: New Insight into Mechanical
and Durability Properties.” Journal of Materials in Civil Engineering, 33(7), 0003726.
Aly, T., and Sanjayan, J. G. (2008). “Factors contributing to early age shrinkage cracking
of slag concretes subjected to 7-days moist curing.” Materials and Structures/Materiaux et
Constructions, 41(4), 633–642.
AS/NZS 3582.2. (2016). Supplementary cementitious materials for use with portland and blended
cement - Slag—Ground granulated blastfurnace.
AS/NZS 3582.3. (2016). Supplementary cementitious materials for use with portland and blended
cement - amorphous silica.
AS 2350.13. (2016). Methods of testing portland, blended And masonry cements - Determination of
drying shrinkage of cement mortars.
ASTM C150 / C150M-16. (2016). Standard specification for portland cement. West
Conshohocken, PA.
ASTMC109 / C109M-16a. (2016). Standard test method for compressive strength of hydraulic
cement mortars (using 2-in. or [50-mm] cube specimens). West Conshohocken, PA.
Elchalakani, M., Basarir, H., and Karrech, A. (2017). “Green concrete with high-volume fly
ash and slag with recycled aggregate and recycled water to build future sustainable
cities.” Journal of Materials in Civil Engineering, 29(2), 04016219.
Fang, C., Ali, M., Xie, T., Visintin, P., and Sheikh, A. H. (2020). “The influence of steel
fibre properties on the shrinkage of ultra-high performance fibre reinforced concrete.”
Construction and Building Materials, 242.
Ghafari, E., Ghahari, S. A., Costa, H., Júlio, E., Portugal, A., and Durães, L. (2016). “Effect
of supplementary cementitious materials on autogenous shrinkage of ultra-high

160
Appendix A

performance concrete.” Construction and Building Materials, 127, 43–48.


Liu, J., Ou, Z., Mo, J., Wang, Y., and Wu, H. (2017). “The effect of SCMs and SAP on the
autogenous shrinkage and hydration process of RPC.” Construction and Building
Materials, 155, 239–249.
Mardani-Aghabaglou, A., Felekoğlu, B., and Ramyar, K. (2017). “Effect of cement C3A
content on properties of cementitious systems containing high-range water-reducing
admixture.” Journal of Materials in Civil Engineering, 29(8), 1–12.
Matschei, T., Bellmann, F., and Stark, J. (2005). “Hydration behaviour of sulphate-activated
slag cements.” Advances in Cement Research, 17(4), 167–178.
Soliman, A. M., and Nehdi, M. L. (2013). “Effect of partially hydrated cementitious
materials and superabsorbent polymer on early-age shrinkage of UHPC.” Construction
and Building Materials, 41, 270–275.
Staquet, S., and Espion, B. (2004). “Early-age autogenous shrinkage of UHPC incorporating
very fine fly ash or metakaolin in replacement of silica fume.” International Symposium
on UHPC, M. Schmidt, E. Fehling, and C. Geisenhanslüke, eds., Kassel university press
GmbH, Kassel, 587–599.
Wu, L., Farzadnia, N., Shi, C., Zhang, Z., and Wang, H. (2017). “Autogenous shrinkage
of high performance concrete: A review.” Construction and Building Materials, 149, 62–
75.
Xie, T., Fang, C., Mohamad Ali, M. S., and Visintin, P. (2018). “Characterizations of
autogenous and drying shrinkage of ultra-high performance concrete (UHPC): An
experimental study.” Cement and Concrete Composites, 91, 156–173.
Yoo, D. Y., Banthia, N., and Yoon, Y. S. (2015). “Effectiveness of shrinkage-reducing
admixture in reducing autogenous shrinkage stress of ultra-high-performance fiber-
reinforced concrete.” Cement and Concrete Composites, 64, 27–36.
Yoo, D. Y., Banthia, N., and Yoon, Y. S. (2016). “Mitigating early-age cracking in thin
UHPFRC precast concrete products using shrinkage-reducing admixtures.” PCI Journal,
61(1), 39–50.
Yoo, D. Y., Park, J. J., Kim, S. W., and Yoon, Y. S. (2014). “Combined effect of expansive
and shrinkage-reducing admixtures on the properties of ultra high performance fiber-
reinforced concrete.” Journal of Composite Materials, 48(16), 1981–1991.

