0% found this document useful (0 votes)
108 views134 pages

Temperature Dependency of The Rheological Properties and Strength of Cemented Paste Backfill That Contains Sodium Silicate

This document summarizes a thesis that studied the effect of time, temperature, binder content, and additives on the rheological properties and mechanical strength of cemented paste backfill (CPB). Experiments were conducted on CPB samples prepared at different temperatures and curing times with various binder blends and sodium silicate dosages. The results showed the rheology and strength of CPB increases with curing time, temperature, and sodium silicate content. The findings contribute to safer and more cost-effective CPB design in mines.
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
Download as pdf or txt
0% found this document useful (0 votes)
108 views134 pages

Temperature Dependency of The Rheological Properties and Strength of Cemented Paste Backfill That Contains Sodium Silicate

This document summarizes a thesis that studied the effect of time, temperature, binder content, and additives on the rheological properties and mechanical strength of cemented paste backfill (CPB). Experiments were conducted on CPB samples prepared at different temperatures and curing times with various binder blends and sodium silicate dosages. The results showed the rheology and strength of CPB increases with curing time, temperature, and sodium silicate content. The findings contribute to safer and more cost-effective CPB design in mines.
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
Download as pdf or txt
Download as pdf or txt
You are on page 1/ 134

Temperature Dependency of the Rheological

Properties and Strength of Cemented Paste Backfill


that Contains Sodium Silicate

Ghada Abdulbaqi Ali

A thesis submitted in partial fulfilment of the


requirements for the degree of
Master of Applied Science
in Civil Engineering

under the supervision of


Dr. Mamadou Fall

Department of Civil Engineering


Faculty of Engineering
University of Ottawa

© Ghada Abdulbaqi Ali, Ottawa, Canada, 2021


ABSTRACT

Over the past decades, cemented paste backfill (CPB) has become a common, environmentally
friendly method of managing mine wastes (such as tailings). This technology allows up to 60% of
the total amount of tailings to be reused and filled in the mine stopes after converting them into
cemented material. Beside reducing the environmental risks associated with the traditional
disposal of these materials, turning them into cemented material and placing them in the
underground mine stopes can also provide secondary support for these stopes in addition to
minimizing the risk of ground subsidence in the mine area. CPB is an engineered mixture of
tailings, water, and hydraulic binder (such as cement, blast furnace slag, and fly ash) that is mixed
in the paste plant and delivered into the mine stopes through a gravity or pumping based
transportation system. During the transportation of CPB through the delivery system pipelines, the
flowability of CPB depends on the rheology of the transported CPB, which is affected by different
factors, such as the transportation time, temperature variation, binder type, and chemical
composition of these mixtures. In addition, the performance of CPB, after placing the CPB mixture
into the mine stopes, is mainly dependent on the role of the hydraulic binder, as it increases the
mechanical strength of the mixture through the process of cement hydration. The mechanical
strength is also influenced by different factors, such as time progress, temperature variation, and
presence of chemical additives. It has previously been found that fresh CPB transported and/or
placed in the mine stopes can be susceptible to temperature variation of different sources, such as
the climatic effects, heat generated from the surrounding rocks, and heat generated during the
process of cement hydration. Unsuitable flowability of CPB through the delivery system might
lead to significant financial losses due to clogging of pipelines with unexpected hardening of CPB
during transportation, which will cause delay in work and possible damages to the pipelines. Also,
failure of CPB structure in the mine stopes due to inappropriate mechanical strength may cause
casualties to the mine workers as well as significant environmental and economic damages. Many
researchers studied the rheological properties and/or strength development of CPB under the
individual effect of any of the aforementioned factors. Additionally, many researchers have
evaluated the coupled effect of some of these factors on the rheology and mechanical strength of
CPB material. Hitherto, there are currently no studies that addressed the combined effect of all
these conditions on the rheological properties and strength development of CPB. At the first stage
of this M.A.Sc. study, a series of experimental tests was conducted on fresh CPB in order to
determine the combined effect of time, temperature, binder content, and chemical additives on the
rheological properties of CPB. These experiments include rheological properties test (yield stress
and viscosity), microstructural analysis (thermal analysis and XRD), chemical analysis (pH and
Zeta potential), and monitoring tests (electrical conductivity), which were conducted on 125 CPB
samples that were mixed and prepared at different temperatures (2oC, 20oC, 35oC) and cured for
different curing time (0 hrs., 0.25 hrs., 1 hr., 2hrs, and 4 hrs.). These samples were prepared with
different blends of hydraulic binders (PCI, PCI/Slag, and PCI/FA) and contained different dosages
of sodium silicate (0%, 0.1%, 0.3%, and 0.5%). The results obtained show that rheology of CPB

ii
increases with the progress of curing time. It also increases with the increase in the initial (mixing
and curing) temperature and content of sodium silicate. It was also found that the partial usage of
slag and FA reduces the rheological properties. However, CPBs containing PCI/FA as binder have
lower rheological properties, and thus better flowability, than those that contain PCI/Slag as
binder. At the second stage of this M.A.Sc. study, in order to understand the combined effect of
time, temperature and sodium silicate content on the strength development of slag-CPB,
unconfined compression (UCS) test, microstructural analysis (thermal analysis and MIP), and
monitoring tests (electrical conductivity, suction, and volumetric water content) were conducted
on 72 CPB samples that were prepared with PCI-Slag as a binder, cured for different times (1 day,
3 days, 7 days, and 28 days) under different curing temperatures of (2oC, 20oC, 35oC), and
contained different dosages of sodium silicate (0%, 0.3% and 0.5%). The results obtained at this
stage showed that the strength development of slag-CPB increases with the progress of curing time
and temperature. It also increases with the increase in the sodium silicate content. Also, the
combined effect of high temperature, high dosage of sodium silicate and longer curing time
showed significant enhancement in the mechanical strength of slag-CPB. The findings of this
M.A.Sc. research will contribute to cost effective, efficient, and safer design of CPB structures in
the mine areas. It will also help in minimizing financial loss associated with unsuitable flowability
of CPB transported in the CPB delivery system besides reducing the risks of human loss, and the
environmental and economic damages associated with the failure of CPB structures.

iii
DEDICATION

To the souls of my father “Abdulbaqi Ali” and my brother “Hussain Ali”.

To my mother “Zahra Ghadeer”

To my husband, Imad, my children, Hazim, Ahmed, and Mayar.

To my treasured family.

Thank you all for your love, support, and patience during my study.

iv
ACKNOWLEDGMENTS

First and foremost, praise and thanks to my God (Alhamdulillah), for providing me with the
opportunity to study, and ability to learn and to fulfil this achievement.

I would like also to express my great thanks and appreciation to my supervisor,


Dr. Mamadou Fall, for his encouragement and guidance during my research study.

I would like to sincerely acknowledge and give special thanks to Dr. Imad Alainachi for
his mentorship, and Dr. Mohammad Al-Umar and Mr. Sada Haruna for their help and support
during my study.

I would like also to acknowledge and express my thanks and appreciation to all of those who
provided friendship, support, assistance and encouragement during my study. They include
Dr. Ahmad Jrade, Mr. Jean-Claude Célestin, Mr. Luc Cloutier, and Mrs. Najlaa Abdul-Hussain.

Last, but not least, I am extremely thankful for my mother “Zahra”, Sisters “Suhair, Suhad,
and Hind” and their families for their endless love and encouragement. I also want to thank my
husband “Imad” and children “Hazim, Ahmed, and Mayar”, this achievement would not have been
possible without your love and support. Finally, I want to pray for the souls of my father
“Abdulbaqi” and my brother “Hussain”, I hope both of you are happy and proud.

v
LIST OF CONTENTS

ABSTRACT ................................................................................................................................................. ii
DEDICATION............................................................................................................................................ iv
ACKNOWLEDGMENTS .......................................................................................................................... v
LIST OF CONTENTS ............................................................................................................................... vi
LIST OF FIGURES ................................................................................................................................... ix
LIST OF TABLES ..................................................................................................................................... xi
LIST OF SYMBOLS AND ABBREVIATIONS .................................................................................... xii
CHAPTER 1 General Introduction ...................................................................................................... 1
1.1 Background and Problem Statement ................................................................................................. 2
1.2 Objectives ......................................................................................................................................... 3
1.3 Research Approaches and Methods .................................................................................................. 4
1.3.1 Research Approaches ................................................................................................................ 4
1.4 Thesis Outline ................................................................................................................................... 6
1.5 References ......................................................................................................................................... 7
CHAPTER 2 Technical Background and Literature Review .......................................................... 13
2.1 Introduction ..................................................................................................................................... 14
2.2 Background of Cemented Paste Backfill (CPB) ............................................................................. 14
2.2.1 Mix design of CPB.................................................................................................................. 15
2.2.2 Geotechnical design of CPB ................................................................................................... 16
2.3 Source of Temperature in Mine Backfill ........................................................................................ 17
2.3.1 Hydration heat ......................................................................................................................... 17
2.3.2 Climatic influence and geographic location ............................................................................ 17
2.3.3 Geological condition and mine stope depth ............................................................................ 18
2.3.4 Rock mass heat........................................................................................................................ 18
2.3.5 Paste-pipe friction ................................................................................................................... 20
2.3.6 Temperature variation produced by the mining operations .................................................... 20
2.4 Background of Hydraulic Binders, Its Hydration and Key Influential Factors .............................. 20
2.4.1 Ordinary Portland Cement (OPC) ........................................................................................... 20
2.4.2 Blast Furnace Slag (Slag) ........................................................................................................ 20
2.4.3 Flay Ash (FA) ......................................................................................................................... 20
2.4.4 Hydration Process ................................................................................................................... 21

vi
2.4.5 Key Factors that Affect Binder Hydration .............................................................................. 22
2.5 Sodium Silicate ............................................................................................................................... 26
2.5.1 Cement-Sodium Silicate Interaction ....................................................................................... 26
2.5.2 Slag-Sodium Silicate Interaction............................................................................................. 26
2.5.3 Fly Ash-Sodium Silicate Interaction ....................................................................................... 27
2.5.4 Effect of sodium silicate on CPB ............................................................................................ 27
2.6 Rheology ......................................................................................................................................... 29
2.6.1 Yield Stress ............................................................................................................................. 29
2.6.2 Viscosity ................................................................................................................................. 32
2.6.3 Rheology of CPB .................................................................................................................... 34
2.7 Compressive Strength ..................................................................................................................... 36
2.7.1 Background ............................................................................................................................. 36
2.7.2 Compressive Strength of CPB ................................................................................................ 37
2.8 Conclusions ..................................................................................................................................... 40
2.9 References ....................................................................................................................................... 41
CHAPTER 3 Paper I: Time- and Temperature-Dependence of Rheological Properties of
Cemented Tailings Backfill with Sodium Silicate .................................................................................. 53
3.1 Abstract ........................................................................................................................................... 54
3.2 Introduction ..................................................................................................................................... 54
3.3 Materials and Experimental Program ............................................................................................. 58
3.3.1 Materials ................................................................................................................................. 58
3.3.2 Sample Preparation ................................................................................................................. 60
3.3.3 Test Methods ........................................................................................................................... 61
3.4 Results and Discussion ................................................................................................................... 63
3.4.1 Effect of temperature on time-dependent changes of rheological properties of CPB ............. 63
3.4.2 Combined effect of type of binder and curing temperature on rheological properties of CPB
68
3.4.3 Combined effect of amount of sodium silicate and curing temperature ................................. 73
3.5 Summary and Conclusion ............................................................................................................... 76
3.6 Data Availability Statement ............................................................................................................ 77
3.7 Acknowledgements ......................................................................................................................... 77
3.8 References ....................................................................................................................................... 78
CHAPTER 4 Paper II: Temperature Dependency of the Strength Development of Slag-
Cemented Paste Backfill that Contains Sodium Silicate ....................................................................... 87

vii
4.1 Abstract ........................................................................................................................................... 88
4.2 Introduction ..................................................................................................................................... 88
4.3 Materials and Equipment Used in the Experiment ......................................................................... 91
4.3.1 Materials ................................................................................................................................. 91
4.3.2 Sample Preparation ................................................................................................................. 93
4.3.3 Test Methods ........................................................................................................................... 94
4.4 Results and Discussion ................................................................................................................... 95
4.4.1 Effect of temperature on the strength development of Slag-CPB that contains Sodium
Silicate. 95
4.4.2 Coupled effect of sodium silicate content and curing temperature on CPB strength. ........... 100
4.5 Summary and Conclusion ............................................................................................................. 104
4.6 Data Availability Statement .......................................................................................................... 105
4.7 Acknowledgements ....................................................................................................................... 105
4.8 References ..................................................................................................................................... 105
CHAPTER 5 Synthesis and Integration of Results ......................................................................... 112
5.1 Introduction ................................................................................................................................... 113
5.2 Effect of Time and Temperature on the Rheological Properties of Cemented Paste Backfill that
Contrains Sodium Silicate ...................................................................................................................... 114
5.3 Effect of Temperature on the Strength Development of Cemented Paste Backfill that Contrains
Sodium Silicate....................................................................................................................................... 115
5.3.1 Mechanical Stability of CPB Structures ............................................................................... 116
5.3.2 Barricade Stability................................................................................................................. 117
5.3.3 Mining Cycles ....................................................................................................................... 118
5.3.4 Binder Consumption ............................................................................................................. 118
5.4 References ..................................................................................................................................... 119
CHAPTER 6 Conclusions and Recommendations .......................................................................... 120
6.1 Conclusions ................................................................................................................................... 121
6.2 Recommendations ......................................................................................................................... 122

viii
LIST OF FIGURES

Figure 1.1 Combined conditions studied in this research ............................................................................. 4


Figure 1.2 Flowchart of research approach ................................................................................................... 5
Figure 1.3 Organization of the thesis ............................................................................................................ 7
Figure 2.1 Components of cemented paste backfill (CPB) ......................................................................... 15
Figure 2.2 Underground delivery systems of CPB ..................................................................................... 16
Figure 2.3 Sources of temperature in mine backfill .................................................................................... 18
Figure 2.4 Variation of hydration heat evolution rate for different binder composite (modified from: Han
et al. 2017). ............................................................................................................................. 19
Figure 2.5 Temperature variation with depth in South African gold mine (modified from: Fall et al. 2014).
................................................................................................................................................ 19
Figure 2.6 Progress of cement hydration with time (modifed from: Bullard et al. 2011). .......................... 22
Figure 2.7 Effect of water-to-cement (w/c) ratio on the rheological properties of self compacted concrete
(modifed from: Ramezani et al. 2016). .................................................................................. 23
Figure 2.8 Effect of water-to-cement (w/c) ratio on the mechanical strength of cemented Paste Backfill. 24
Figure 2.9 Effect of curing temperature on the strength development of cemented Paste Backfill (modified
from: Fall et al. 2010). ............................................................................................................ 25
Figure 2.10 Effect of curing temperature on the rheological properties of cemented Paste Backfill: (a)
Yield stress, and (b) Viscosity (modifed from: Haruna and Fall 2019). ................................ 25
Figure 2.11 Effect of sodium silicate content on the strength development of CPB (modified
from:Kermani et al. 2015). ..................................................................................................... 28
Figure 2.12 Effect of sodium silicate content on the strength development of CPB (modified from: Abdul-
Hussain 2011). ........................................................................................................................ 28
Figure 2.13 Predicted yield stress and flowable zone of CBP of different binder and solid content
(modified from: Belem and Benzaazoua 2008). ..................................................................... 30
Figure 2.14 Schematic view of the Vane shear test method (modified from: Bauer et al. 2007). .............. 31
Figure 2.15 Vane torque spring and electrical transducer (ASTM D4648/D4648M-16 2015). ................. 31
Figure 2.16 Vane shear apparatus used in this study. ................................................................................. 32
Figure 2.17 Vane shear apparatus used in this study. ................................................................................. 33
Figure 2.18 Unconfined Compression Strength (UCS) testing apparatus used in this study. ..................... 37
Figure 2.19 Effect of initial temperature on the strength development of CPB that contains sodium silicate
at early age (modified from: Wang et al. 2016). .................................................................... 39
Figure 3.1 CPB transport system and sources of temperature change. ....................................................... 56
Figure 3.2 Grain size distribution of silica tailings used in this study vs. average grain size distribution of
tailings from nine Canadian mines. ........................................................................................ 58
Figure 3.3 Temperature vs. yield stress of PCI-CPB samples with 0.3% sodium silicate. ........................ 65
Figure 3.4 Temperature vs. viscosity of PCI-CPB samples with 0.3% sodium silicate. ............................ 65
Figure 3.5 DT/DTG diagrams for PCI-CPB samples with 0.3% sodium silicate cured for 4 h vs. different
curing temperatures. ............................................................................................................... 66
Figure 3.6 Zeta potential of PCI-CPB samples with 0.3% of sodium silicate prepared and cured at
different temperatures for 4 h. ................................................................................................ 67

ix
Figure 3.7 Changes in pH of PCI-CPB samples with 0.3% sodium silicate prepared and cured at different
temperatures for 4 h. ............................................................................................................... 67
Figure 3.8 Effect of temperature and binder type on yield stress of CPB samples with 0.3% sodium
silicate cured for 2 hrs. ........................................................................................................... 69
Figure 3.9 Effect of temperature and binder type on viscosity of CPB samples with 0.3% sodium silicate
cured for 2 hrs......................................................................................................................... 69
Figure 3.10 DT/DTG diagrams for CPB samples with 0.3% of sodium silicate and cured for 2 h with
different binder type at 35oC. ................................................................................................. 71
Figure 3.11 XRD intensities for CPB samples with 0.3% of sodium silicate cured for 2 h with different
binder type at 35oC. ................................................................................................................ 72
Figure 3.12 Development of EC in CPB samples with 0.3% of sodium silicate and different binder type at
35oC. ....................................................................................................................................... 73
Figure 3.13 Effect of temperature and sodium silicate content on yield stress of PCI-CPB cured for 2 h. 74
Figure 3.14 Effect of temperature and sodium silicate content on viscosity of PCI-CPB cured for 2 h. ... 75
Figure 3.15 DT/DTG of PCI-CPB-SS with different sodium silicate contents prepared and cured for 4 h at
35oC. ....................................................................................................................................... 75
Figure 3.16 Time-dependent changes of electric conductivity of PCI-CPB-SS with different sodium
silicate contents cured at 35ºC. ............................................................................................... 76
Figure 4.1 Sources of temperature change in CPB placed in mine stope. .................................................. 90
Figure 4.2 Grain size distribution of silica tailings used in this study and the average grain size
distribution of tailings from nine Canadian mines. ................................................................ 92
Figure 4.3 Effect of temperature on the strength development of Slag-CPB that contain 0.3% sodium
silicate..................................................................................................................................... 96
Figure 4.4 DT/DTG diagrams for 7 days Slag-CPB specimens that contain 0.3 % sodium silicate, cured at
various temperatures (20oC and 35oC). .................................................................................. 97
Figure 4.5 Differential pore size distribution curves of 7 days CPB specimens that contain 0.3% sodium
silicate, cured at various temperatures (20oC and 35oC). ....................................................... 98
Figure 4.6 EC of the 28 days Slag-CPB samples that contain 0.3% sodium silicate, cured at different
temperatures of 20oC and 35oC. ............................................................................................. 99
Figure 4.7 Coupled development of VWC and suction (negative PWP) of the 28 days Slag-CPB samples
that contain 0.3% sodium silicate, cured at different temperatures of 20oC and 35oC. .......... 99
Figure 4.8 Effect of temperature and sodium silicate content on the strength development of Slag-CPB
that contain sodium silicate and cured at: (a) 35oC; (b) 2oC................................................. 101
Figure 4.9 Effect of sodium silicate content on TG/DTG diagrams for 7 days Slag-CPB specimens that
cured at 35oC. ....................................................................................................................... 102
Figure 4.10 Effect of sodium silicate content on the differential pore size distribution for 7 days Slag-CPB
specimens that cured at 35oC. ............................................................................................... 103
Figure 4.11 Effect of sodium silicate content on the EC of 7 days Slag-CPB specimens that cured at 35oC.
.............................................................................................................................................. 104
Figure 4.12 Effect of sodium silicate content on the VWC and suction of 7 days Slag-CPB specimens that
cured at 35oC. ....................................................................................................................... 104
Figure 5.1 Critical UCS of CPB with respect to the width variation of a 60 m height stope. .................. 117

x
LIST OF TABLES

Table 3.1. Primary physical properties of tailings used in this study......................................................... 58