161
APPENDIX B
(Data related to Chapter 4)

Table B.1. Data used from this study for Figures 4.12, 4.13, 4.14, and 4.15

Sl. no. Mix UHPC type Fly ash Reactivity moduli b/a w/b fc Ec fr fsp
% MPa MPa MPa MPa
1 F0 Conventional 0 HM = 2.54, SM = 1.15 0.16 198.9 53145 27.9 11.7
101.03, AM = 0.99
2 F20 With Class-F 20 HM = 2.13, SM = 179.2 49671 24.7 10.1
fly ash 100.86, AM = 1.35
3 F40 40 HM = 1.71, SM = 152.6 44001 23.5 8.6
100.69, AM = 1.71
4 F60 60 HM = 1.29, SM = 123.5 38266 18.3 7.9
100.52, AM = 2.07
5 F70 70 HM = 1.08, SM = 106.9 34378 15.6 6.4
100.43, AM = 2.25
6 W0 Conventional 0 HM = 2.54, SM = 0.19 172.1 - 23.6 9.8
101.03, AM = 0.99
7 W60 With Class-F 60 HM = 1.29, SM = 99.1 - 13.6 4.8
fly ash 100.52, AM = 2.07

Table B.2. Data collected from studies on standard-cured UHPC with cement – silica fume or
cement – silica fume – fly ash binder system for Figures 4.12, 4.13, 4.14, and 4.15

Sl. no. Reference UHPC type Fly ash Reactivity moduli b/a w/b fc Ec fr fsp
% MPa MPa MPa MPa
8 Chen et al. Conventional* 0 HM = 2.97, SM = 0.91 0.2 114 - 16.3 -
(2018) 13.18, AM = 1.51
9 With Class-F 10 HM = 2.78, SM = 115.8 - 16.7 -
fly ash 13.12, AM = 3.28
10 20 HM = 2.49, SM = 126 - 20 -
13.05, AM = 5.06
11 30 HM = 2.25, SM = 121.3 - 18.4 -
12.99, AM = 6.84
12 Liu et al. With Class-F 21 HM = 2.91, SM = 1.1 0.18 108.5 - 10.4 -
(2018) fly ash & river 38.46, AM = 2.16
sand
13 Song et al. With Class-F 18 HM = 3.03, SM = 0.82 0.2 105 - 9.4 -
(2018) fly ash & river 15.3, AM = 3.10
sand
14 Huang et Conventional 0 HM = 3.16, SM = 1.18 0.2 119 - 20 -
al. (2019) 24.67, AM = 1.02
15 Mueller et With Class-F 25 - 0.74† 0.16 147.2 49700 - -
al. (2016) fly ash
16 31 0.17 135.7 49900 - -
17 Wang et With Class-F 29 HM = 1.88, SM = 1 0.18 110 - - -
al. (2016) fly ash 7.32, AM = 2.07
18 Wu et al. With river sand 0 HM = 2.97, SM = 1 0.18 107 - 21 -
(2016b) 18.26, AM = 1.25
19 0 HM = 2.8, SM = 105 - 22 -
26.08, AM = 1.18
20 0 HM = 2.64, SM = 116.5 - 24.3 -
33.91, AM = 1.11

Journal of Materials in Civil Engineering, 33(7), 0003726


(https://doi.org/10.1061/(asce)mt.1943-5533.0003726)
Appendix B

21 0 HM = 2.48, SM = 112.6 - 24.1 -


41.74, AM = 1.04
22 Zhu et al. With natural 0 HM = 1.9, SM = 1 0.2 107 - 15.2 -
(2016) sand 21.31, AM = 4.33
23 Li et al. With Class-F 9 HM = 2.34, SM = 1 0.16 125.6 - 22.2 -
(2015) fly ash 26.30, AM = 2.05
24 9 HM = 2.34, SM = 0.17 120.8 - 18.5 -
26.30, AM = 2.06
25 Mounanga Conventional 0 HM = 2.6, SM = 1.14 0.16 155 - 19.2 -
et al. 5.02, AM = 3.40
(2012)
26 0 HM = 2.6, SM = 150 - 18.6 -
5.02, AM = 3.40
27 0 HM = 2.6, SM = 140 - 18.3 -
5.02, AM = 3.40
28 0 HM = 2.73, SM = 1.02 170 - 20.6 -
4.56, AM = 2.89
29 0 HM = 2.73, SM = 175 - 22.3 -
4.56, AM = 2.89
30 Peng et al. Conventional 0 HM = 2.43, SM = 0.91 0.18 113.1 - - -
(2010) 9.56, AM = 2.16
31 With Class-F 12 HM = 2.22, SM = 97.8 - - -
fly ash 9.47, AM = 2.46
32 18 HM = 2.11, SM = 92 - - -
9.43, AM = 2.60
33 24 HM = 2.00, SM = 91.5 - - -
9.38, AM = 2.75
34 35 HM = 1.79, SM = 85.8 - - -
9.29, AM = 3.04
35 47 HM = 1.57, SM = 88.3 - - -
9.20, AM = 3.33
36 Han With fly ash 20 - 1.14 0.24 100 - 11.6 6.75
(2017a) (type not
reported)
*Conventional UHPC contains cement, silica fume, quartz sand, quartz powder or crushed quartz, water, and chemical
admixtures

by volume

Table B.3. Data collected from studies in the literature for Figures 4.13, 4.14, and 4.15

Sl. Reference UHPC type fc Ec fr fsp


no.
MPa MPa MPa MPa
37 Abbas et al. (2015) Conventional 151 42600 - 9.4
38 Abdulkareem et Conventional 140 - 19 8.7
al. (2018)
39 With GGBS 141 - 18 8.5
40 125 - 17.5 7.7
41 107 - 14 8.2
42 Ahmad et al. Conventional 130 38000 - -
(2014)
43 Alsalman et al. With river sand and/or fly ash as aggregate 136.4 43400 - -
(2017b)
44 135 37600 - -
45 124.1 37200 - -
46 135.5 36900 - -

163
Appendix B

47 Courtial et al. Conventional 155 - 19 -


(2013)
48 150 - 19 -
49 140 - 19 -
50 175 - 22 -
51 170 - 21 -
52 Cwirzen et al. Conventional 146 41609 14 -
(2008)
53 127 40316 12 -
54 142 42944 15 -
55 152 41575 15 -
56 137 45365 15 -
57 129 40932 13 -
58 147 41209 15 -
59 152 43210 15 -
60 Dhundasi & Conventional 144 - 22 10.2
Khadiranaikar
(2019)
61 Gesoglu et al. Conventional 137 39300 - 8.4
(2016b)
62 Gesoglu et al. Conventional 114 - - 7.1
(2016a)
63 121 - - 7.95
64 With nano-silica 118 - - 8.12
65 123.5 - - 8.24
66 124 - - 7.95
67 131.7 - - 8.97
68 137 - - 9.58
69 115 - - 7.5
70 134 - - 9.25
71 127.5 - - 8.45
72 Goaiz et al. (2018) Conventional 73.41 - - 6.26
73 Gupta (2016) With river sand 110 - 13.2 -
74 With river sand, ultra-fine GGBS 96.5 - 12.1 -
75 130 - 16.4 -
76 107.5 - 13.7 -
77 102 - 12.33 -
78 136 - 17 -
79 116 - 14.6 -
80 Han (2017) With nano-zirconia 115.5 - 13.4 6
81 115 - 14 6.625
82 114.85 - 15.6 8.6
83 114.8 - 14.5 6.8
84 Hannawi et al. Conventional 141 45300 - -
(2016)
85 Holschemacher et Conventional 148 47100 - 12.2
al. (2004)
86 Huang et al. With rice husk ash 121.5 - 22.5 -
(2017)
87 125 - 19.7 -
88 130 - 21.2 -
89 136.7 - 19.5 -

164
Appendix B

90 130.5 - 21 -
91 Huang et al. Conventional (contains very fine quartz sand with 87 - 17.5 -
(2019) maximum aggregate size = 0.38 mm)
92 90 - 16.3 -
93 87.5 - 17.3 -
94 93 - 16.2 -
95 90.5 - 16.2 -
96 87.5 - 15.5 -
97 85 - 15 -
98 80.5 - 16.2 -
99 85 - 14.8 -
100 95 - 15.5 -
101 98 - 14.5 -
102 99.6 - 13.7 -
103 100 - 17 -
104 104 - 18.3 -
105 100.5 - 19.5 -
106 98 - 20 -
107 99.8 - 15.5 -
108 85 - 14.8 -
109 94 - 15 -
110 95 - 13.7 -
111 98 - 17.8 -
112 107 - 17.1 -
113 109 - 17.8 -
114 109.5 - 19.1 -
115 110 - 17.5 -
116 105 - 15.9 -
117 100 - 18 -
118 90 - 15.8 -
119 95 - 19.1 -
120 100 - 18.7 -
121 98 - 19.1 -
122 85 - 17.2 -
123 90 - 18 -
124 96 - 19.5 -
125 95 - 17.8 -
126 94 - 18.7 -
127 Karihaloo & Conventional 100 - - 5.2
Vriese (1999)
128 85 - - 4.5
129 100 - - 6.4
130 95 - - 6
131 Li et al. (2015) With nano-silica and/or nano-limestone 128.7 - 25.6 -
132 130.3 - 25.8 -
133 129.4 - 25 -
134 130.6 - 24.5 -
135 124.6 - 23.1 -
136 126.5 - 23.7 -