Table 3.2. Main chemical elements of tailings used in this study............................................................... 59
Table 3.3. Primary physical and chemical properties of binders used in this study ................................... 59
Table 3.4. Properties of sodium silicate (National Silicate Ltd.) ................................................................ 60
Table 3.5. Summary of CPB mix proportion and stages of experimental program .................................... 61
Table 4.1. Primary physical properties of tailings used in this study......................................................... 92
Table 4.2. Main chemical elements of tailings used in this study............................................................... 92
Table 4.3. Primary physical and chemical properties of binders used in this study ................................... 93
Table 4.4. Properties of the sodium silicate used in this study (national Silicate Ltd.). ............................. 93
Table 4.5. Summary of CPB mix proportion and stages of experimental program .................................... 94
Table 5.1 Summary of effective factors that were investigated in this thesis ........................................... 113
Table 5.2 Summary of experimental tests conducted in this thesis .......................................................... 113

xi
LIST OF SYMBOLS AND ABBREVIATIONS

9MT : Nine Mine Tailings


AMD : Acid Mine Drainage
CPB : Cemented Paste Backfill
DTG : Differential Thermogravimetry
EC : Electrical Conductivity
FA : Fly Ash
hrs : Hours
min : Minutes
MIP : Mercury Intrusion Porosimetry
OPC : Ordinary Portland Cement
PC : Portland Cement
PCI : Portland Cement – type I
sec : Seconds
Slag : Blast Furnace Slag
SS : Sodium Silicate
ST : Silica Tailings
Temp. : Temperature
TG : Thermal Gravimetry
UCS : Unconfined Compression Strength
UCSc : Critical Unconfined Compression Strength
VWC : Volumetric Water Content
w/c : Water-to-Cement Ratio
wt% : percentage of the total dry weight
XRD : X-ray Diffraction
𝜃 : Diffraction Angle

xii
CHAPTER 1
General Introduction

1
1.1 Background and Problem Statement

Mining operations were established in Canada in 1577. Since then, mining industry has become a
lead in the economic growth in Canada. For instance, the total Canada-wide mineral production in
2019 was more than 48 billion CAD (Chang 2016, Statista 2020). Moreover, the Canadian mining
industry offered around 626,000 jobs in 2018, with an estimation of around 80,000 additional jobs
to be offered over the next decade, i.e. up to 2030 (Marshall 2020).

Ore extraction during mining operations create large underground openings (stopes) that
might cause several geotechnical engineering issues endangering the safety of the mine workers
and the nearby public congregation. These issues include ground subsidence and the instability of
these openings (Alainachi and Fall 2019). In addition, mining operations annually generate large
amount of mining wastes. The worldwide generation of mine solid waste (e.g. tailings) was
estimated to be over 100 billion tons per year (Tayebi-Khorami et al. 2019). Poor disposal of these
tailings may harm the environment, as they can generate acid mine drainage (AMD). The AMD
has always been a main problem associated with mining activities, as it causes contamination of
underground and surface water bodies if deported from mine disposal site (Heikkinen et al. 2009,
Abdul-Hussain 2011).

An environmentally friendly underground disposal method of mine waste has been


extensively used around the world in order to lessen the engineering and environmental problems
associated with mining operations. This technology is called the cemented paste backfill (CPB),
which is summarized by the reuse of large quantities of the tailings and pumping them back to the
mine stope after mixing them with water and binder material (Belem and Benzaazoua 2004, Bull
2019). CPB has been found to be beneficial in increasing the stability of the underground mine in
long term (Alainachi and Fall 2019, Ercikdi et al. 2017)

CPB is often prepared and mixed in the surface paste plants, and hydraulically pumped
and/or transported by gravity to the underground mine stopes. However, if the transported CPB
material does not have suitable flowability, it might cause significant financial losses and
production interruption due to the possible flow delay and possible clogging-induced pipelines’
damages (Li and Fall 2016, Haruna and Fall 2020). Also, the mechanical strength performance of
CPB is significantly affected when exposed to variation of static loading, dynamic loading,
temperature variation, and potential chemical interaction of the chemical elements within the CPB
components. Inadequate mechanical performance of CPB causes failure of CPB structure and thus
endangers the safety of the mine workers and causes negative environmental impacts and
economic damages (Fall and Pokharel 2010, Cao et al. 2018, Xue et al. 2018, Alainachi 2020)

Considerable amount of research, (Simon and Grabinsky 2013, Cui and Fall 2016,
Haiqiang et al. 2016, Jiang and Fall 2017, Haruna and Fall 2019), has been undertaken to evaluate
the flowability and/or the rheological properties of CPB. Other studies (e.g. Kesimal et al. 2005,
Nasir and Fall 2010, Li and Aubertin 2012, Ghirian and Fall 2013, Ghirian and Fall 2014, Cui and
Fall 2016, Ghirian and Fall 2016, Cihangir et al. 2018, Fang and Fall 2019, Haruna and Fall 2020,
Xu et al. 2020) have studied the mechanical behavior or stability of CPB exposed to different

2
conditions, such as temperature variation, curing time, different CPB mix design, and availability
of different chemical components within the CPB. Yet, there is paucity of studies that addressed
the combined effect of these conditions on the flowability and/or the mechanical strength of CPB
at early age. Thus, the need of more studies to understand the combined effect of these conditions
on the flowability and strength of early aged CPB has become critically important.

The engineering properties (such as the mechanical, thermal, and hydraulic behavior etc.)
of cementitious material, including CPB, was found to be affected by the progress of cement
hydration (i.e. progress of curing time) (Bullard et al. 2011, Scrivener et al. 2015, Aldhafeeri and
Fall 2016) and the effect of partial usage of different hydraulic binders (such as slag and flay ash)
(Maltais and Marchand 1997, Elakneswaran et al. 2009, Kwan and Chen 2013, Kondraivendhan
and Bhattacharjee 2015, Le et al. 2019). Moreover, it was previously observed that the temperature
of fresh CPB placed in mine stopes is affected by several heat sources, which might affect the
rheological, mechanical, thermal, and hydraulic properties of CPB (Maltais and Marchand 1997,
Fall et al. 2010, Nasir and Fall 2010, Aldhafeeri et al. 2016, Cui and Fall 2016, Wang et al. 2016,
Fang and Fall 2018, Wang et al. 2018, Haruna and Fall 2019, Liu et al. 2019, Cheng et al. 2020,
Haruna and Fall 2020). On the other hand, previous studies have revealed that adding some
chemical additives (such as sodium silicate) to the CPB may affect its flow behavior and
mechanical strength. It is expected that the flowability of cementitious material drops when it
contains sodium silicate, while the mechanical behavior is expected to enhance (Brough and
Atkinson 2002, Živica 2007, Abdul-Hussain and Fall 2012, Kermani et al. 2015, Wang et al. 2016,
Veenstra et al. 2017, Cihangir et al. 2018, Jiang et al. 2019).

Accordingly, the problem statement of this research, as shown in Figure 1.1, can be
summarized as the need of comprehensive study of the rheological properties and mechanical
strength of CPB at the early ages by assessing the combined effect of time, temperature, binder
type, and chemical additives on these properties.

1.2 Objectives

The primary goal of this M.A.Sc. study is to investigate the temperature dependency of the
rheological properties and strength development of CPB under the combined effect of time
progress, binder type, and the content of sodium silicate as a chemical additive. Silica tailings are
used to determine how the initial (mixing and curing) temperature affects the rheological properties
and the mechanical strength of samples prepared with different types of hydraulic binders and
different dosages of sodium silicate. The experiment program of this research includes a
combination of different tests that are conducted at different temperature conditions. These tests
include yield stress and viscosity tests, unconfined compression tests (UCS), microstructure
analysis, chemical tests, and monitoring tests. These experiments aim to investigate:

1. The influence of initial (mixing and curing) temperature and the progress of cement
hydration process (progress of curing time) on the rheological properties of CPB that
contains sodium silicate.

3
Figure 1.1 Combined conditions studied in this research

2. Whether the type of hydraulic binders of the CPB mixture changes the rheological
properties of CPB that contains sodium silicate.

3. The effect of different dosages of sodium silicate on the rheological properties of CPB
samples.

4. The influence of curing temperature on the mechanical strength of slag-CPB material that
contains sodium silicate.

5. The change in mechanical strength of slag-CPB material with the variation of sodium
silicate content.

1.3 Research Approaches and Methods

1.3.1 Research Approaches

The research approaches adopted in this study have been illustrated in the schematic flowchart in
Figure 1.2. To better understand the temperature and time effect on the rheological properties and
the development of the mechanical strength of CPB that is made with different binders and
contains different dosages of sodium silicate, this study has been conducted in three main phases:

4
Figure 1.2 Flowchart of research approach

Phase 1 - A comprehensive literature review of (i) the material to be tested (CBP), (ii) different
hydraulic binder types, (iii) sodium silicate as chemical additive, (iv) the rheology of CPB, and (v)
the compressive strength (UCS).

Phase 2 - Conducting the experimental program that will be divided into two sub-phases to better
understand each of the studied properties. These sub-phases are:

1. Conducting rheological (yield stress and viscosity) tests, microstructure (Thermal and
XRD) analysis, chemical (pH and zeta potential) tests, and monitoring (electrical
conductivity) test on CPB samples that consist of different binders (Portland cement, Slag,
and Fly Ash) blends, contain different amount of sodium silicate (0, 0.1%, 0.3%, and
0.5%), and were prepared at different temperatures (2oC, 20oC, and 35oC) and cured to
different maturity ages (curing time) of 0, 0.25, 1, 2, and 4 hours.

2. Conducting unconfined compressive strength (UCS) test, microstructure (Thermal and


MIP) analysis, and monitoring (suction, volumetric water content and electrical
conductivity) tests on CPB samples that consist of 50% of slag and 50% of Portland cement

5
as a binder, contain different amount of sodium silicate (0, 0.3%, and 0.5%), and cured
under different temperatures (2oC, 20oC, and 35oC) to different maturity ages (curing time)
of 1, 3, 7, and 28 days.

Phase 3 - Analyzing the experimental results and assessing the rheological properties and strength
development of CPB with respect to the tested conditions.

1.4 Thesis Outline

This thesis is organized in the form of technical papers and contains six chapters (Figure 1.3):

Chapter 1: includes a general introduction to the study, the problem statements, objectives, and
research approaches and methods adopted in this study.

Chapter 2: reviews technical background and literature review on Cemented Paste Backfill (CPB),
hydraulic binders, sodium silicate, rheology, and compressive strength. This information is
necessary to facilitate the understanding of the technical issues addressed in this thesis. Moreover,
previous studies of rheological properties and strength development of cemented paste backfill
have been reviewed and discussed.

Chapter 3: discusses the combined effect of time and temperature on the rheological properties of
cemented tailings backfill that contains sodium silicate (Technical Paper I).

Chapter 4: describes the temperature effect on the strength development of slag-cemented paste
backfill that contains sodium silicate (Technical Paper II).

Chapter 5: syntheses the overall thesis results.

Chapter 6: presents the conclusions and recommendations of this thesis.

It should be emphasized that since a paper-based thesis format has been adopted, some of the
contents in the papers may be repeated as each paper is independently written and crafted
according to the manuscript instructions for the specified publication.

6
Chapter 1
• General Introduction
Chapter 2
• Techniccal Background and Literature Review
Chapter 3
• Paper 1
• Time- and temperature-dependence of rheological properties of
cemented tailings backfill with sodium silicate
Chapter 4
• Paper 2
• Temperature Dependency of the Strength Development of Slag-
Cemented Paste Backfill that Contain Sodium Silicate
Chapter 5
• Synthesis of Results and Discussion
Chapter 6
• Conclusions and Recommendations
Figure 1.3 Organization of the thesis

1.5 References

Abdul-Hussain, N. (2011) Experimental Study on the Engineering Properties of Gelfil, Masters


thesis, University of Ottawa.

Abdul-Hussain, N. and Fall, M. (2012) 'Thermo-hydro-mechanical behaviour of sodium silicate-


cemented paste tailings in column experiments', Tunnelling and Underground Space
Technology, 29, 85-93.

Alainachi, I. and Fall, M. (2019) 'Liquefaction potential of cementing tailing backfill: shaking table
test results', in Proceedings of 16th Panamerican conference on soil mechanics and
geotechnical engineering, Cancun, Mexico, IOS Press, 1949-1956.

Alainachi, I. H. (2020) Shaking Table Testing of Cyclic Behaviour of Fine-Grained Soils


Undergoing Cementation: Cemented Paste Backfill, Doctoral Dissertation, University of
Ottawa.

7
Aldhafeeri, Z. and Fall, M. (2016) 'Time and damage induced changes in the chemical reactivity
of cemented paste backfill', Journal of Environmental Chemical Engineering, 4(4), 4038-
4049.

Aldhafeeri, Z., Fall, M., Pokharel, M. and Pouramini, Z. (2016) 'Temperature dependence of the
reactivity of cemented paste backfill', Applied Geochemistry, 72, 10-19.

Belem, T. and Benzaazoua, M. (2004) An overview on the use of paste backfill technology as a
ground support method in cut and fill mines, translated by Perth, Australia: CRC Press,
637–650.

Brough, A. R. and Atkinson, A. (2002) 'Sodium silicate-based, alkali-activated slag mortars: Part
I. Strength, hydration and microstructure', Cement and Concrete Research, 32(6)(6), 865-
879.

Bull, A. (2019) Temperature Dependence of the Leachability of Cemented Paste Backfill, M.A.Sc.
Thesis: University of Ottawa.

Bullard, J. W., Jennings, H. M., Livingston, R. A., Nonat, A., Scherer, G. W., Schweitzer, J. S.,
Scrivener, K. L. and Thomas, J. J. (2011) 'Mechanisms of cement hydration', Cement and
Concrete Research, 41(12), 1208-1223.

Cao, S., Yilmaz, E. and Song, W. (2018) 'Dynamic response of cement-tailings matrix composites
under SHPB compression load', Construction and Building Materials, 186, 892-903.

Chang, S. (2016) Strength and Deformation Behaviour of Cemented Paste Backfill in Sub-zero
Environment, M.A.Sc. thesis: University of Ottawa.

Cheng, H., Wu, S., Li, H. and Zhang, X. (2020) 'Influence of time and temperature on rheology
and flow performance of cemented paste backfill', Construction and Building Materials,
231.

Cihangir, F., Ercikdi, B., Kesimal, A., Ocak, S. and Akyol, Y. (2018) 'Effect of sodium-silicate
activated slag at different silicate modulus on the strength and microstructural properties
of full and coarse sulphidic tailings paste backfill', Construction and Building Materials,
185, 555-566.

8
Cui, L. and Fall, M. (2016) 'Mechanical and thermal properties of cemented tailings materials at
early ages: Influence of initial temperature, curing stress and drainage conditions',
Construction and Building Materials, 125, 553-563.

Elakneswaran, Y., Nawa, T. and Kurumisawa, K. (2009) 'Zeta potential study of paste blends with
slag', Cement and Concrete Composites, 31(1), 72-76.

Ercikdi, B., Cihangir, F., Kesimal, A. and Deveci, H. (2017) 'Practical Importance of Tailings for
Cemented Paste Backfill' in Yilmaz, E. and Fall, M., eds., Paste Tailings Management,
Cham: Springer International Publishing, 7-32.

Fall, M., Célestin, J. C., Pokharel, M. and Touré, M. (2010) 'A contribution to understanding the
effects of curing temperature on the mechanical properties of mine cemented tailings
backfill', Engineering Geology, 114(3-4), 397-413.

Fall, M. and Pokharel, M. (2010) 'Coupled effects of sulphate and temperature on the strength
development of cemented tailings backfills: Portland cement-paste backfill', Cement and
Concrete Composites, 32(10), 819-828.

Fang, K. and Fall, M. (2018) 'Effects of curing temperature on shear behaviour of cemented paste
backfill-rock interface', International Journal of Rock Mechanics and Mining Sciences,
112, 184-192.

Fang, K. and Fall, M. (2019) 'Shear Behavior of the Interface Between Rock and Cemented
Backfill: Effect of Curing Stress, Drainage Condition and Backfilling Rate', Rock
Mechanics and Rock Engineering, 53(1), 325-336.

Ghirian, A. and Fall, M. (2013) 'Coupled thermo-hydro-mechanical–chemical behaviour of


cemented paste backfill in column experiments. Part I: Physical, hydraulic and thermal
processes and characteristics', Engineering Geology, 164, 195-207.

Ghirian, A. and Fall, M. (2014) 'Coupled thermo-hydro-mechanical–chemical behaviour of


cemented paste backfill in column experiments', Engineering Geology, 170, 11-23.

Ghirian, A. and Fall, M. (2016) 'Influence of Curing Stress on Long-term Hydro-Mechanical


Response of Cemented Paste Backfill', in Tailings and Mine Waste, Keystone, Colorado,
USA,

9
Haiqiang, J., Fall, M. and Cui, L. (2016) 'Yield stress of cemented paste backfill in sub-zero
environments: Experimental results', Minerals Engineering, 92, 141-150.