165
Appendix B

137 126 - 22.9 -


138 125.7 - 22.3 -
139 130.3 - 25.3 -
140 134.7 - 25.8 -
141 138.2 - 26.1 -
142 136.4 - 24.7 -
143 122.4 - 21.4 -
144 130.5 - 22.3 -
145 133.1 - 22.7 -
146 130.3 - 22.1 -
147 127.4 - 22.2 -
148 122.5 - 21 -
149 115.3 - 19 -
150 125.6 - 21.6 -
151 116.7 - 20.9 -
152 113.4 - 19 -
153 109.5 - 17.9 -
154 120.7 - 21.4 -
155 116.6 - 19.9 -
156 111.2 - 17 -
157 121.4 - 20.3 -
158 Liu et al. (2018) With river sand, porous pumice (internal curing agent), 107.5 - 16.4 -
and fly ash
159 103.5 - 15.3 -
160 96 - 15 -
161 91.5 - 14 -
162 86.5 - 13 -
163 88.5 - 12.9 -
164 95 - 13.5 -
165 92.5 - 10.3 -
166 111.5 - 10.4 -
167 109.6 - 14.7 -
168 103 - 13.2 -
169 104.5 - 13.2 -
170 Máca et al. (2012) Conventional 132.2 40000 20.8 -
171 110 - 17.6 -
172 141.9 - 22.1 -
173 Mao et al. (2019) Conventional 129.5 - 23.5 -
174 With recycled powder 118 - 22.2 -

175 116 - 22.1 -


176 114 - 21.8 -
177 113 - 20.8 -
178 115 - 21 -
179 105 - 19 -
180 135 - 21 -
181 120 - 19.2 -
182 125 - 22.4 -
183 114 - 22 -

166
Appendix B

184 Mueller et al. With GGBS 135 48300 - -


(2016)
185 Ng et al. (2014) Conventional 150 - 19 -
186 Rong et al. (2010) With ultra-fine slag and ultra-fine fly ash 138 - 17.5 -
187 Ruan et al. (2018) With nano-zirconia 93.7 - 9.5 -
188 Schwartzentruber Conventional 135 - 20.5 -
et al. (2004)
189 Soutsos et al. Conventional 106.6 - 17.1 -
(2005)
190 89.2 - 11.4 -
191 With GGBS, pulverized fuel ash (PFA), and/or ground 117.6 - 10.5 -
glass cullet
192 119.2 - 11.2 -
193 78.2 - 10.8 -
194 89.9 - 13.2 -
195 76.4 - 9.8 -
196 72.6 - 12.2 -
197 96.2 - 11.7 -
198 83.6 - 7.3 -
199 89.8 - 9.2 -
200 Sovják et al. Conventional 132 41000 - 14
(2013)
201 Tafraoui et al. Conventional 98 - 16 -
(2009)
202 120 - 16 -
203 With metakaolin 109 - 17 -
204 119 - 17 -
205 Wu et al. (2016a) With slag and river sand 81.4 33600 - -

206 Wu et al. (2016b) With river sand 90 - 19 -


207 Yu et al. (2014) With natural sand 69.7 - 10.3 -
208 With natural sand and nano-silica 78.0 - 11.9 -
209 84.0 - 13.1 -
210 87.8 - 13.7 -
211 89.2 - 13.8 -
212 88.3 - 13.3 -
213 Yu et al. (2015a) With natural sand and nano-silica 102 - 14.1 -
214 100.5 - 13.6 -
215 110 - 17 -
216 114 - 19 -
217 101 - 14 -
218 105 - 14.7 -
219 110.5 - 17.5 -
220 116 - 20 -
221 100 - 12.5 -
222 99 - 13 -
223 105 - 15.5 -
224 109 - 17.5 -
225 Yu et al. (2015b) With natural sand and nano-silica 100 - 15 -
226 Yunsheng et al. With river sand, ultra-fine slag, ultra-fine fly ash 120 - 15.5 -
(2008)
227 130 - 16.5 -

167
Appendix B

228 140 - 17.1 -


229 Zhu et al. (2016) With natural sand and/or recycled powder from brick and 115 - 15.2 -
cement solids
230 102 - 14.75 -
231 84.5 - 13.4 -
232 78.5 - 11.25 -
233 68 - 9.6 -
234 72 - 12.42 -
235 80 - 13.7 -
236 88 - 14.2 -
237 74 - 13.3 -
238 74 - 12.5 -
239 79 - 11.4 -
240 76 - 11.2 -