Haruna, S. and Fall, M. (2019) 'Time- and temperature-dependent rheological properties of


cemented paste backfill that contains superplasticizer', Powder Technology.

Haruna, S. and Fall, M. (2020) 'Strength Development of Cemented Tailings Materials Containing
Polycarboxylate ether-based Superplasticizer: Experimental Results on the Effect of Time
and Temperature', Canadian Journal of Civil Engineering.

Heikkinen, P., Räisänen, M. and Johnson, R. (2009) 'Geochemical characterisation of seepage and
drainage water quality from two sulphide mine tailings impoundments: acid mine drainage
versus neutral mine drainage', Mine Water and the Environment, 28(1), 30-49.

Jiang, H. and Fall, M. (2017) 'Yield stress and strength of saline cemented tailings in sub-zero
environments: Portland cement paste backfill', International Journal of Mineral
Processing, 160, 68-75.

Jiang, H., Qi, Z., Yilmaz, E., Han, J., Qiu, J. and Dong, C. (2019) 'Effectiveness of alkali-activated
slag as alternative binder on workability and early age compressive strength of cemented
paste backfills', Construction and Building Materials, 218, 689-700.

Kermani, M., Hassani, F. P., Aflaki, E., Benzaazoua, M. and Nokken, M. (2015) 'Evaluation of
the effect of sodium silicate addition to mine backfill, Gelfill − Part 1', Journal of Rock
Mechanics and Geotechnical Engineering, 7(3), 266-272.

Kesimal, A., Yilmaz, E., Ercikdi, B., Alp, I. and Deveci, H. (2005) 'Effect of properties of tailings
and binder on the short-and long-term strength and stability of cemented paste backfill',
Materials Letters, 59(28), 3703-3709.

Kondraivendhan, B. and Bhattacharjee, B. (2015) 'Flow behavior and strength for fly ash blended
cement paste and mortar', International Journal of Sustainable Built Environment, 4(2),
270-277.

Kwan, A. K. H. and Chen, J. J. (2013) 'Adding fly ash microsphere to improve packing density,
flowability and strength of cement paste', Powder Technology, 234, 19-25.

10
Le, T. H. M., Park, D.-W., Park, J.-Y. and Phan, T. M. (2019) 'Evaluation of the Effect of Fly Ash
and Slag on the Properties of Cement Asphalt Mortar', Advances in Materials Science and
Engineering, 2019, 1-10.

Li, L. and Aubertin, M. (2012) 'A modified solution to assess the required strength of exposed
backfill in mine stopes', Canadian Geotechnical Journal, 49(8), 994-1002.

Li, W. and Fall, M. (2016) 'Sulphate effect on the early age strength and self-desiccation of
cemented paste backfill', Construction and Building Materials, 106, 296-304.

Liu, Y., Li, H. and Wu, H. (2019) 'Experimental Study on Mechanical Properties of Cemented
Paste Backfill under Temperature-Chemical Coupling Conditions', Advances in Materials
Science and Engineering, 2019, 1-10.

Maltais, Y. and Marchand, J. (1997) 'Influence of curing temperature on cement hydration and
mechanical strength development of fly ash mortars', Cement and Concrete Research,
27(7)(7), 1009-1020.

Marshall, B. (2020) The State of Canada’s Mining Industry: Facts and Figures 2019, The Mining
Association of Canada.

Nasir, O. and Fall, M. (2010) 'Coupling binder hydration, temperature and compressive strength
development of underground cemented paste backfill at early ages', Tunnelling and
Underground Space Technology, 25(1), 9-20.

Scrivener, K. L., Juilland, P. and Monteiro, P. J. M. (2015) 'Advances in understanding hydration


of Portland cement', Cement and Concrete Research, 78, 38-56.

Simon, D. and Grabinsky, M. (2013) 'Apparent yield stress measurement in cemented paste
backfill', International Journal of Mining, Reclamation and Environment, 27(4), 231-256.

Statista (2020) Canada's mining industry.

Tayebi-Khorami, M., Edraki, M., Corder, G. and Golev, A. (2019) 'Re-Thinking Mining Waste
through an Integrative Approach Led by Circular Economy Aspirations', Minerals, 9(5).

11
Veenstra, R., Slattery, T., Chauvier, C. and van der Merwe, S. (2017) Evaluation of Mount Heaney
leads slag to produce a cemented backfill, translated by University of Science and
Technology Beijing, 294-301.

Wang, Y., Fall, M. and Wu, A. (2016) 'Initial temperature-dependence of strength development
and self-desiccation in cemented paste backfill that contains sodium silicate', Cement and
Concrete Composites, 67, 101-110.

Wang, Y., Wu, A., Ruan, Z., Wang, H., Wang, Y. and Jin, F. (2018) 'Temperature Effects on
Rheological Properties of Fresh Thickened Copper Tailings that Contain Cement', Journal
of Chemistry, 2018, 1-8.

Xu, X., Fall, M., Alainachi, I. and Fang, K. (2020) 'Characterisation of fibre-reinforced
backfill/rock interface through direct shear tests', Geotechnical Research, 7(1)(1), 1-15.

Xue, G., Yilmaz, E., Song, W. and Cao, S. (2018) 'Compressive Strength Characteristics of
Cemented Tailings Backfill with Alkali-Activated Slag', Applied Sciences, 8(9).

Živica, V. (2007) 'Effects of type and dosage of alkaline activator and temperature on the properties
of alkali-activated slag mixtures', Construction and Building Materials, 21(7), 1463-1469.

12
CHAPTER 2
Technical Background and Literature Review

13
2.1 Introduction
In order to better evaluate the effect of temperature on the rheological properties and strength
development of cemented paste backfill (CPB), this chapter provides a review of the fundamental
technical background information on CPB. Section 2.2 provides background information on CPB
technology, including the mix design and preparation, delivery methods, backfilling strategy, and
the design of CPB from geotechnical point of view as well as the source of temperature in backfill
operations. In section 2.3, the current knowledge on different hydraulic binders (Portland Cement,
Slag, and Fly Ash) and some of the factors that affect the hydration process of these binders have
been deliberated. After that, the background of sodium silicate and its effect on cementitious
material have been elucidated in section 2.4. Afterwards, the background information to
understand the theory of rheology, the equipment and testing methods for rheology, and a review
of the previous studies on the rheology of CPB have been described in section 2.5. Furthermore,
background of compressive strength and a review of the previous studies that used this technique
on CPB have been described in sections 2.6.

2.2 Background of Cemented Paste Backfill (CPB)


Underground mining operations produce a huge quantity of mine wastes. Beside not having any
benefit to the industry, mishandling of these wastes makes them highly dangerous to the
environment as they may generate several hazards, such as acid mine drainage (AMD). Also,
mining activities generate large underground voids that may aggravate the risks of ground
subsidence (Chang 2016, Cui 2017). Although there are several waste management techniques that
are often used, but each type of these “traditional” methods is associated with its own
environmental, geotechnical, and/or economic challenges. For instance, surface disposal, such as
stacking tailings in a series of dams, is one of the common tailing disposal methods. However,
beside the fact that this method requires a large (over ground) area, these dams might fail resulting
in multiple environmental and economic damages, as the tailings might rapidly drift for several
kilometers (Chang 2016, Bull 2019). Therefore, in order to minimize these challenges, a need for
new solution has been felt. One of these solutions is the cemented paste backfill (CPB), which is
a modern technology of mine waste management that was developed in the early 1980s in
Germany. This technology has been extensively used in several underground mines around the
world as it allows big quantities of tailings to be turned into cemented backfill and returned to the
mine voids. This method is environmentally friendly as it reduces the risk of generating acid mine
drainage (AMD) and/or other toxic materials. Furthermore, from the geotechnical point of view,
the use of CPB prevents the collapse and/or roof fall of the mine stope by providing ground support
to the walls and pillars. Economically, on the other hand, CPB can be produced and delivered to
the mine stope in a short period, which reduces the backfilling process (mining cycle) in a matter
of days and consequently increases the mine productivity (Belem and Benzaazoua 2004, Abdul-
Hussain 2011, Cui 2017, Aldhafeeri 2018, Alainachi and Fall 2019a).

14
2.2.1 Mix design of CPB
The identical CPB mixture consists of mine tailings with a solid percentage of 70% - 85%, fresh
or mine processed water with a water-to-cement ratio (w/c) that can range between 5.5 and 9.6,
and around 2% - 9% (by total weight of solid) of hydraulic binder (Figure 2.1). The content of
these ingredients is often selected based on the primary and tertiary requirements of the mine stope
as well as the tailings fineness and density which affect CPB transportability (slump) (Alakangas
et al. 2013, Yilmaz et al. 2013, Aldhafeeri and Fall 2016, Haiqiang et al. 2016, Koohestani et al.
2018). Ordinary Portland Cement (OPC) is perceived to be the most hydraulic binder used in CPB
mixture. However, a partial replacement for the OPC with pozzolanic materials, such as blast
furnace slag (Slag) and/or fly ash (FA), has become a common practice in CPB technology in order
to reduce the cost due to cement consumption without affecting the desired strength of CPB (Tariq and
Yanful 2013, Cui 2017).

Figure 2.1 Components of cemented paste backfill (CPB)


CPB mixture is often prepared in paste backfill plant, which is usually located at the surface
of the mine. Produced CPB is then transported through underground distribution system. Based on
the amount of energy required to deliver it, CPB is often delivered into the mine stope by pumping,
gravity and/or combination of both (Figure 2.2) (Belem and Benzaazoua 2004, Cui 2017).
Therefore, the flowability of fresh CPB is a key concern in CPB transportation to the mine stope.

15
If transported material did not have suitable flowability, pipelines’ clogging might occur, which
will cause flow delays, temporary disruption of the progress of CPB production, and/or possible
replacement of clogged pipelines, resulting in significant financial losses (Wu et al. 2013, Li and
Fall 2016, Haruna and Fall 2019). To ensure better flowability of transported fill in the pipelines,
it is imperative to study the rheological properties of these fills, such as yield stress and viscosity.
To maintain appropriate flowability, the pumping pressure of fresh CPB must be larger than the
yield stress of the produced paste. In order to optimize the delivery system, appropriate balance
between pumping velocity, the pipe diameters, and relative density of the fill has to be taken into
consideration. It is also worth noting that the rheological properties of CPB, and consequently its
flowability through the pipelines, are significantly affected by several factors, such as temperature,
binder type, the presence of chemical additives, and the elapsed time after mix preparation (Wu et
al. 2013, Haiqiang et al. 2016, Cui 2017, Wang et al. 2018, Haruna and Fall 2019, Alainachi 2020).
Further details on the rheological properties of CPB will be discussed in section 2.5.

Figure 2.2 Underground delivery systems of CPB

2.2.2 Geotechnical design of CPB


While designing a CPB structure, it is vitally important to take into consideration the economic
efficiency of the CPB structure as well as the stability of the structure, and consequently the safety

16
of the workers in the mine workplace and the surrounding area. Economically, it is important to
shorten the completion time of the backfill process in order to reduce the mining cycle time that
will consequently boost the mine production. From the safety standpoint, the CPB structures
should be designed to have a sufficient strength in order to provide a safe and stable platform for
mine workers besides preventing caving of the mine stope and/or subsidence of the ground of the
mine and the neighboring area. Accordingly, the strength and stability of the CPB structures play
the major role in achieving the best CPB performance with respect to workers’ safety and mine
productivity, and thus in designing the CPB structures (Belem and Benzaazoua 2004, Chang 2016,
Lu 2017). Strength performance of CPB structures is also found to be significantly affected by the
temperature, binder type, the post-mix and preparation time, and the presence of chemical
additives (Fall et al. 2010, Nasir and Fall 2010, Ghirian and Fall 2013, Chang 2016, Wang et al.
2016, Fang and Fall 2018). Further details on the strength of CPB will be discussed in section 2.6.

2.3 Source of Temperature in Mine Backfill


During its delivery and/or when it is placed in mine stopes, several sources might affect the
temperature of CPB in the field (Figure 2.3). These sources might be internal, such as the heat
generated during hydration of the hydraulic binders, or external, such as the climatic influence
based on the geographic location, temperature variation with stope depth, heat from the interaction
between CPB and the rock mass, and heat from paste-pipe friction as along with the temperature
variation produced by the mining operations (Fall et al. 2007, Fall et al. 2010, Wu et al. 2013,
Aldhafeeri et al. 2016, Wang et al. 2016, Alainachi and Fall 2019b, Bull 2019).

2.3.1 Hydration heat


There is a common consensus that exothermic reaction of the components of cement or any other
hydraulic binder with water generates a significant amount of heat. This generated heat is
associated with the formation of cement hydration products that are responsible for the hardening
of cementitious materials, such as CPB (Fall et al. 2010). Due to the large size of CPB structures,
the generated heat does not easily dissipate through these structures. So, there will be a noticeable
increase in temperature of the CPB placed in the mine stope through its curing process (Bull 2019).
However, the rate of heat evolution or the amount of released heat, in other words, varies with the
progress of time and/or the type and content of the hydraulic binder used in the mix (Figure 2.4)
(Schindler and Folliard 2003, Bullard et al. 2011, Han et al. 2017).

2.3.2 Climatic influence and geographic location


The temperature of CPB transported through the pipeline and/or placed in the mine stope is
significantly affected by the climate fluctuations (e.g. winter/summer) and the geographic location
of the mine itself, especially in the extremely hot regions and/or permafrost regions (Wang et al.
2016, Aldhafeeri 2018, Alainachi and Fall 2019b).

17
Figure 2.3 Sources of temperature in mine backfill

2.3.3 Geological condition and mine stope depth


Simultaneously with the high demand on the extraction of more ores, there is a gradual reduction
of ore available at shallow depths in many underground mines around the world. Hence,
underground mine operations are being extensively undertaken at greater depths (Johnson 2015).
Generally, the depth increase is naturally associated with increase in heat flux, which is attributed
to the geothermal gradient (Wu et al. 2013). As shown in Figure 2.5, in deep mines (more than
2500 m depth), the temperature is expected to exceed 45oC, while it might be less than 20oC at
shallow depths (0 m – 500 m) (Fall et al. 2014).

2.3.4 Rock mass heat


The surrounding rock mass is another major heat source for CPB in any deep level mining
operation. The thermal interaction between the CPB transported/placed within the mined-out stope
and the exposed rock mass will obviously change the temperature of the CPB (Wu et al. 2013).

18
12
Hydration Heat Evolution Rate (J/g.h)
Cement
10 Cement-Slag
Cement-FA
8

0
0 10 20 30 40 50 60 70 80
Hydration Time (h)

Figure 2.4 Variation of hydration heat evolution rate for different binder composite (modified from: Han
et al. 2017).

80

70

60

50
Temperature (°C)

40

30

20

10

0
0 1,000 2,000 3,000 4,000 5,000
Depth (m)

Figure 2.5 Temperature variation with depth in South African gold mine (modified from: Fall et
al. 2014).

19
2.3.5 Paste-pipe friction
Regardless of the adopted delivery and distribution system of the CPB, there is always friction
between the transported CPB and the inner sidewalls of the pipelines during the transportation of
the fill from the production plant and the stope. This friction often elevates the temperature of the
fresh transported CPB (Wu et al. 2013).

2.3.6 Temperature variation produced by the mining operations


Mining operations may use some technologies that can contribute to temperature variation of CPB
within the mine stope. These technologies include mining machinery, blasting operations, fires,
ventilation, and lighting. However, these heat sources may cause insignificant changes to the
temperature of CPB because of the considerably large size of the mine stopes and/or the CPB
structure (Fall et al. 2010).

2.4 Background of Hydraulic Binders, Its Hydration and Key Influential Factors
Hydraulic binders are the most important ingredient of any cementitious materials, such as CPB.
The main task of hydraulic binders is to increase the strength of these materials by bonding their
particles together (Landriault et al. 1997, Persson 1997). In CPB, there are quite a few types of
hydraulic binders that are usually used, such as Ordinary Portland Cement (OPC), blast furnace
slag (Slag), and Fly Ash (FA) (Haiqiang et al. 2016). An overview of these hydraulic binders, the
basic hydration reactions, and the key influential factors have been briefly described in the
following subsections.

2.4.1 Ordinary Portland Cement (OPC)


Ordinary Portland cement (OPC) is a hydraulic cement that experiences a dissolution and
precipitation while reacting with water, forming different water-resistant hydrates that are
responsible for the setting and hardening of the cement. The products of cement hydration include
poorly crystalline gel of calcium silicate hydrate (C-S-H), thin to large hexagonal crystals of
portlandite (CH), and needle-like morphological crystals of ettringite (Marchon and Flatt 2016,
Cui 2017).

2.4.2 Blast Furnace Slag (Slag)


Blast-furnace slag cement (Slag) is a pozzolanic mineral admixture that has been extensively used
in the past decades as a partial replacement of Portland Cement. In CPB operations, the utilization
of pozzolanic admixtures (such as slag) can reduce the binder cost due to cement consumption,
while maintaining the desired strength of CPB mixture (Daube and Bakker 1986, Cui 2017). The
reactivity of slag is often activated by alkalinity which results from the hydration of clinker during
the early hydration process of cement-slag blends (Roy and Idorn 1982, Daube and Bakker 1986).

2.4.3 Flay Ash (FA)


Fly Ash (FA) is a by-product pozzolanic material that is usually generated from coal firing for
power generation. FA consists of fine, powdery particles that have spherical shape. There is an
20
increasing demand of using FA as it is considered to be highly contaminating material. However,
the current utilization of FA in the form of alternative to cement and concrete products depends on
several factors, such as the deign strength and workability of concrete as well as the water demand.
As pozzolanic admixture, FA can also be activated by alkalinity, and it is significantly affected by
temperature (Palomo et al. 1999, Siddique 2004, Ahmaruzzaman 2010).

2.4.4 Hydration Process


Several researches (e.g. Schindler and Folliard 2003, Gruskovnjak et al. 2006, Bullard et al. 2011,
Kolani et al. 2012, Scrivener et al. 2015, Marchon and Flatt 2016, Le et al. 2019) have explained
the mechanism of hydration process of cement, slag, FA, and different blends of these binders. In
general, there are four periods (Figure 2.6), which the progress of cement hydration with time
passes through. The heat flow and the amount of cement hydration change with time at each stage.
These four periods are:

1) Initial reaction period, which is the period of immediate reaction of cement when exposed to
water. In this stage, significant amount of heat is released because of the rapid dissolution of
gypsum and aluminite. This period lasts up to 1 hour after mixing.
2) Slow reaction period, wherein the hydration products (C-S-H and ettringite) will be formed
around the calcium silicate. So, the magnitude of cement hydration will slowly increase, and
low rate of hydration heat will be released. This period occurs between 1 and 2 hours after
mixing.

3) Acceleration period, wherein a significant amount of hydration heat will be released. Thus,
the hydration process rapidly increases. In this period, the peak level of hydration process is
reached, the pore spaces will be further refined, and the bonds strength between particles will
significantly improve. This period occurs between 2 to 9 hours after mixing.

4) Deceleration period, wherein the heat evolution rate will decrease because the hydration
products (C-S-H and CH) will continue to form around the particles. Hence, there will be a
noticeable reduction in cement hydration. This period begins 9 hours after mixing and
continue until the hydration process reaches its minimum magnitude at around 24 hours after
mixing.

When Slag is partially added to the blend, the hydration of slag begins during the initial
reaction period of the cement hydration process. The early hydration of clinker releases alkalinity,
which is responsible for activating the slag, synergistic with temperature rise and provides energy
to activate alkali-hydroxide attack on the slag particles (Kolani et al. 2012).

21
4
(1) Initial Period
(2) Slow Reaction Period
(3) Acceleration Period
(4) Deceleration Period
3
Heat Flow (mW/g)

(1) (2) (3) (4)

0
0 2 4 6 8 10 12 14 16 18 20 22 24
Hydration Time (h)

Figure 2.6 Progress of cement hydration with time (modifed from: Bullard et al. 2011).

On the other hand, when FA is partially added to the cement blend, the early hydration of
cement alite will be faster than its hydration with the pure OPC mixture, because the fine particles
of FA will result in the formation of hydration products on the surface of the FA and the thin layer
of the hydrates formed on the surface of alite (Sakai et al. 2005). However, the slow reaction of
FA within the cement-FA blend makes the heat released during the initial hydration uncountable
(Han et al. 2017).

2.4.5 Key Factors that Affect Binder Hydration


There are several factors that affect binder hydration, for example fineness of the binder material,
amount of mixing water (i.e. w/c ratio), mixing and curing temperature, and chemistry of mixing
water (Sakai et al. 2005, Bentz 2006, Lin and Meyer 2009, Fall et al. 2010, Li and Fall 2016, Wu
et al. 2016, Han et al. 2017).

Several studies (Bentz 2006, Lin and Meyer 2009) have found that fineness of cement
particles improves the mechanical strength of cemented materials as it accelerates the cement
hydration process because the finer cement particles have larger surface area. This will provide
more contact area for the mixing water that increases the amount of reaction between the water
and cement particles. In addition, the more hydration of fine cement particles, the more production

22
of less thick cement hydration products. It reduces the setting time of cemented particles, and
accelerates the reaction between water and fine cement particles. Also, beside the acceleration of
the water-binder reaction due to the increase in production of thin layers of hydrates, the fine size
and spherical shape of the FA particles will increase the packing density of the soils that are
cemented with cement-FA blend and reduce the friction forces between these particles. This will
reduce the volume of pore-water required to fill the voids between particles, and thus enhance the
workability and flowability of the cemented material by reducing its yield stress and viscosity
(Siddique 2004, Sakai et al. 2005, Kwan and Chen 2013) . On the other hand, it was found that the
presence of coarse slag particles within the cement-slag blends increases the reactivity of slag and
significantly increases the hydration rate of slag particles (Tan et al. 2014), which will
consequently increase the mechanical strength of soils cemented with cement-slag blend (Pokharel
and Fall 2011).

Similarly, the degree of cement hydration was found to be another factor that significantly
affects the hydration of soil-hydraulic binder mixes. It was found that the increase in w/c increases
the porosity and hydraulic conductivity of the cemented material, and thereby increases the
flowability (reduces the rheological properties) (Figure 2.7) and reduces the mechanical strength
of these materials (Figure 2.8) (Bentz 2006, Fall et al. 2008, Harini et al. 2011, Ramezani et al.
2016)..

10 0.025
Viscosity
9
Yield Stress
8 0.02
7
Yield Stress (Pa)

Viscosity (Pa.s)
6 0.015
5
4 0.01
3
2 0.005
1
0 0
0.35 0.45 0.6
w/c Ratio
Figure 2.7 Effect of water-to-cement (w/c) ratio on the rheological properties of self compacted concrete
(modifed from: Ramezani et al. 2016).

23
1700
W/C = 6
1600 W/C = 7
W/C = 8
1500
W/C = 9
1400
1300
UCS (kPa)

1200
1100
1000
900
800
700
3.0 3.5 4.0 4.5 5.0
Cement Content (%)
Figure 2.8 Effect of water-to-cement (w/c) ratio on the mechanical strength of cemented Paste Backfill
(modifed from: Fall et al. 2008).
The initial (mixing and curing) temperature of cementitious material is considered to be
another important factor that influences the hydration process. It has been found that high initial
and/or curing temperature will reduce the pore spaces between particles and reduce the water
content. Thus, the cement hydration process will accelerate, and more cement hydration products
will be generated within the pores between the particles. Consequently, the high initial and/or
curing temperature will improve the mechanical strength of cemented soil at early ages (Figure
2.9) (Lin and Meyer 2009, Fall et al. 2010, Wang et al. 2016). Moreover, the higher temperature
has greater effect on the reactivity of slag, as compared to other hydraulic binders, and it can also
enhance the pozzolanic reaction of FA (Deschner et al. 2013, Ogirigbo and Black 2016). However,
the increase in initial temperature will negatively affect the flowability of cemented materials by
increasing its rheological properties (Figure 2.10), as it will cause significant reduction in the water
content of these materials (Haiqiang et al. 2016, Haruna and Fall 2019). Also, the increase in curing
temperature to more than 85oC will significantly reduce the mechanical strength of cementitious
material because this range of curing temperature will lead to less uniform microstructure, higher
porosity, and coarser pore structure within the soil particles (Maltais and Marchand 1997, Elkhadiri
et al. 2009).

24
4000
7 days
3500 28 days

3000

2500
UCS (kPa)

2000

1500

1000

500

0
0 10 20 30 40 50 60
Temperature (oC)
Figure 2.9 Effect of curing temperature on the strength development of cemented Paste Backfill (modified
from: Fall et al. 2010).

Figure 2.10 Effect of curing temperature on the rheological properties of cemented Paste Backfill: (a)
Yield stress, and (b) Viscosity (modifed from: Haruna and Fall 2019).

Hydration process of different hydraulic binders was found to be significantly influenced


by the chemistry of mixing water. The presence of uncontrolled chemical components (such as
sulphate) might slowdown the hydration process because of the reaction between the sulphate
anions in the mixture and C3A grains within the binder which will form a thin coating of
anhydrated cement particles and prevent the C3A from quickly reacting with water (Li and Fall

25
2016, Alainachi and Fall 2020). On the other hand, the presence of other chemical components
(such as sodium silicate) might increase the rate of the cement hydration process and improve the
mechanical strength of cementitious soils (Abdul-Hussain and Fall 2012). It was also found that
sodium silicate potentially contributes to the activation of pozzolanic binders, such as slag and FA
(Jawed and Skalny 1978, Qing-Hua et al. 1992, Palomo et al. 1999, Gruskovnjak et al. 2006,
Cihangir et al. 2018).

2.5 Sodium Silicate


The main components of alkaline binders were discovered in 1967. Since then, many studies have
been conducted to understand the activation role of these activators (Garcia-Lodeiro et al. 2015).
Soluble sodium silicates, also known as water soluble glass, are silicate polymer, which are an
inorganic chemical, and a clear, colorless, and viscous liquid. They are classified as Alkali
activators that are generally produced by varying proportions of alkali metal and SiO2 (Razavi and
Hassani 2007). The main application of sodium silicate, from geotechnical point of view, is to
reduce the water content and/or the w/c ratio of the soil grout and mine tailings backfill. This will
reduce the quantity of free water in mine stopes, and thereby improve the strength of the backfill
(Abdul-Hussain 2011). Also, adding sodium silicate to cemented material helps in lowering the
threat of thermal cracking as it produces lower hydration heat as compared to cement (Hassani et
al. 2007). Nevertheless, there is lack of information on the effect of sodium silicate on the rheology
and mechanical strength of soils, potentially tailings, which are cemented with different hydraulic
binders.

In the following subsection, the effect of sodium silicate on the hydration process of
different hydraulic binders is briefly discussed.

2.5.1 Cement-Sodium Silicate Interaction


Mixing cementitious material with alkaline activators (such as sodium silicate) results in alkaline
attack on the aluminosilicate of these materials. This will form the alkaline cement, which
potentially sets and hardens, and thereby improves the binding properties of these materials
(Garcia-Lodeiro et al. 2015). Sodium silicate is widely used as an accelerator and low-density
lightning admixture in Portland cement. It is also used as self-healing agent in self-healing cement
(Guo et al. 2019). Moreover, it was found that rate of hydration of the cement clinker minerals
significantly increases with the presence of sodium silicate, because adding sodium silicate to
cemented mixture will result in the reduction of the initial pH of the pore-solution of cement paste
and potential generation of hydration products (such as CH) (Brykov et al. 2002).

2.5.2 Slag-Sodium Silicate Interaction


The initial hydration process of slag-water mixture is much slower than the initial hydration
process while mixing Portland cement with water. However, the rate of hydration of mixing water
with cement-slag blend depends on the dissolution of the glassy structure of slag during the initial
hydration of cement (Sajedi and Razak 2010). There is a common agreement that adding sodium

26
silicate to pozzolans (such as slag) will effectively activate these pozzolans and accelerate its
hydration process, because the early hydration process of cement clinker releases alkali and
increases temperature, which will provide high energy to activate the alkali-hydroxide attacks on
the slag particles (Roy and Idorn 1982, Kolani et al. 2012, Kermani et al. 2015, Azadi et al. 2017).
Moreover, Zhao et al. (2020) found that the hydration and mechanical strength of slag are
significantly affected by adding sodium silicate, specially at the early age, as it shortens the setting
time, increases the production of C-S-H, and enhances the compressive strength. However, it was
reported that the hydration process, the production of hydration products and consequently the
strength of the slag-cement pastes activated by alkali activators (such as sodium silicate) are
influenced by several factors. These factors include the nature of slag and activator, water/slag
ratio, fineness of slag, dosage of activator, and curing temperature etc. (Sajedi and Razak 2010).

2.5.3 Fly Ash-Sodium Silicate Interaction

The partial replacement of Portland cement with FA has become common as it lowers the carbon
dioxide emission as a result of the manufacturing process of Portland cement. However, the
strength development of cement-FA paste, at ambient temperature conditions, was found to be
slower, producing material of relatively lower strength. Using sodium silicate as alkali activator
was found to enhance the mechanical performance of the cement-FA paste as it promotes the
condensation process of the mixture (Palomo et al. 2007, Panias et al. 2007, Rashad 2013, Phoo-
ngernkham et al. 2017). However, the cement/FA ratio and/or the type of dosage of the sodium
silicate controls the strength development of cement-FA blends activated by sodium silicate,
because the products resulted from the reaction between the active silica (from sodium silicate
solution) and the calcium (from the cement) are responsible for the strength performance (Morsy
et al. 2014, Phoo-ngernkham et al. 2017).

2.5.4 Effect of sodium silicate on CPB

The effect of adding sodium silicate, as alkali activator, on the properties of CPB was studied by
only few researchers. For instance, (Kermani et al. 2015) evaluated the effect of sodium silicate
on the strength of gelfill, which is the mixture of tailings, hydraulic binder, water, and chemical
additives (such as sodium silicate gel). They found that the UCS values of gelfill increase with
increase in the dosage of sodium silicate up to 0.3% of the total dry weight (wt%), while these
values significantly decrease when the sodium silicate content exceeds 0.5%. Also, they found that
CPB materials that contain 0.7% or more of sodium silicate have no measurable strength within
the first 14 days of curing (Figure 2.11). They attribute to the increase in the total porosity of
samples and the amount of water trapped in the samples.

Abdul-Hussain (2011) experimentally simulated the coupled behavior of the thermal,


hydraulic, and mechanical (THM) properties of CPB that contains sodium silicate by building two
columns of cemented paste tailings under drained and undrain conditions. Both columns were
cured at room temperature for 28 days, and it was found that the strength development (UCS

27
values) is significantly coupled with the heat development, saturation degree, and suction
development within the CPB that contains sodium silicate. They also found that the chemical
interaction between the cemented tailings particles, mixing water, and sodium silicate has
significant effect on the cement hydration process and rate. So, the strength development is
significantly affected by the dosage of sodium silicate (Figure 2.12).

1600
wt% Sodium silicate:
1400 0.0% 0.1% 0.2% 0.3%
0.4% 0.5% 0.7% 0.9%
1200
1000
UCS (kPa)

800
600
400
200
0
0 5 10 15 20 25 30
Curing Time (days)
Figure 2.11 Effect of sodium silicate content on the strength development of CPB (modified
from:Kermani et al. 2015).

800
700
600
500
UCS (kPa)

400
300 0.1% Sodium Silicate
200 0.3% Sodium Silicate
100 0.2% Sodium Silicate
0.4% Sodium Silicate
0
0 10 20 30 40 50 60
Curing Time (days)
Figure 2.12 Effect of sodium silicate content on the strength development of CPB (modified from: Abdul-
Hussain 2011).

28
2.6 Rheology

Rheology, in general, refers to the deformation and flow of a material in terms of the relationship
between stress, strain, and time (Barnes et al. 1989). From civil engineering point of interest,
rheology can provide valuable and practical information about the properties of engineered fluids
and admixtures (such as concrete, mortars, cement based suspensions, cemented slurries, and
cemented paste backfill) that exhibit a time-dependent response to stress (Bannister 1980, Struble
et al. 1998, Wallevik and Wallevik 2011). Rheology is also considered as a key factor for
determining the flowability of these materials, which is a potential factor in the
transportation/delivery systems of cemented slurry. If transported material does not have
appropriate flowability, clogging of the pipelines might occur, which will lead to possible flow
delays and/or temporary interruption of the progress of these materials as well as possible
replacement of the pipelines. Hence, the lack of understanding of the rheological properties of the
transported material might lead to significant financial losses (Bannister 1980, Li and Fall 2016,
Haruna and Fall 2019). The scientific approach in assessing the rheology, and thus the workability
and flowability of cemented slurries, is based on measuring the yield stress and viscosity of these
materials (Szecsy 1997).

In the following subsection, an overview of the yield stress and viscosity and a brief
discussion on the previous research that studied the rheology of CPB have been presented.

2.6.1 Yield Stress


Many fluids often show plastic behavior, in which they require a certain level of stress to initiate
flow. This level of stress is known as “Yield Stress” (Struble et al. 1998). In other words, the term
“yield stress” refers to minimum pressure or shear stress that is required to overcome the static
friction of the fluid and make it flowable (Sofra 2017). In CPB technology, the yield stress of the
backfill should range between 200 and 700 Pa to better work within the operational pipeline
transportation system (Figure 2.13) (Potvin et al. 2005, Belem and Benzaazoua 2008).

There are several (indirect and direct) approaches to determine the shear stress of the
cemented slurries. The indirect methods, which are based on the interpretation of fundamental
shear stress-shear rate data and obtain the shear stress in the limit of zero shear rate, include the
direct extrapolation of the rheological shear stress, and shear rate data as well as the extrapolation
of the flow curves. While the direct methods rely on measurement of the shear stress, at which
flow first begins. These methods are experimentally conducted under the condition of (a)
controlled shear stress, at which deformation of material is observed as unction of time when
exposed to constant shear stress, or (b) controlled shear rate, at which the shear stress-time
response is measured while exposing the material to low and constant shear rate. The most
common direct methods in engineering practice are the stress relaxation method and vane shear
method (Dzuy and Boger 1983). In the current study, Vane shear method was adopted to determine
the yield stress of CPB.

29
1200
1100 Binder Content:
3.0 wt% 3.5 wt%
Predicted Yield Stress (Pa)
1000
900 4.0 wt% 4.5 wt%
800 5.0 wt% 7.0 wt%
700
600
500 CPB Flowable Zone
400
300
200
100
0
74 74.5 75 75.5 76 76.5 77 77.5 78
CPB Solid Mass Concentration (wt%)
Figure 2.13 Predicted yield stress and flowable zone of CBP of different binder and solid content
(modified from: Belem and Benzaazoua 2008).
Vane shear strength testing is considered as one of the most efficient and accurate methods
among the direct measuring methods for the yield stress of soil and Newtonian fluids, which can
be performed in the field and/or laboratory (Dzuy and Boger 1985, Abd Elaty and Ghazy 2012).
The yield stress values provided by vane shear methods were found to be greatly synchronized
with the most of the other rheological testing methods. In the recent years, this method has been
extensively used for different construction materials, such as the concrete and mortars (Bauer et
al. 2007). The vane shear test can be summarized by inserting a four-bladed vane (H:D=2:1) in the
end of the remolded cylindrical sample (Figure 2.14). Blades rotation is then applied by a
calibrated torque spring (or electrical torque transducer) that is attached directly to the vane (Figure
2.15), and the critical torque that can cause shear in the cylindrical surface of the sample is
determined. By converting the torque to a unit shearing resistance, the yield stress can be
calculated.

In the current study, a Wykeham-Farrance vane shear (Figure 2.16) has been used to
determine the yield stress of the CPB samples. The vane shear consists of an electrical motor that
rotates the four-bladed vane with a diameter of 2.5 cm and the height of 2.5 cm, which is connected
to the deflection spring and a dial gage to measure the resistance of the sample to the applied
torque. The test was performed in accordance with ASTM D4648/D4648M-16 (2015). Sample
preparation and the test procedure are discussed in detail in Chapter 3 (Paper I) of this dissertation.

30
Figure 2.14 Schematic view of the Vane shear test method (modified from: Bauer et al. 2007).

Figure 2.15 Vane torque spring and electrical transducer (ASTM D4648/D4648M-16 2015).

31
Figure 2.16 Vane shear apparatus used in this study.

2.6.2 Viscosity
Theoretically, viscosity can be defined as the measure of the internal friction of a material and its
resistance to flow (Barnes et al. 1989). It might also refer to the force per unit area that produces a
motion of a fluid (Szecsy 1997). In CPB technology, the viscosity represents the ratio between the
shear stress and the shear rate at a specific shear rate, and it governs the dynamic flow
characteristics of the tested material as it measures the resistance of the material against the shear
stress-induced flow (Ercikdi et al. 2017, Bian et al. 2019).