168
APPENDIX C
(Miscellaneous)

Table C.1. Constituents of UHPCs and UHPC pastes, presented in Chapter 2, in kg/m3 (based on a target density of 2425 kg/m3)

Mix type Mix ID. Constituents Mix design parameters


MAC VLAC SF S QS W HRWR* DA* b/a w/b
Preliminary UHPC M-SF0WB181BA139 1266 - - - 912 229 16.5 1.4 1.39 0.181
SF0WB181BA139 - 1266 - - 912 229 16.5 1.4 1.39 0.181
SF0WB166BA139 - 1276 - - 919 212 16.6 1.4 1.39 0.166
SF10WB166BA139 - 1149 128 - 919 212 16.6 1.4 1.39 0.166
SF20WB166BA139 - 1021 255 - 919 212 16.6 1.4 1.39 0.166
SF30WB166BA139 - 893 383 - 919 212 16.6 1.4 1.39 0.166
SF20WB166BA125 - 980 245 - 980 203 15.9 1.3 1.25 0.166
SF20WB166BA114† - 942 235 - 1036 195 15.3 1.3 1.14 0.166
SF20WB166BA104 - 906 227 - 1088 188 14.7 1.2 1.04 0.166

ECO-UHPC S0† - 942 235 - 1036 195 15.3 1.3 1.14 0.166
S15 - 800 235 141 1036 195 15.3 1.3 1.14 0.166
S30 - 659 235 283 1036 195 15.3 1.3 1.14 0.166
S45 - 518 235 424 1036 195 15.3 1.3 1.14 0.166
S60 - 377 235 565 1036 195 15.3 1.3 1.14 0.166

MAC or VLAC based P-M-SF0WB181 2029 - - - - 367 26.4 2.2 - 0.181


paste P-SF0WB181 - 2029 - - - 367 26.4 2.2 - 0.181

ECO-UHPC paste P-S0 - 1644 411 - - 341 26.7 2.3 - 0.166


P-S15 - 1397 411 247 - 341 26.7 2.3 - 0.166
P-S60 - 658 411 986 - 341 26.7 2.3 - 0.166
MAC = moderate-C3A cement, VLAC = very-low-C3A cement, SF = silica fume, S = GGBS, QS = quartz sand, W = water, HRWR = high-range water reducer, DA = defoaming
agent, b/a = binder to aggregate ratio, w/b = water to binder ratio
*Solid content

SF20WB166BA114 and S0 represent same mix
Appendix C

Table C.2. Constituents of UHPCs and UHPFRCs, presented in Chapter 3, in kg/m3 (based on
a target density of 2425 kg/m3 for UHPC, 2450 kg/m3 for UHPFRC with 1% steel fibre content,
and 2500 kg/m3 for UHPFRC with 2% steel fibre content)

Mix ID. R-F0 R-F1 R-F2 MK-F0 MK-F2 S-F0 S-F2


Constituents VLAC 921 901 891 921 891 645 624
SF 249 243 241 147 143 249 241
MK - - - 101 98 0 -
S - - - - - 276 267
QP 203 198 196 203 196 203 196
QS 811 793 784 811 784 811 784
W 226 221 218 226 218 226 218
HRWRa 14.7 14.4 14.3 14.7 14.3 14.7 14.3
DAa 1.3 1.3 1.2 1.3 1.2 1.3 1.2
F - 78.4 155 - 155 - 155
Mix design mk/sf 0/100 0/100 0/100 40/60 40/60 0/100 0/100
parameters s/c 0/100 0/100 0/100 0/100 0/100 30/70 30/70
w/b 0.193 0.193 0.193 0.193 0.193 0.193 0.193
VLAC = very-low-C3A cement, SF = silica fume, MK = metakaolin, S = GGBS, QP = quartz powder,
QS = quartz sand, W = water, HRWR = high-range water reducer, DA = defoaming agent, F = fibre,
mk/sf = metakaolin to silica fume ratio, s/c = GGBS to VLAC ratio, w/b = water to binder ratio
a
Solid content

(a) (b)

(c) (d) (e)


Figure C.1. Compression test specimen at 28 days: (a) before failure, and after failure –
(b) S60 (Chapter 2), (c) S-F2 (Chapter 3), (d) F70 (Chapter 4), (e) T60 (Chapter 5)

170
Appendix C

(a) (b)

(c) (d)

Figure C.2. Flexural strength test specimen at 28 days: (a) before failure, (b) typical UHPC specimen at peak load (Chapters 2, 3, and 4), (c) UHPFRC
specimen (S-F2) at around peak load (Chapter 3), (d) UHPFRC specimen (S-F2) after peak load (Chapter 3)

171

You might also like