There are different empirical methods for determining the viscosity of fluid suspensions,
such as the oscillating-vessel method, at which the viscosity is measured by conducting an
analytical mathematical model to back calculate the viscosity from the damping of the vessel
(Roscoe 1958). However, there are several factors that might result in discrepancy of the results
determined by this method, such as the surface tension and slippage at the liquid/vessel interface
etc.(Cherne and Deymier 1998), as well as the possible uncertainty of results obtained by these

32
methods as they are operator-sensitive. Accordingly, there is a great need for adopting new
methods that relay on fundamental physical quantities, rather than the methods that are based on
the measuring apparatus (Tattersall and Bloomer 1979, Tattersall and Banfill 1983). For this
purpose, an advanced viscosity testing apparatus was developed by the community of concrete
science, which is named “Viscometer”. In general, the viscometer measures the shear rate by
observing the torque as a function of the predetermined rotational speed. Thus, the viscosity is
determined as the ratio of the shear stress to the shear rate at the specific shear rate (Sofra 2017).
There are several types of viscometer; couette viscometer, coaxial cylinders, and Brookfield digital
viscometer (which has been used in the current study). The Brookfield digital viscometer (Figure
2.17) consists of a rotating spindle connected to a calibrated spring. To measure the viscosity, the
spindle is immersed into the fluid to be tested and is then rotated in the liquid. The liquid will then
resist the rotation, which is known as viscous drag that is determined by the spring deflection. The
spring deflection is measured by the rotary transducer (Brookfield 2014). Sample preparation and
the test procedure are discussed in detail in Chapter 3 (Paper I) of this dissertation.

Figure 2.17 Vane shear apparatus used in this study.

33
2.6.3 Rheology of CPB
The rheology of CPB has received great attention in the literature, as it provides good indication
of the workability and flowability of the backfill being transported through the delivery system to
the underground mine stops. For instance, Huynh et al. (2006) studied different factors that might
affect the yield stress of CPB, and they found that the yield stress of CPB is affected by the
percentage of solid concentration. It was found that the CPB mix that is prepared with solid
concentration of 70 – 85 wt% produces high yield stress that makes it difficult to transport in real
plant operations. It was also found from this study that adding high concentration of chemical
additives, such as polyphosphate, is required to disperse the paste to a suitable level of flowability.

Belem and Benzaazoua (2008) developed predictive models to compare different mix
designs of CPB, based on the practical experience of the authors and some empirical methods, by
determining the rheological properties, binder content, and soil mass concentration. The result of
this study predicted that the shear yield stress of CPB material increases with the increase in the
solid concentration and/or the increase in the binder content. However, the study found that
extensive increase in these components (such as 77 (wt%) of solid mass concentration and 7% of
binder content) will lead to a non-flowable CPB mix, as the magnitude of yield stress will fall
beyond the CPB flowable zone.

Zhang et al. (2010) discussed the effect of superplasticizer on the workability and initial
setting of CPB, and found that the yield stress of CPB increases as the time progresses due to
cement hydration. However, it was also noted that by adding superplasticizer, the initial setting
time of CPB material will be delayed, that will slow down the increase in yield stress.

Kwan and Chen (2013) discussed the benefits of adding fly ash (FA) to the CPB mixture in terms
of improving the packing density, flowability, and strength. The results of this study showed that
the cement-FA blends significantly increase the packing density of CPB, which causes large
reduction in void ratio. This reduction in void ratio would potentially reduce the amount of water
needed to fill these voids that increases the amount of unused water, which can coat the solid
particles and act as lubricant. This will thereby reduce the viscosity and yield stress of CPB.

Simon and Grabinsky (2013) assessed the apparent yield stress of CPB material under the
effect of the variation of binder type and content, presence of different chemical admixtures
(superplasticizer), and pore fluid chemistry (such as pH and zeta potential). It was learnt from this
study that the apparent yield stress increases with the partial usage (30%) of FA, as the mixture
became denser. Also, the usage of chemical admixtures decreases the apparent yield stress, but the
amount of reduction is significantly affected by the type and concentration of these admixtures.
Moreover, it was noted that the apparent yield stress decreases at low pH and with the increase in
magnitude of zeta potential.

Haiqiang et al. (2016) and Jiang and Fall (2017) studied the development of yield stress of
CPB in sub-zero environment. Haiqiang et al. (2016) found that the yield stress of CPB will

34
significantly decrease when the CPB is prepared and cured at -6oC as compared to the CBSs that
are prepared and cured at 23.5oC, and the further decrease in ambient temperature causes more
reduction in the yield stress of CPB. Also, the yield stress in sub-zero condition will furtherly
decrease with the reduction in binder content, increase in w/c ratio, and/or the partial usage of FA.
While the partial usage of slag will increase the yield stress after 1 hour of curing time, Jiang and
Fall (2017), on the other hand, found that salinity of CPB mixture (in sub-zero conditions)
decreases the yield stress, as the yield stress decreases with the increase in the NaCl concentration.

The yield stress and apparent viscosity of CPB were investigated by Panchal et al. (2018)
as a function of hydration age, binder content, and superplasticizer dosage, and it was found that
at a given shear rate, the yield stress and apparent viscosity decrease with the increase in hydration
age for specific binder content and concentration of superplasticizer.

Haruna and Fall (2019) studied the time and temperature effect on the rheological
properties of CPB that contains superplasticizer, and they found that yield stress and viscosity of
CPB increase with the increase in the initial (mixing and curing) temperature, and/or the reduction
in the superplasticizer concentration. It was also noted that the yield stress and viscosity are also
affected by the change in tailing type with different dosages of superplasticizer and different curing
temperatures.

Xiapeng et al. (2019) investigated the sulphate-induced changes of the yield stress and
viscosity of CPB. The results show that the rheological properties decrease with the increase in
sulphate concentration, regardless it was prepared with silica tailings (ST) or natural tailings (NT).
Also, the increase in binder content will increase the yield stress and viscosity, regardless of the
presence of sulphate.

Kou et al. (2020) experimentally investigated the time-dependent rheological behavior of


CPB that contains alkali-activated slag as a binder. It was found that both yield stress and viscosity
gradually increase with progress of curing time and the increase in binder content (up to 8%).
Further increase in binder content may cause reduction in initial yield stress and viscosity.
Moreover, it was found that the increase in the fineness of slag promotes the reactivity of slag,
which will produce more cement hydration products and consume more pore-water, and thus
increase the rheological properties of CPB that contains alkali activated slag.

Roshani and Fall (2020) found that the addition of non-silica particles to the CPB
admixture or increasing the dosage of the non-silica particles will increase the yield stress and
viscosity. Also, the increase in the rheological properties of CPB in the presence of non-silica
particles becomes more significant with the increase in curing time and temperature. Moreover,
the increase in these properties of CPB that contains non-silica particles is dependent on the binder
type, as the yield stress of CPB that is made with PCI increases (with the increase in the dosage of
non-silica particles) at faster rate as compared to CPBs made of cement-Slag and cement-FA
blends.

35
In the current study, the rheological properties (yield stress and viscosity) of CPB were
investigated under the combined effect of curing time, initial (mixing and curing) temperature,
different binder (PCI, PCI-Slag, and PCI-FA) blends, and different dosages of alkali activator
(sodium silicate). To the best knowledge of the author, the approach of studying the coupled effect
of these factors has not yet been addressed in the literature. Further details of the experimental
program adopted in this research as well as the methodology, experimental conditions, and
instrumentations are briefly described in the Chapter 3 of this dissertation.

2.7 Compressive Strength

2.7.1 Background
The Unconfined Compression Strength (UCS) test, also known as uniaxial compression test, is a
simple and effective method to assess the deformability and behavior of a material against
compressive loading. UCS is generally determined by subjecting samples to compressive load
along a single (usually vertical) axis and recording the compressive displacement (change in
dimensions) of the sample along that axis (Wittke 2014, Pan 2019). It was previously found that
the unconfined compression of a soil might vary based on the material itself (such as the cemented
or non-cemented material and the saturation condition of the material etc.) and the testing
characteristics and conditions, such as performing the test under stable or unstable confining stress,
drained or undrained conditions, and static or cyclic compressive loading etc. (Mitachi et al. 2001,
Chae et al. 2010, Fredlund et al. 2012).

There are several approaches, where the UCS can be conducted. These approaches are
mainly based on the type of the tested material. For instance, ASTM D2166/D2166M - 16 is
usually used for determining the UCS of cohesive soils, ASTM D5102-09 is used for compacted
soil-lime mixture, ASTM D2938-95 is used for intact rock specimens, and ASTM C39/C39M is
often used for determining the UCS for cylindrical concrete specimens. In this study, the
unconfined compression strength tests of CPB material were conducted according to ASTM
C39/C39M (2018), computer-controlled mechanical press of ELE Digital Tritest (Figure 2.18),
and the tests were conducted with loading capacity and deformation rate of 50 kN and 1 mm/min
respectively. The tested specimen had a height-to-diameter ratio of 2:1. Before conducting the test,
the two ends of the samples were first flattened to get their surfaces normal to the plates of the
mechanical press. Sample preparation and the detailed test procedure are discussed in Chapter 4
(Paper II) of this dissertation.

36
Figure 2.18 Unconfined Compression Strength (UCS) testing apparatus used in this study.

2.7.2 Compressive Strength of CPB


In CPB technology, UCS has become a widely accepted and important property for controlling the
quality of CPB (Chen et al. 2019). However, the strength development of CPB is affected by
several factors, such as progress of cement hydration process and intensity of cement self-
desiccation (Ghirian and Fall 2016). Furthermore, the strength of CPB is significantly governed
by the change in backfill temperature, type of hydraulic binder and binder blend ratio used in the
CPB mix, and the presence of chemical additives and/or activators (such as sodium silicate) (Fall
and Benzaazoua 2005, Fall et al. 2010, Nasir and Fall 2010, Cui and Fall 2016, Li and Fall 2018,
Xue et al. 2018).

The UCS of CPB has been extensively discussed in the literature. For instance, Belem and
Benzaazoua (2004) mentioned that the UCS of CPB to suitably perform as ground support of
underground mines should be at least 5000 kPa.

Benzaazoua et al. (2004) used the UCS test as a tool to better understand the hardening
processes of CPB, and they found that the strength development of CPB is effected by the tailings
properties, water-to-binder ratio, binder type and content, and the chemistry of the water used in

37
the CPB mix. The result of this study demonstrated that the grain size distribution of tailings
critically influences the hardening process, and thus the strength, of CPB. The increase in water
content may affect the hardening process by effecting the precipitation reactions during the
cementing phase. Also, different binders were found to give different strength development, and
the presence of some chemical components (such as sulphate) may negatively affect the strength
of CPB.

Kesimal et al. (2005) conducted the UCS tests on different CPB samples, and it was found
that the short-term strength of CPB tends to increase with the decrease in water-to-cement ratio
and the long term CPB strength.

Fall et al. (2010) discussed the relationship of UCS development of CPB with the curing
temperature, binder type, and curing time. They noted that the UCS values increase with the
progress of curing time, and the increase in curing temperature accelerates the cement hydration
process and thereby increases the UCS values. They also found that using PCI-slag blends as a
binder will yield higher UCS values as compared to PCI only, while the usage of PCI-FA blend
will reduce these values.

Nasir and Fall (2010) developed a numerical model for predicting the UCS of undrained
CPB. In their model, the strength development was coupled with temperature and degree of
hydration, and they found that different mixes of CPB have different strength development with
respect to different temperatures.

Abdul-Hussain (2011) used UCS as a tool to determine the right dosage of sodium silicate
to prepare a suitable Gelfill (sodium silicate-activated CPB mixtures), and she found that 0.4% wt
would be the best proportion of sodium silicate as it results in higher UCS values.

Yilmaz et al. (2011) assessed the UCS of consolidated and unconsolidated CPB materials.
The UCS results showed that the consolidated backfills consistently show higher strength than the
unconsolidated ones.

Yin et al. (2012) studied the effect of solid components on the mechanical properties of
CPB, and found that the magnitude of UCS of CPB samples without slag is almost two times larger
than the samples prepared with adding 25% of slag. It was also noted that unground slag would
also diminish the mechanical properties of CPB, while the addition of sodium silicate would
efficiently help achieve the desired initial strength performance of CPB.

The results of (Ghirian and Fall 2014) showed that the time-dependent increase in UCS of
CPB is due to the combined effect of the progress of cement hydration and suction development.
Also, the UCS results indicated that the CPB material in the long term (higher curing time) shows
higher UCS values and more brittle behavior as compared to those at early age.

38
Wang et al. (2016) experimentally investigated the influence of initial temperature on the
strength development and self-desiccation in cemented paste backfill that contains sodium silicate.
In this study, the mechanical strength, particularly UCS, was tested for CPB that contains sodium
silicate as the admixture at early age under the effect of various initial temperatures (2 oC, 20oC,
35oC, and 50oC). Beside the effect of sodium silicate on activating the hydration of CPB material,
this study found that the high initial temperature leads to more intense rapid self-desiccation in the
CPB material. It was also found that the rate of strength development at early age of CPB material
that contains sodium silicate is significantly enhanced at high initial temperature. However, the
overly high initial temperature was found to reduce the compressive strength of CPB that contains
sodium silicate after 7 days of curing age (Figure 2.12). Wu et al. (2016), on the other hand,
investigated the effect of initial temperature on the deformation (stress-strain) behavior of CPB
that contains sodium silicate. They also evaluated modulus of elasticity as a function of the UCS
and the degree of hydration. They found that the high initial temperature-induced increase in UCS
and hydration degree leads to an increase in the modulus of elasticity, which is not directly affected
by the initial temperature.

1200
2°C
20°C
1000
35°C
800 50°C
UCS (kPa)

600

400

200

0
0 5 10 15 20 25 30
Curing Time (days)
Figure 2.19 Effect of initial temperature on the strength development of CPB that contains sodium silicate
at early age (modified from: Wang et al. 2016).

Cihangir et al. (2018) conducted UCS tests to determine the short- and long-term strength
of CPB materials that are prepared with different binder and/or dosage of sodium silicate. The
results showed that Portland cement-CPBs have higher strength at early age (<14 days) and lower
strength at longer age (around 360 days) as compared to sodium silicate activated slag-CPBs.

39
Jiang et al. (2019) also studied the UCS of CPB that is prepared with different dosages of
alkali-activated slag as binder, and they found that the increase in the binder dosage would yield
significant increase in the UCS values.

Jafari (2020) conducted UCS test on CPB samples to investigate the change in UCS with
curing time and cement content. The UCS results showed that the UCS values of CPB generally
increase with the increase in curing time. However, the rate of increase is potentially higher at
early ages and it gets lower with the passage of time, which was attributed to the fact that with the
passage of time, most of the cement particles have been used to gain strength. It was also found
that the UCS increases with the increase in cement content.

In the current study, the UCS of Slag-CPB (CPB that was prepared with PCI-Slag blend)
was investigated under the combined effect of curing time, initial (mixing and curing) temperature,
and different dosages in alkali activator (sodium silicate). To the best knowledge of the author, the
approach of studying the coupled effect of these factors has yet not been addressed in the literature.
Further details of the experimental program adopted in this research as well as the methodology,
experimental conditions, and instrumentations are briefly described in the Chapter 4 of this
dissertation.

2.8 Conclusions
Underground mining operations were found to cause potential geotechnical problems, such as
ground subsidence, and produce significant amounts of anti-ecological wastes, such as tailings.
Traditionally, these tailings are stocked as surface structures (e.g. tailing dams), but they were
found to be exposed to more environmental challenges (such as acid mine drainage) and
geotechnical issues (such as tailing dam failure), which can cause significant environmental and
economic damages and lead to great amount of casualties in the mine area and the surrounding
areas. Over the past few decades, cemented paste backfill (CPB) was found to be a successful
modern method of tailing management around the world, as it allows around 60% of the hazardous
mining waste to be returned into the underground mine opening and reused as geotechnical support
for the underground mine openings (stope). CPB is a mixture of tailings and water (fresh and/or
mine processing) that is undergoing cementation, as it contains different types of hydraulic binder
(such as Portland cement, slag, and fly ash). CPB mixture is often prepared in paste plants located
in the mine area, and the fresh CPB mixture is delivered through transportation pipeline system
into the mine stope. So, it is important for the CPB mixture to have suitable flowability in order to
avoid pipelines clogging, which might cause production delay and/or possible damages to the
transportation system that leads to significant financial losses. In addition, CPB mixtures have to
have suitable mechanical strength to achieve its desired performance.

From the technical background information, it was learnt that there are several sources of
heat that can affect the temperature of CPB in the delivery pipelines when placed in the mine
stopes. This heat potentially influences the rheological properties (and hence the flowability) of
transported CPB as well as the mechanical behavior of CPB in the mine stope. Also, it was

40
understood that using different binder blends in the CPB mixtures can significantly change the
rheological properties and the mechanical response of CPB. In addition, using chemical additives,
such as sodium silicate, has potential influence on the behavior of CPB at early age.

From the reviewed studies, it can be observed that many researchers studied the rheological
properties, potentially yield stress and viscosity, of CPB. These studies investigated the effect of
initial temperature, binder content, or the usage of chemical additives on the rheological properties
of CPB. However, these studies only investigated the individual or coupled effect of some of these
conditions. Since no study till now (as per the knowledge of the author) has assessed the combined
effect of all these conditions on the rheological properties of CPB, so there is a great need to assess
the rheological properties of fresh CPB under the effect of initial temperature, binder type, and the
presence and initial content of sodium silicate. Furthermore, similar to the rheological properties
of the fresh CPB, the mechanical strength of CPB might also be affected by some parameters (such
as the initial temperature and type and content of chemical additives.). There is, therefore, a great
need to understand the combined effect of these parameters on the mechanical strength of fresh
CPB in order to fill the gap of previous studies. Also, a better understanding will contribute to
more safe and efficient design of CPB structures at early age.

2.9 References

Abd Elaty, M. A. and Ghazy, M. F. (2012) 'Flow properties of fresh concrete by using modified
geotechnical Vane shear test', HBRC Journal, 8(3), 159-169.

Abdul-Hussain, N. (2011) Experimental Study on the Engineering Properties of Gelfil, Masters


thesis, University of Ottawa.

Abdul-Hussain, N. and Fall, M. (2012) 'Thermo-hydro-mechanical behaviour of sodium silicate-


cemented paste tailings in column experiments', Tunnelling and Underground Space
Technology, 29, 85-93.

Ahmaruzzaman, M. (2010) 'A review on the utilization of fly ash', Progress in Energy and
Combustion Science, 36(3), 327-363.

Alainachi, I. and Fall, M. (2019a) 'Liquefaction potential of cementing tailing backfill: shaking
table test results', in Proceedings of 16th Panamerican conference on soil mechanics and
geotechnical engineering, Cancun, Mexico, IOS Press, 1949-1956.

Alainachi, I. and Fall, M. (2019b) 'Shaking table testing of liquefaction potential of cemented paste
backfill: Effects of backfill temperature', in Geo St-John's 2019 – the 72th Canadian

41
Geotechnical Conference (CGC), Sept. 29- Oct 2, 2019, St. John's, NL, Canada, September
29 - October 2, 2019,

Alainachi, I. and Fall, M. (2020) 'Chemically Induced Changes in the Geotechnical Response of
Cementing Paste Backfill in Shaking Table Test', Journal of Rock Mechanics and
Geotechnical Engineering.

Alainachi, I. H. (2020) Shaking Table Testing of Cyclic Behaviour of Fine-Grained Soils


Undergoing Cementation: Cemented Paste Backfill, Doctoral Dissertation, University of
Ottawa.

Alakangas, L., Dagli, D. and Knutsson, S. (2013) Literature review on potential geochemical and
geotechnical effects of adopting paste technology under cold climate conditions, Luleå
tekniska universitet.

Aldhafeeri, Z. (2018) Reactivity of Cemented Paste Backfill, University of Ottawa.

Aldhafeeri, Z. and Fall, M. (2016) 'Time and damage induced changes in the chemical reactivity
of cemented paste backfill', Journal of Environmental Chemical Engineering, 4(4), 4038-
4049.

Aldhafeeri, Z., Fall, M., Pokharel, M. and Pouramini, Z. (2016) 'Temperature dependence of the
reactivity of cemented paste backfill', Applied Geochemistry, 72, 10-19.

ASTM C39/C39M (2018) 'Standard test method for compressive strength of cylindrical concrete
specimens',

ASTM D4648/D4648M-16 (2015) 'Standard test method for laboratory miniature vane shear test
for saturated fine-grained clayey soil',

Azadi, M. R., Taghichian, A. and Taheri, A. (2017) 'Optimization of cement-based grouts using
chemical additives', Journal of Rock Mechanics and Geotechnical Engineering, 9(4), 623-
637.

Bannister, C. E. (1980) 'Rheological evaluation of cement slurries: Methods and models', in SPE
Annual Technical Conference and Exhibition, Society of Petroleum Engineers.,

Barnes, H. A., Hutton, J. F. and Walters, K. (1989) An introduction to rheology, Elsevier.

42
Bauer, E., de Sousa, J. G. G., Guimarães, E. A. and Silva, F. G. S. (2007) 'Study of the laboratory
Vane test on mortars', Building and Environment, 42(1), 86-92.

Belem, T. and Benzaazoua, M. (2004) An overview on the use of paste backfill technology as a
ground support method in cut and fill mines, translated by Perth, Australia: CRC Press,
637–650.

Belem, T. and Benzaazoua, M. (2008) Predictive models for prefeasibility cemented paste backfill
mix design, translated by Nancy, France.

Bentz, D. P. (2006) 'Influence of water-to-cement ratio on hydration kinetics: Simple models based
on spatial considerations', Cement and Concrete Research, 36(2), 238-244.

Benzaazoua, M., Fall, M. and Belem, T. (2004) 'A contribution to understanding the hardening
process of cemented pastefill', Minerals Engineering, 17(2), 141-152.

Bian, J., Fall, M. and Haruna, S. (2019) 'Sulphate induced changes in rheological properties of
fibre-reinforced cemented paste backfill', Magazine of Concrete Research, 1-35.

Brookfield (2014) 'Brookfield Dial Reading Viscometer with Electronic Drive: Model DV‐E
Operating Instructions.', [online], available: https://pim-
resources.coleparmer.com/instruction-manual/98945-xx.pdf].

Brykov, A., Danilov, V., Korneev, V. and Larichkov, A. (2002) 'Effect of hydrated sodium
silicates on cement paste hardening', Russian journal of applied chemistry, 75(10), 1577-
1579.

Bull, A. (2019) Temperature Dependence of the Leachability of Cemented Paste Backfill, M.A.Sc.
Thesis: University of Ottawa.

Bullard, J. W., Jennings, H. M., Livingston, R. A., Nonat, A., Scherer, G. W., Schweitzer, J. S.,
Scrivener, K. L. and Thomas, J. J. (2011) 'Mechanisms of cement hydration', Cement and
Concrete Research, 41(12), 1208-1223.

Chae, J., Kim, B., Park, S.-w. and Kato, S. (2010) 'Effect of suction on unconfined compressive
strength in partly saturated soils', KSCE Journal of Civil Engineering, 14(3), 281-290.

43
Chang, S. (2016) Strength and Deformation Behaviour of Cemented Paste Backfill in Sub-zero
Environment, M.A.Sc. thesis: University of Ottawa.

Chen, X., Shi, X., Zhou, J., Du, X., Chen, Q. and Qiu, X. (2019) 'Effect of overflow tailings
properties on cemented paste backfill', J Environ Manage, 235, 133-144.

Cherne, F. and Deymier, P. A. (1998) 'Calculation of viscosity of liquid nickel by molecular


dynamics methods', Scripta materialia, 39(11), 1613-1616.

Cihangir, F., Ercikdi, B., Kesimal, A., Ocak, S. and Akyol, Y. (2018) 'Effect of sodium-silicate
activated slag at different silicate modulus on the strength and microstructural properties
of full and coarse sulphidic tailings paste backfill', Construction and Building Materials,
185, 555-566.

Cui, L. (2017) Multiphysics Modeling and Simulation of the Behavior of Cemented Tailings
Backfill, Doctoral dissertation: University of Ottawa.

Cui, L. and Fall, M. (2016) 'Mechanical and thermal properties of cemented tailings materials at
early ages: Influence of initial temperature, curing stress and drainage conditions',
Construction and Building Materials, 125, 553-563.

Daube, J. and Bakker, R. (1986) 'Portland blast-furnace slag cement: a review' in Blended Cements,
ASTM International.

Deschner, F., Lothenbach, B., Winnefeld, F. and Neubauer, J. (2013) 'Effect of temperature on the
hydration of Portland cement blended with siliceous fly ash', Cement and Concrete
Research, 52, 169-181.

Dzuy, N. Q. and Boger, D. (1985) 'Direct yield stress measurement with the vane method', Journal
of Rheology, 29(3), 335-347.

Dzuy, N. Q. and Boger, D. V. (1983) 'Yield stress measurement for concentrated suspensions',
Journal of Rheology, 27(4), 321-349.

Elkhadiri, I., Palacios, M. and Puertas, F. (2009) 'Effect of curing temperature on cement
hydration', Ceram Silik, 53(2)(2), 65-75.

44
Ercikdi, B., Cihangir, F., Kesimal, A. and Deveci, H. (2017) 'Practical Importance of Tailings for
Cemented Paste Backfill' in Yilmaz, E. and Fall, M., eds., Paste Tailings Management,
Cham: Springer International Publishing, 7-32.

Fall, M. and Benzaazoua, M. (2005) 'Modeling the effect of sulphate on strength development of
paste backfill and binder mixture optimization', Cement and Concrete Research, 35(2),
301-314.

Fall, M., Benzaazoua, M. and Saa, E. G. (2008) 'Mix proportioning of underground cemented
tailings backfill', Tunnelling and Underground Space Technology, 23(1), 80-90.

Fall, M., Célestin, J. C., Pokharel, M. and Touré, M. (2010) 'A contribution to understanding the
effects of curing temperature on the mechanical properties of mine cemented tailings
backfill', Engineering Geology, 114(3-4), 397-413.

Fall, M., Nasir, O. and Celestine, J. (2007) 'Paste backfill responses in deep mine temperature
conditions', in Symposium minefill, Montreal, Canada, CD-Room,

Fall, M., Wu, D. and Pokharel, M. (2014) Effect of deep mine temperature conditions on the heat
development in cemented paste backfill and its properties, translated by Australian Centre
for Geomechanics, 559-573.

Fang, K. and Fall, M. (2018) 'Effects of curing temperature on shear behaviour of cemented paste
backfill-rock interface', International Journal of Rock Mechanics and Mining Sciences,
112, 184-192.

Fredlund, D. G., Rahardjo, H. and Fredlund, M. D. (2012) Unsaturated soil mechanics in


engineering practice, Hoboken, New Jersey, USA: John Wiley & Sons.

Garcia-Lodeiro, I., Palomo, A. and Fernández-Jiménez, A. (2015) 'An overview of the chemistry
of alkali-activated cement-based binders' in Handbook of alkali-activated cements, mortars
and concretes, Elsevier, 19-47.

Ghirian, A. and Fall, M. (2013) 'Coupled thermo-hydro-mechanical–chemical behaviour of


cemented paste backfill in column experiments. Part I: Physical, hydraulic and thermal
processes and characteristics', Engineering Geology, 164, 195-207.

Ghirian, A. and Fall, M. (2014) 'Coupled thermo-hydro-mechanical–chemical behaviour of


cemented paste backfill in column experiments', Engineering Geology, 170, 11-23.

45
Ghirian, A. and Fall, M. (2016) 'Strength evolution and deformation behaviour of cemented paste
backfill at early ages: Effect of curing stress, filling strategy and drainage', International
Journal of Mining Science and Technology, 26(5), 809-817.

Gruskovnjak, A., Lothenbach, B., Holzer, L., Figi, R. and Winnefeld, F. (2006) 'Hydration of
alkali-activated slag: comparison with ordinary Portland cement', Advances in cement
research, 18(3), 119-128.

Guo, S., Zhang, Y., Wang, K., Bu, Y., Wang, C., Ma, C. and Liu, H. (2019) 'Delaying the hydration
of Portland cement by sodium silicate: Setting time and retarding mechanism',
Construction and Building Materials, 205, 543-548.

Haiqiang, J., Fall, M. and Cui, L. (2016) 'Yield stress of cemented paste backfill in sub-zero
environments: Experimental results', Minerals Engineering, 92, 141-150.

Han, F., He, X., Zhang, Z. and Liu, J. (2017) 'Hydration heat of slag or fly ash in the composite
binder at different temperatures', Thermochimica Acta, 655, 202-210.

Harini, M., Shaalini, G. and Dhinakaran, G. (2011) 'Effect of size and type of fine aggregates on
flowability of mortar', KSCE Journal of Civil Engineering, 16(1), 163-168.

Haruna, S. and Fall, M. (2019) 'Time- and temperature-dependent rheological properties of


cemented paste backfill that contains superplasticizer', Powder Technology.

Hassani, F., Razavi, S. and Isagon, I. (2007) 'A study of physical and mechanical behaviour of
gelfill', CIM Bulletin, 2(5), 1-7.

Huynh, L., Beattie, D. A., Fornasiero, D. and Ralston, J. (2006) 'Effect of polyphosphate and
naphthalene sulfonate formaldehyde condensate on the rheological properties of dewatered
tailings and cemented paste backfill', Minerals Engineering, 19(1), 28-36.

Jafari, M. (2020) Experimental study of physical and mechanical properties of a cemented mine
tailings, Ph. D. Thesis: University of Toronto.

Jawed, I. and Skalny, J. (1978) 'Alkalies in cement: A review: II. Effects of alkalies on hydration
and performance of Portland cement', Cement and Concrete Research, 8(1)(1), 37-51.

46
Jiang, H. and Fall, M. (2017) 'Yield stress and strength of saline cemented tailings in sub-zero
environments: Portland cement paste backfill', International Journal of Mineral
Processing, 160, 68-75.

Jiang, H., Qi, Z., Yilmaz, E., Han, J., Qiu, J. and Dong, C. (2019) 'Effectiveness of alkali-activated
slag as alternative binder on workability and early age compressive strength of cemented
paste backfills', Construction and Building Materials, 218, 689-700.

Johnson, D. B. (2015) 'Biomining goes underground', Nature Geoscience, 8(3)(3), 165-166.

Kermani, M., Hassani, F. P., Aflaki, E., Benzaazoua, M. and Nokken, M. (2015) 'Evaluation of
the effect of sodium silicate addition to mine backfill, Gelfill − Part 1', Journal of Rock
Mechanics and Geotechnical Engineering, 7(3), 266-272.

Kesimal, A., Yilmaz, E., Ercikdi, B., Alp, I. and Deveci, H. (2005) 'Effect of properties of tailings
and binder on the short-and long-term strength and stability of cemented paste backfill',
Materials Letters, 59(28), 3703-3709.

Kolani, B., Buffo-Lacarrière, L., Sellier, A., Escadeillas, G., Boutillon, L. and Linger, L. (2012)
'Hydration of slag-blended cements', Cement and Concrete Composites, 34(9), 1009-1018.

Koohestani, B., Khodadadi Darban, A. and Mokhtari, P. (2018) 'A comparison between the
influence of superplasticizer and organosilanes on different properties of cemented paste
backfill', Construction and Building Materials, 173, 180-188.

Kou, Y., Jiang, H., Ren, L., Yilmaz, E. and Li, Y. (2020) 'Rheological Properties of Cemented
Paste Backfill with Alkali-Activated Slag', Minerals, 10(3).

Kwan, A. K. H. and Chen, J. J. (2013) 'Adding fly ash microsphere to improve packing density,
flowability and strength of cement paste', Powder Technology, 234, 19-25.

Landriault, D., Verburg, R., Cincilla, W. and Welch, D. (1997) Paste technology for underground
backfill and surface tailings disposal applications., Short Course notes––Technical
Workshop, Vancouver, British Columbia: unpublished.

Le, T. H. M., Park, D.-W., Park, J.-Y. and Phan, T. M. (2019) 'Evaluation of the Effect of Fly Ash
and Slag on the Properties of Cement Asphalt Mortar', Advances in Materials Science and
Engineering, 2019, 1-10.

47
Li, W. and Fall, M. (2016) 'Sulphate effect on the early age strength and self-desiccation of
cemented paste backfill', Construction and Building Materials, 106, 296-304.

Li, W. and Fall, M. (2018) 'Strength and self-desiccation of slag-cemented paste backfill at early
ages: Link to initial sulphate concentration', Cement and Concrete Composites, 89, 160-
168.

Lin, F. and Meyer, C. (2009) 'Hydration kinetics modeling of Portland cement considering the
effects of curing temperature and applied pressure', Cement and Concrete Research, 39(4),
255-265.

Lu, G. (2017) Modelling the Response of Evolutive Granular Media to Blast Loading: Cemented
Tailings Backfill, Doctoral dissertation, University of Ottawa.

Maltais, Y. and Marchand, J. (1997) 'Influence of curing temperature on cement hydration and
mechanical strength development of fly ash mortars', Cement and Concrete Research,
27(7)(7), 1009-1020.

Marchon, D. and Flatt, R. J. (2016) 'Mechanisms of cement hydration' in Science and Technology
of Concrete Admixtures, 129-145.

Mitachi, T., Kudoh, Y. and Tsushima, M. (2001) 'Estimation of in-situ undrained strength of soft
soil deposits by use of unconfined compression test with suction measurement', Soils and
Foundations, 41(5), 61-71.

Morsy, M., Alsayed, S., Al-Salloum, Y. and Almusallam, T. (2014) 'Effect of sodium silicate to
sodium hydroxide ratios on strength and microstructure of fly ash geopolymer binder',
Arabian journal for science and engineering, 39(6), 4333-4339.

Nasir, O. and Fall, M. (2010) 'Coupling binder hydration, temperature and compressive strength
development of underground cemented paste backfill at early ages', Tunnelling and
Underground Space Technology, 25(1), 9-20.

Ogirigbo, O. R. and Black, L. (2016) 'Influence of slag composition and temperature on the
hydration and microstructure of slag blended cements', Construction and Building
Materials, 126, 496-507.

48
Palomo, A., Fernández-Jiménez, A., Kovalchuk, G., Ordoñez, L. and Naranjo, M. (2007) 'OPC-
fly ash cementitious systems: study of gel binders produced during alkaline hydration',
Journal of Materials Science, 42(9), 2958-2966.

Palomo, A., Grutzeck, M. W. and Blanco, M. T. (1999) 'Alkali-activated fly ashes: a cement for
the future', Cement and Concrete Research, 29(8)(8), 1323-1329.

Pan, A. N. (2019) Mechanical Properties of Cemented Paste Backfill under Low Confining Stress,
Master of Applied Science Thesis: University of Toronto.

Panchal, S., Deb, D. and Sreenivas, T. (2018) 'Variability in rheology of cemented paste backfill
with hydration age, binder and superplasticizer dosages', Advanced Powder Technology,
29(9), 2211-2220.

Panias, D., Giannopoulou, I. P. and Perraki, T. (2007) 'Effect of synthesis parameters on the
mechanical properties of fly ash-based geopolymers', Colloids and Surfaces A:
Physicochemical and Engineering Aspects, 301(1-3), 246-254.

Persson, B. (1997) 'Self-desiccation and its importance in concrete technology', Materials and
Structures, 30(5), 293-305.

Phoo-ngernkham, T., Hanjitsuwan, S., Damrongwiriyanupap, N. and Chindaprasirt, P. (2017)


'Effect of sodium hydroxide and sodium silicate solutions on strengths of alkali activated
high calcium fly ash containing Portland cement', KSCE Journal of Civil Engineering,
21(6), 2202-2210.

Pokharel, M. and Fall, M. (2011) 'Coupled Thermochemical Effects on the Strength Development
of Slag-Paste Backfill Materials', Journal of Materials in Civil Engineering, 23(5), 511-
525.

Potvin, Y., Thomas, E. G. and Fourie, A. B. (2005) Handbook on Mine fill, Australian Centre for
Geomechanics – University of Western Australia.

Qing-Hua, C., Tagnit-Hamou, A. and Sarkar, S. L. (1992) 'Strength and Microstructural Properties
of Water Glass Activated Slag', MRS Proceedings, 245.

Ramezani, M., Kim, Y. H., Hasanzadeh, B. and Sun, Z. (2016) Influence of carbon nanotubes on
SCC flowability, translated by 397-406.

49
Rashad, A. M. (2013) 'Properties of alkali-activated fly ash concrete blended with slag', Iranian
Journal of Materials Science and Engineering, 10(1), 57-64.

Razavi, S. and Hassani, F. (2007) Preliminary Investigation Into Gel Fill: Strength Development
and Characteristics of Sand Paste Fill With Sodium Silicate, translated by American Rock
Mechanics Association.

Roscoe, R. (1958) 'Viscosity determination by the oscillating vessel method I: theoretical


considerations', Proceedings of the Physical Society, 72(4), 576.

Roshani, A. and Fall, M. (2020) 'Flow ability of cemented pastefill material that contains nano-
silica particles', Powder Technology, 373, 289-300.

Roy, D. M. and Idorn, G. M. (1982) 'Hydration, structure, and properties of blast furnace slag
cements, mortars, and concrete', ACI Journal Proceedings, 79(6)(6), 444-457.

Sajedi, F. and Razak, H. A. (2010) 'The effect of chemical activators on early strength of ordinary
Portland cement-slag mortars', Construction and Building Materials, 24(10), 1944-1951.

Sakai, E., Miyahara, S., Ohsawa, S., Lee, S.-H. and Daimon, M. (2005) 'Hydration of fly ash
cement', Cement and Concrete Research, 35(6), 1135-1140.

Schindler, A. K. and Folliard, K. J. (2003) Influence of supplementary cementing materials on the


heat of hydration of concrete, translated by.

Scrivener, K. L., Juilland, P. and Monteiro, P. J. M. (2015) 'Advances in understanding hydration


of Portland cement', Cement and Concrete Research, 78, 38-56.

Siddique, R. (2004) 'Properties of concrete incorporating high volumes of class F fly ash and san
fibers', Cement and Concrete Research, 34(1), 37-42.

Simon, D. and Grabinsky, M. (2013) 'Apparent yield stress measurement in cemented paste
backfill', International Journal of Mining, Reclamation and Environment, 27(4), 231-256.

Sofra, F. (2017) 'Rheological properties of fresh cemented paste tailings' in Yilmaz, E. and Fall,
M., eds., Paste tailings management, Cham: Springer International Publishing, 33-57.

50
Struble, L., Szecsy, R., Lei, W.-G. and Sun, G.-K. (1998) 'Rheology of cement paste and concrete',
Cement, concrete and aggregates, 20(2), 269-277.

Szecsy, R. S. (1997) Concrete Rheology, University of Illinois at Urbana-Champaign.

Tan, Z., De Schutter, G., Ye, G., Gao, Y. and Machiels, L. (2014) 'Influence of particle size on the
early hydration of slag particle activated by Ca(OH)2 solution', Construction and Building
Materials, 52, 488-493.

Tariq, A. and Yanful, E. K. (2013) 'A review of binders used in cemented paste tailings for
underground and surface disposal practices', J Environ Manage, 131, 138-49.

Tattersall, G. and Bloomer, S. (1979) 'Further development of the two-point test for workability
and extension of its range', Magazine of Concrete Research, 31(109), 202-210.

Tattersall, G. H. and Banfill, P. F. (1983) The rheology of fresh concrete.

Wallevik, O. H. and Wallevik, J. E. (2011) 'Rheology as a tool in concrete science: The use of
rheographs and workability boxes', Cement and Concrete Research, 41(12), 1279-1288.

Wang, Y., Fall, M. and Wu, A. (2016) 'Initial temperature-dependence of strength development
and self-desiccation in cemented paste backfill that contains sodium silicate', Cement and
Concrete Composites, 67, 101-110.

Wang, Y., Wu, A., Ruan, Z., Wang, H., Wang, Y. and Jin, F. (2018) 'Temperature Effects on
Rheological Properties of Fresh Thickened Copper Tailings that Contain Cement', Journal
of Chemistry, 2018, 1-8.

Wittke, W. (2014) Rock mechanics based on an anisotropic jointed rock model (AJRM), John
Wiley & Sons.

Wu, A., Wang, Y., Zhou, B. and Shen, J. (2016) 'Effect of Initial Backfill Temperature on the
Deformation Behavior of Early Age Cemented Paste Backfill That Contains Sodium
Silicate', Advances in Materials Science and Engineering, 2016, 1-10.

Wu, D., Fall, M. and Cai, S. J. (2013) 'Coupling temperature, cement hydration and rheological
behaviour of fresh cemented paste backfill', Minerals Engineering, 42, 76-87.

51
Xiapeng, P., Fall, M. and Haruna, S. (2019) 'Sulphate induced changes of rheological properties
of cemented paste backfill', Minerals Engineering, 141.

Xue, G., Yilmaz, E., Song, W. and Cao, S. (2018) 'Compressive Strength Characteristics of
Cemented Tailings Backfill with Alkali-Activated Slag', Applied Sciences, 8(9).

Yilmaz, E., Belem, T. and Benzaazoua, M. (2013) 'Study of physico-chemical and mechanical
characteristics of consolidated and unconsolidated cemented paste backfills', Gospodarka
Surowcami Mineralnymi - Mineral Resources Management, 29(1), 81-100.

Yilmaz, E., Belem, T., Bussière, B. and Benzaazoua, M. (2011) 'Relationships between
microstructural properties and compressive strength of consolidated and unconsolidated
cemented paste backfills', Cement and Concrete Composites, 33(6), 702-715.

Yin, S., Wu, A., Hu, K., Wang, Y. and Zhang, Y. (2012) 'The effect of solid components on the
rheological and mechanical properties of cemented paste backfill', Minerals Engineering,
35, 61-66.

Zhang, M.-H., Sisomphon, K., Ng, T. S. and Sun, D. J. (2010) 'Effect of superplasticizers on
workability retention and initial setting time of cement pastes', Construction and Building
Materials, 24(9), 1700-1707.

Zhao, Y., Qiu, J., Zhang, S., Guo, Z., Ma, Z., Sun, X. and Xing, J. (2020) 'Effect of sodium sulfate
on the hydration and mechanical properties of lime-slag based eco-friendly binders',
Construction and Building Materials, 250, 118603.

52
CHAPTER 3
Paper I: Time- and Temperature-Dependence of
Rheological Properties of Cemented Tailings Backfill
with Sodium Silicate

53
Paper I: Time- and Temperature-Dependence of Rheological Properties of Cemented
Tailings Backfill with Sodium Silicate

Published 2020 in the ASCE’s Journal of Materials in Civil Engineering 33(3): 04020498

Ghada Ali, Mamadou Fall and Imad Alainachi

3.1 Abstract
Cemented paste backfilling is a novel tailings management method that is widely used in the
mining industry to minimize the environmental and geotechnical risks associated with traditional
mine waste management techniques, ensure safe work conditions in underground mines, and
increase mine productivity. To enhance the early age strength of cement paste backfill (CPB),
sodium silicate has been proposed and adopted as an admixture in CPB systems due to its
activation ability. However, no studies have been done to gain insight into the effect of temperature
on the rheological properties (yield stress and viscosity) of CPB with sodium silicate and different
types of binders, although a CPB with sufficient flowability is critical for its successful application,
and CPB is subjected to various temperatures in the field during its transport. Thus, this study aims
to investigate the effect of temperature (2oC, 20oC, and 35oC) and time (0 to 4 hrs) on the
rheological properties (yield stress and viscosity) of CPB samples with different amounts of
sodium silicate (0, 0.1%, 0.3%, and 0.5%) and different binders. The results show that the yield
stress and viscosity of CPB with sodium silicate increase with time and temperature. It is also
found that the partial replacement of Portland cement with blast furnace slag or fly ash enhances
the flowability (reduces the rheological properties) of CPB, regardless of the temperature. Also,
an increase in the sodium silicate dosage will increase the yield stress and viscosity of CPB at any
temperature. The findings of this research will contribute to a more efficient design of CPB
transportation systems.

Keywords: Cemented paste backfill; Tailings; Cement; Rheology, Yield Stress, Viscosity,
Sodium Silicate, Temperature

3.2 Introduction
Mining is an important industry around the world and has been making significant contributions
to the economy in many regions and countries. In Ontario (Canada), for instance, mining is the
source of around $10 billion CAD of revenue every year (Dungan and Murphy 2012, Ontario
Mining Association 2017). However, during ore rock extraction processes, an immense quantity
of mine wastes (e.g., tailings, which are human-made soils generated by mining activities) are
produced (Aydın and Kızıltepe 2019, Esmaeili et al. 2020). Improper disposal of such wastes has
negative impacts on the environment (Dimitrova and Yanful 2011, Cihangir et al. 2012, Farkish
and Fall 2013, Ahmari et al. 2015, Alainachi and Fall 2019a, Chen et al. 2019). To manage tailings
in a more environmentally friendly and safer way, a novel technology of tailings disposal has been
developed by mixing tailings with water and binders. This disposal technology is called cemented

54
paste backfilling (Fall et al. 2005, Tariq and Nehdi 2007, Abdul-Hussain and Fall 2012, Niroshan
et al. 2017, Cihangir et al. 2018). Recently, this technology has been extensively used in
underground mining operations because the process allows for a large volume of tailings to be
reused and returned to the mine stopes, and thus, reduce the volume of such wastes that need to be
managed on the ground surface. Also, cemented paste backfill (CPB) can be produced and
delivered to the mine stopes in a short period of time. This facilitates the completion of the stope
backfilling process (stope cycles) in a matter of only days, while weeks or months might be
required with older backfilling methods. Reducing the stope cycle time allows for additional
revenue to be generated for the mine (Fall and Benzaazoua 2005, Fall et al. 2005, Kesimal et al.
2005, Ercikdi et al. 2009a, Ercikdi et al. 2009b, Thompson et al. 2009, Nasir and Fall 2010, Yin et
al. 2012, Wu et al. 2013, Haiqiang et al. 2016). In addition, using CPB can increase the stability
of the underground mine openings or rock mass which, otherwise, might be susceptible to several
geotechnical problems, such as ground subsidence, rockburst and rock mass failure (Kesimal et al.
2005). The aforementioned stability increase due to CPB is mainly a function of the mechanical
strength performance of CPB. The later can be significantly affected by several conditions or
factors, such as static and dynamic loading conditions, potential chemical interaction of the
ingredients within the CPB materials and curing temperatures (Cao et al. 2018, Xue et al. 2018,
Alainachi and Fall 2019a).

Typically, CPB mixtures consist of around 70% - 85% tailings, fresh or mine processed
water, and often 2% - 9% (by total weight of solid) of hydraulic binder (usually cement), based on
the primary and tertiary requirements of the mine stope (Alakangas et al. 2013, Yilmaz et al. 2013,
Aldhafeeri and Fall 2016, Haiqiang et al. 2016, Koohestani et al. 2018). The CBP is mixed and
prepared in a paste backfilling plant usually located on the mine surface and then delivered into
the mine stopes through pipelines either through gravity and/or pumping. Therefore, the
flowability of CPB is a very important factor in transporting CPB to the mine stopes. If the
transported material does not have suitable flowability, clogging of the pipelines might occur
which will lead to significant financial losses due to the possible flow delays and/or temporary
interruption of the progress of CPB production as well as possible replacement of the pipelines (Li
and Fall 2016, Haruna and Fall 2019).

It is a well known fact that the flowability of cemented slurries depends on their rheological
properties (Bannister 1980). Hence, to ensure the better flowability of fresh CPB in the pipelines,
it is important to study the rheological properties (such as yield stress and viscosity) of the CPB
material.

Yield stress and viscosity are the most important rheological parameters of paste flow
through pipelines. Yield stress represents the minimum shear stress that is required to make the
material flowable. In order to flow in the pipeline transport systems, the yield stress of CPB has to
be within its flowable range, which is usually between 200 and 700 Pa (Potvin et al. 2005, Yilmaz
and Fall 2017). The viscosity, on the other hand, is defined as the ratio between the shear stress
and the shear rate at a specific shear rate. The viscosity of CPB is a measure of the resistance of

55
the CPB material towards the flow induced by the shear stress (Yilmaz and Fall 2017, Bian et al.
2019).

Past studies have shown that regardless of the method of delivery in CPB transport systems,
it is evident that CPB material transported from the above ground paste plant to the underground
mine stopes is subjected to several different temperatures from various sources (Figure 3.1) which
might significantly affect its flowability. These sources include: (i) the temperature generated from
the friction between the interior sidewall of the pipeline and the fresh CPB (Fall et al. 2010); (ii)
the heat generated from the rock mass that is surrounding the pipeline pathway and the mine stopes,
especially in deep mines. The thermal interaction between the CPB within the pipelines and the
temperature of these rocks will increase the temperature of the transported CPB (Wang et al. 2016);
(iii) the geographic location of the mine itself, which greatly affects the temperature of the
transported CPB, especially in the extremely hot regions and/or the permafrost regions (Wang et
al. 2016, Alainachi and Fall 2019b); and (iv) the internal heat generated due to the cement
hydration of the fresh CPB, which might change the temperature of the CPB in the field. Indeed,
the hydration process and thermal increase start immediately after the CPB components are mixed
(Aldhafeeri et al. 2016, Alainachi and Fall 2019b). This heat will affect the flowability of the CPB
in the pipelines, especially if the CPB is placed into a stope that is in a deep mine as the transport
distance (in some cases) can be extremely long.

Figure 3.1 CPB transport system and sources of temperature change.

56
Other than the temperature, there are other external and internal factors that might also
affect the rheological characteristics of fresh CPB. However, most of the previous studies (e.g.
Wang et al. 2004, Huynh et al. 2006, Yin et al. 2012, Wang et al. 2018) have studied the rheology
of CPB under the effect of an individual factor (e.g. binder content, type of binder, and pH value).

Some other studies have been conducted recently to understand the coupled effect of time
and temperature on the rheological properties of different mixtures of CPB. For instance, Wu et
al. (2013) developed a mathematical model to predict and assess the changes in the rheological
properties of CPB under the coupled effect of curing time and temperature. The simulation showed
that the temperature and progression of cement hydration (curing time) considerably influence the
rheological behavior of CPB. In line with the work in (Wu et al. 2013), Haiqiang et al. (2016)
investigated the combined effect of time and sub-zero temperatures on the yield stress of CPB.
Their study found that the yield stress of CPB samples studied in sub-zero temperatures is much
lower than that of the sample tested in room temperature. Jiang and Fall (2017), on the other hand,
investigated the effect of the salinity of CPB on the time-temperature coupled behavior of CPB. It
was found that an increase in salinity reduces the yield stress of CPB in subzero temperatures.
Moreover, Haruna and Fall (2019) added different dosages of superplasticizer to a CPB mixture
to determine the temperature-time induced changes in the rheological properties of CPB. Their
results are in agreement with those of previous studies in that an increase in time and temperature
will jointly increase the yield stress and viscosity of CPB. Also, they found that adding
superplasticizer will reduce the yield stress and viscosity of CPB in high temperatures. Cheng et
al. (2020) further analyzed the effect of time and temperature on CPB rheology by establishing a
calculation model implemented with numerical software, and established a 3D simulation model
to determine the effect of the velocity distribution of CPB with time and temperature. They
concluded that the rheological behavior of CPB has a time-temperature equivalent effect. It was
also noted that, along with the decrease in yield stress and viscosity, the flow velocity also
decreases with an increase in time and temperature.

On the other hand, some studies (e.g. Brough and Atkinson 2002, Chang 2003, Abdul-
Hussain and Fall 2012, Kermani et al. 2015a, Kermani et al. 2015b, Cihangir et al. 2018, Jiang et
al. 2019, Kou et al. 2020) have used sodium silicate as a chemical additive in different CPB
mixtures. They used sodium silicate as an activator in different types of CPB binders (such as
Portland cement type I (PCI), blast furnace slag (Slag) and Fly ash (FA)), and also examined the
effect of sodium silicate on the mechanical strength of CPB in the long term. However, no study
to date has investigated the effect of sodium silicate on the rheological properties of CPB at the
very early ages (up to 4 hours). Also, no study has evaluated the combined effects of time,
temperature, binder content, and sodium silicate content on the rheological properties of CPB.
Accordingly, the goal of this study is to experimentally examine the combined effects of
temperature (mixing and curing), type of binder (PCI, Slag, and FA), chemical additive (sodium
silicate), and progression of cement hydration (curing time) on the rheological properties of fresh
CPB mixtures.

57
3.3 Materials and Experimental Program

3.3.1 Materials

3.3.1.1 Tailings
The tailings material used in this study is ground silica tailings (ST) which are a synthetic tailings
material (manufactured by U.S. Silica Co.). The grain size distribution of the ST is similar to the
average grain size distribution of the mine tailings extracted from nine different mines (9MT) in
eastern Canada (Figure 3.2). ST are used in this study because they comprise mostly quartz, which
is the predominant mineral in Canadian hard rock mine tailings. Also, ST have a high percentage
of silica (SiO2; 99.8%) which makes them a chemically inter material. The use of ST will allow
the mineralogical and chemical compositions of the tailings to be controlled and the uncertainties
related to the use of natural tailings to be minimized (Aldhafeeri and Fall 2016, Alainachi and Fall
2019a). The primary physical properties and main chemical elements of the ST are listed in Tables
3.1 and 3.2.

100
9MT
ST
80

60
Passing (%)

52

40

20

0
20
0.01 0.1 1 10 100 1000
Grain diameter (μm)
Figure 3.2 Grain size distribution of silica tailings used in this study vs. average grain size distribution of
tailings from nine Canadian mines.

Table 3.1. Primary physical properties of tailings used in this study

Element Gs D10 (μm) D30 (μm) D50 (μm) D60 (μm)


ST 2.7 1.9 9.0 22.5 31.5

58
Table 3.2. Main chemical elements of tailings used in this study

Element Al wt.% Ca wt.% Si wt.% Fe wt.% Na wt.% Pb wt.% S wt.% K wt.%


ST 0.10 <0.01 99.80 <0.01 <0.01 0.00 0.00 0.00

3.3.1.2 Water and binders


To enhance the mechanical properties of CPB, a sufficient amount of binder is often used in the
mixture. In this study, the binder agent used in the preparation of the CPB mixtures is Portland
cement type I (PCI), Fly Ash (FA), and blast furnace slag (Slag). These binders are the most
frequently used in preparing CPB worldwide (Haiqiang et al. 2016). The properties of these
binders are listed in Table 3.3. PCI is used alone or blended with Slag or FA. The blending ratio
of PCI and Slag or FA is 50/50. This ratio is selected based on: (i) the need of the industrial partner
in this project; (ii) cost savings; (iii) previous studies on performance of this ratio in CPB systems
(e.g., Fall et al. 2010). Tap water is used in this study to mix the binders with the tailings.

Table 3.3. Primary physical and chemical properties of binders used in this study

Binder SSA* (m2/g) R.D** S (wt%) Ca (wt%) Si (wt%) Al (wt%) Mg (wt%) Fe (wt%)

PCI 1.32 3.2 1.5 44.9 8.4 2.4 1.6 1.9


FA 2.20 2.6 1.3 13.3 15.2 9.2 2.6 4.1
Slag 2.10 2.8 12 26.6 18.9 3.9 6.9 0.5

* Specific surface area and ** Relative density

3.3.1.3 Sodium silicate


Soluble sodium silicates are silicate polymers which are an inorganic chemical, and a clear,
colourless, and viscous liquid. Sodium silicate are also known as water soluble glasses which are
generally produced from varying proportions of an alkali metal and SiO2. Sodium silicates have
many applications; for instance, they are used in the geotechnical engineering field in soil grout
and mine tailings backfill, and as an admixture for cement products. In this study, the primary
reason for using sodium silicates is to evaluate their effect on the rheological properties of CPB.
In this regard, a commercial solution of sodium silicate (Type N) was added in the liquid form to
the CPB mixture as the admixture for the binders. The properties of the sodium silicate used in
this study are listed in Table 3.4.

59
Table 3.4. Properties of sodium silicate (National Silicate Ltd.)

Property Value
Na2O% by weight 8.90
SiO2% by weight 28.66
Weight ratio, %SiO2/%Na2O 3.22
Specific gravity @ 20oC 1.39
Solids % 37.56

3.3.2 Sample Preparation


The CPB ingredients (tailings, water, and binder) were first stored in a temperature controlled
chamber (refrigerator or oven, as the case may be) for a minimum of 24 hours or until they reached
the desired temperature (2C, 20C, and 35C). Then, 125 CPB samples were prepared by mixing
a given amount of tailings, binder (4.5 wt%), and water (w/c ratio =7.8) at the testing temperature
(2C, 20C, and 35C). During the mixing of these ingredients, sodium silicate (SS) was added to
each mixture (each sample) with different concentrations (0, 0.1, 0.3, and 0.5% by weight solid
components; solid components = tailings + binder; %SS = (mass of SS/mass of Solids)x100). The
sodium silicate gel used in these mixtures contains 62.44% of water. This water content was
considered when determining the amount of mixing water required for the preparation of the CPB
samples. The samples were mixed for 7 min to obtain a homogenous mixture (Bian et al. 2019).
Then, the prepared CPB mixtures were poured into a 600 ml low-form Griffin beaker and vibrated
to remove the entrapped air. They were subsequently poured into plastic cylindrical containers (10
cm in diameter × 10 cm in height). To prevent the evaporation of water, each sample was covered
with plastic film. They were then placed aside for curing at the same mixing temperatures. To
reflect the range of temperature that the CPB might be subjected to during its transport through the
pipeline system, the CPB samples were prepared, cured and tested at 2oC, 20oC, and 35oC.
Moreover, to reflect the different transportation times in backfill practice from the CPB plant to
the underground stope, the prepared CPB was cured for 0, 0.25, 1, 2, and 4 h before being tested.
Selecting a curing time up to 4 hrs enables to consider the long transportation time that can be
required in deep or ultra depth mines. To ensure that the CPB remained homogeneous, the samples
were agitated for 1 min before each test. This activity also experimentally models the shearing
process of the backfill during transport and prevents the self-weight settlement of the soil particles
during curing (Haiqiang et al. 2016). Moreover, some of the samples were used for microstructural
analyses or monitoring tests (pH and zeta potential, and electrical conductivity (EC)). Table 3.5
summarizes the experimental program and the mix proportion of CPB for each stage of the
program.

60
Table 3.5. Summary of CPB mix proportion and stages of experimental program
Binder %PCI %Slag %FA SS
Sample W/C Mix/Curing Curing Time
Content in the in the in the content
Label ratio Temp. (oC) (h)
(%) binder binder binder (%)
A. Effect of temperature on the time-dependent changes in rheological properties of CPB
SS-CPB- 4.5 100 0 0 7.8 0.3 2 0, 0.25, 1, 2, 4
2oC

SS-CPB-
4.5 100 0 0 7.8 0.3 20 0, 0.25, 1, 2, 4
20oC

SS-CPB-
4.5 100 0 0 7.8 0.3 35 0, 0.25, 1, 2, 4
35oC

B. Effect of binder content and curing temperature on rheological properties of CPB


PCI-SS-
4.5 100 0 0 7.8 0.3 2, 20, 35 2
CPB
PCI/Slag-
4.5 50 50 0 7.8 0.3 2, 20, 35 2
SS-CPB
PCI/FA-
4.5 50 0 50 7.8 0.3 2, 20, 35 2
SS-CPB
C. Effect of sodium silicate content and curing temperature on rheological properties of CPB
0.0-SS-
4.5 100 0 0 7.8 0.0 2, 20, 35 2
CPB
0.1-SS-
4.5 100 0 0 7.8 0.1 2, 20, 35 2
CPB
0.3-SS-
4.5 100 0 0 7.8 0.3 2, 20, 35 2
CPB
0.5-SS-
4.5 100 0 0 7.8 0.5 2, 20, 35 2
CPB
PCI: Portland cement I, FA: Fly ash, SS: Sodium silicate.

3.3.3 Test Methods

3.3.3.1 Viscosity test


A Brookfield digital viscometer (DV-E model) was used to determine the viscosity of the CPB
samples under the conditions provided above. This viscometer consists of a rotating spindle
connected to a calibrated spring. To measure the viscosity, the spindle is immersed into the fluid
to be tested and then rotates in the liquid. The liquid will then resist the rotation which is known
as viscous drag, and determined by the spring deflection. The spring deflection is measured by the
rotary transducer (Brookfield 2014). In this study, an RV5 spindle is used and a specific speed is
applied to ensure that the readings are between 10 and 100% of scale. The rotation speed of the
rotor is changed until obtaining an acceptable viscosity reading, as described in (Xiapeng et al.
2019)). The rotation speeds, 30 revolution per seconds, is used. Before testing, the guard leg was
mounted onto the viscometer to ensure the stability of the sample during testing. In order to prevent
the trapping of air, the spindle was gently inserted into the center of the CPB samples to be tested
in their titled position. Afterwards, the spindle was attached to the spindle coupling nut of the

61
viscometer, which was subsequently turned on and allowed to run for a specific length of time
until reaching a relatively constant reading. After reaching stabilization, the test was carried out
on each CPB sample (according to the aforesaid experimental program). To ensure the
repeatability and accuracy of the obtained results, each test was repeated at least twice. Further
details on the procedure of viscosity measurements of CPB using this viscometer are given in
Brookfield (2014).

3.3.3.2 Yield stress test


Vane shear testing is considered as one of the most efficient, direct and accurate single-point
methods for measuring the yield stress of soil and non-Newtonian fluids (Bauer et al. 2007, Abd
Elaty and Ghazy 2012, Kondraivendhan and Bhattacharjee 2015). In this study, a Wykeham-
Farrance vane shear is used to determine the yield stress of the CPB samples. The vane shear
consists of an electrical motor that rotates the four-bladed vane with a diameter of 2.5 cm and
height of 2.5 cm which is connected to the deflection spring, and a dial gage to measure the
resistance of the sample to the applied torque. Before testing, the vane was slowly inserted into the
middle of the containers (10 cm in diameter × 10 cm in height) of CPB samples to be tested. A
waiting period of 30 seconds was provided to allow the material to stabilize. Afterwards, the motor
was turned on to rotate the vanes at a constant rate of 0.18 rpm until the maximum torque was
reached. The test was carried out in accordance with ASTM D4648/D4648M-16 (2015) and the
readings obtained from the dial gage were subsequently calculated into the corresponding yield
stress values. Each test was repeated at least twice to ensure the repeatability and accuracy of the
obtained results.

3.3.3.3 Microstructural analysis


To understand the effect of the microstructural changes in the CPB on its flowability when
subjected to the different testing conditions, several microstructural analyses were conducted on
selected CPB samples at different curing times. The microstructural analyses were carried out by
using X-ray diffraction (XRD) to determine the changes in the mineralogical composition of the
different CPB samples, and thermal analyses (differential thermogravimetry (DTG) and thermal
gravimetry (TG)) to monitor the changes in the weight, heat flow, and temperature changes in each
CPB sample. Before conducting these analyses, the samples were first dried at 45oC in a vacuum
oven until mass stabilization. The XRD process was conducted with a Scintag XDS2000 x-ray
diffractometer. The thermal analyses, in the other hand, were undertaken with a TGA Q5000
thermal analyzer (TA Instruments). The various (dried) samples (about 20 mg each) were heated
in an inert nitrogen atmosphere at the rate of 10°C per minute up to a temperature of 1000°C.

3.3.3.4 pH and zeta potential measurements


Zeta potential is often determined by measuring the electrophoretic mobility of suspended particles
in samples and applying the Henry equation on the measured values (Clogston and Patri 2011). In
this study, a Zetasizer Nano series was used to determine the difference in the zeta potential of two
CPB samples that were prepared and cured at 2oC, and 35oC. Both samples were prepared by using

62
PCI as the binder, consisted of 0.3% of sodium silicate and cured for 4 h. All specimens of this
test were prepared with 0.1g/L solid to water ratio, and the basic water of these suspensions was
distilled water (Jiang and Fall 2017). As with all the testing, each test was repeated at least three
times to ensure accuracy and repeatability.

On the other hand, measuring the pH of the different samples provides an idea of the
changes in the pore solution volume and alkalinity of the tested samples in line with the changes
of the testing conditions (Ghirian and Fall 2014). The pH changes were determined by again using
CPB samples with PCI (PCI-CPB) as the binder and 0.3% of sodium silicate. They were prepared
and cured at 2oC, 20oC, and 35oC for 0.25 hr (15 min), 1 h, 2 h, and 4 h. One of the samples was
also not cured for any amount of time. The variations in the pH of the tested CPB samples show
the time-temperature effect on the volume of the pore solution of SS-CPB. An Oakton pH Tester
10 Waterproof BNC Pocket pH Tester with a pH range of -1.00 to 15.00 and accuracy of ±0.01
pH was used in this regard.

3.3.3.5 Electrical conductivity monitoring of samples


The changes in the electrical conductivity (EC) reflect the rate of ion movement from the chemical
reactions that take place between cement and water. Monitoring the EC is an effective way to
assess the progression of cement hydration and the related structural changes (Li and Fall 2016).
This part of the experimental program aims to monitor the changes in the EC under the combined
effects of type of binder and sodium silicate content. Accordingly, two sets of EC tests were
conducted. To determine the effect of the binder content, the EC was monitored for three samples
with different binder contents as follows: (i) 100% PCI, (ii) 50% PCI and 50% Slag, and (iii) 50%
PCI and 50% FA. All three samples consisted of 0.3% of sodium silicate, and their initial and
curing temperatures were both 35oC. On the other hand, to determine the effect of the sodium
silicate content, the EC was monitored on two PCI-CPB samples and consisted of: (i) no sodium
silicate content, and (ii) 0.3% sodium silicate. The initial and curing temperatures of both samples
were 35oC. The EC monitoring of all of the samples was carried out while the samples were cured
for 48 h. An ECH2O 5TE sensor that measures the EC in the range of 0-23 dS/m with an accuracy
of ±0.1 was used in this regard.

3.4 Results and Discussion

3.4.1 Effect of temperature on time-dependent changes of rheological properties of CPB


The effect of temperature on the changes in the yield stress and viscosity of the PCI-CPB samples
with 0.3% sodium silicate as a function of time is illustrated in Figures 3.3 and 3.4, respectively.

As expected, for a given temperature, there is a gradual enhancement of the rheological


properties with time (yield stress and viscosity). This is related to the growth of more cement
hydration products, and process of reduction of free water that occurred in the CPB material due
to self-desiccation with time (Bullard et al. 2011, Haiqiang et al. 2016, Bian et al. 2019). This
behavior is consistent with the findings of previous studies (e.g. Roussel 2007, Simon and

63
Grabinsky 2013, Wu et al. 2013, Haiqiang et al. 2016, Panchal et al. 2018, Haruna and Fall 2019,
Xiapeng et al. 2019) which indicated that the rheological properties of fresh CPB or other
cementitious materials (e.g., fresh concrete) can significant increase in the initial stage of cement
hydration (30 minutes or less) and beyond, as a result of the progress of cement hydration. It is
also evident from results of Figures 3.3 and 3.4 that, for a given curing time, the initial and curing
temperatures have a direct relationship with both the yield stress and viscosity. Samples which are
subjected to a temperature of 2oC have the lowest yield stress and viscosity, while those subjected
to a temperature of 35oC have the highest. This phenomenon is attributed to the acceleration of the
cement hydration process under high temperatures (Aldhafeeri et al. 2016, Cui and Fall 2016).
This acceleration of the cement hydration due to high temperatures will result in the formation of
more cement hydration products (such as calcium silicate hydrate (C-S-H), ettringite, and calcium
hydroxide (CH)) as well as increase the self-desiccation of the backfill material (Fall et al., 2010).
This is beneficial for yield stress increases of the CPB because more of these products will result
in increased cohesion between the tailings particles (the free CPB particles are progressively
bonded to the hydration products to form a 3D network structure, or a gel structure, which leads
to increased cohesion between particles) as well as reduce the size of the pores between the CPB
particles, which increases the yield stress (Wang et al. 2016, Alainachi and Fall 2019b).
Furthermore, the consumption of capillary water by the cement hydration (self-desiccation) will
lead to the reduction of the space between the CPB particles and increased particle inter-friction
(Wang et al., 2016). Accordingly, the interaction and the cementitious bond between the particles
rise, which in turn, leads to higher yield stress values. Moreover, this temperature-induced
formation of more hydration products and higher consumption of water molecules in the capillary
pores mean that the solid volume fraction of the backfill material will increase and the CPB will
become more viscous (Xiapeng et al. 2019), which is obviously associated with an increase of the
viscosity of the CPB. This rationale is supported by the results of the thermal analysis in this study,
and Figure 3.5 provides the TG/DTG diagrams for two PCI-CPB samples (with 0.3% of sodium
silicate) prepared and cured (up to 4 h) at 20oC and 35oC. These diagrams confirm that more
hydration products are produced within the CPB at higher initial and curing temperatures. The first
two peaks (sudden change in weight) show a higher weight loss within the CPB prepared and cured
at 35oC. The first peak is due to the destruction of the C-S-H, ettringite, and gypsum (e.g. Taylor
1990, Fall et al. 2010), while the second peak is associated with the degradation of CH (e.g.
Aldhafeeri et al. 2016). The third peak represents the decomposition of the calcite in the cement
(Taylor 1990, Mehta and Monteiro 2001), and is lower for the CPB prepared at 35oC compared to
the CPB prepared at 20oC.

64
700
2°C 20°C 35°C
600

500
Yield Stress (Pa)
400

300

200

100

0
0 1 2 3 4
Time (hrs)

Figure 3.3 Temperature vs. yield stress of PCI-CPB samples with 0.3% sodium silicate.

9
2°C 20°C 35°C
8

6
Viscocity (Pa.S)

0
0 1 2 3 4
Time (hrs)

Figure 3.4 Temperature vs. viscosity of PCI-CPB samples with 0.3% sodium silicate.

65
100 0.16

0.14
95
0.12

Deriv. Weight (%/°C)


Weight (%) 90 0.1

0.08
85
0.06

80 0.04

0.02
75
0
20°C 35°C
70 -0.02
0 200 400 600 800 1000
Temperature (°C)

Figure 3.5 DT/DTG diagrams for PCI-CPB samples with 0.3% sodium silicate cured for 4 h vs. different
curing temperatures.
The increase in the viscosity of the backfill material at higher temperatures is consistent
with the results of the zeta potential measurements of the PCI-CPB samples (0.3% of sodium
silicate) cured up to 4 h at 2oC and 35oC, as shown in Figure 3.6. The CPB samples prepared and
cured at a temperature of 2oC has the highest absolute zeta potential value compared to those
prepared and cured at a temperature of 35oC. This higher zeta potential value at lower curing
temperatures suggest that stronger repulsive forces are found between the particles in CPB
subjected to lower temperatures, since the zeta potential is the electrostatic surface charge or
magnitude of the electric repulsion or attraction between colloidal particles (Elakneswaran et al.
2009, Jiang and Fall 2017). In other words, the particle dispersion within the CPB is higher in low
temperatures, and thus results in better flowability (Huynh et al. 2006, Haruna and Fall 2019).
These zeta potential measurement results and the resultant conclusion made are in good agreement
with the results of the pH measurements that were conducted on the CPB samples at different
temperatures and shown in Figure 3.7. The pH values are higher for the CPB samples that are
prepared and cured at higher temperatures compared to those at lower temperatures. This is related
to the fact that higher temperatures increase cement hydration, and thereby leads to higher
electrolyte concentration and reduced water content in the CPB system. It is well acknowledged
that increasing the electrolyte concentration or decreasing the water content reduces the thickness
of the diffuse double layers (DDLs) (Mojid 2011), and results in an increase in the pH of the
solution which compresses the diffuse layer due to the well-known concentration effect (Nagele
1986).

66
2°C 35°C
0

-10
Zeta Potential (mv)

-20

-30

-40

-50

Figure 3.6 Zeta potential of PCI-CPB samples with 0.3% of sodium silicate prepared and cured at
different temperatures for 4 h.
12.8
2°C 20°C 35°C
12.7
12.6
12.5
12.4
pH

12.3
12.2
12.1
12.0
11.9
11.8
0 1 2 3 4
Time (hrs)

Figure 3.7 Changes in pH of PCI-CPB samples with 0.3% sodium silicate prepared and cured at different
temperatures for 4 h.

67
3.4.2 Combined effect of type of binder and curing temperature on rheological properties
of CPB
Pozzolanic materials, such as Slag and FA, are often used in paste backfill mixes as a partial
replacement for Portland cement to reduce the cost due to cement consumption, while maintaining
the desired strength (Cui 2017). There is also consensus in the literature that pozzolans can be
effectively activated by adding alkali activators, such as sodium silicate (Kermani et al. 2015a,
Azadi et al. 2017). During the early hydration process of cement-slag blends, the hydration of the
clinker releases alkali and increases the temperature. This provides the energy to activate alkali-
hydroxide attacks on the Slag particles, which activates the Slag and accelerates its hydration
process (Roy and Idorn 1982, Kolani et al. 2012). Several research studies (e.g. Qing-Hua et al.
1992, Brough and Atkinson 2002) have suggested that CH reacts with sodium silicate and leads to
the consumption of its silica components which leaves the sodium oxide (Na2O) relatively free
around the Slag particles, thus activating/accelerating the Slag hydration and producing higher
amounts of C-S-H. In other words, the reaction between CH and sodium silicate increases the
concentration of alkali ions around the Slag particles and produces a high pH (e.g. Qing-Hua et al.
1992), which is indicative of the start of the Slag hydration and strength build-up of the mixture
(e.g. Qing-Hua et al. 1992, Brough and Atkinson 2002). Cement-FA blends, on the other hand,
have an initial reaction that is similar to the cement-Slag blends. However, the exothermic process
between sodium silicate and the FA activates the alkali reaction of the FA, during which the FA
structure is destroyed. This results in the progressive accumulation of reaction products that are
produced during the destruction of the FA structure. Subsequently, condensation occurs on the
surface of the FA particles, which contributes to creating a cementitious material with high
mechanical strength (Palomo et al. 1999). In other words, the presence of sodium silicate increases
the alkali reaction rate, and consequently accelerates the development of the mechanical strength
of the mixture with the blended binders (Palomo et al. 1999, Cihangir et al. 2012). Moreover, it is
widely accepted that compounds of alkali metals (sodium and potassium) stimulate the hydration
of the key phases of Portland cement, at least in the early stages of the hydration process, and
accelerate the setting of cement paste (e.g. Jawed and Skalny 1978, Taylor 1990). For example,
Brykov et al. (2002) indicated that the hydration of Portland cement clinker minerals happens at a
higher rate in the presence of sodium silicates because of a reduction in the initial pH of the pore
solutions of cement pastes and intensive production of CH.

Previous studies (e.g. Simon and Grabinsky 2013, Haiqiang et al. 2016, Jiang and Fall
2017) have observed that addition of Slag or FA to the binder of CPB will significantly affect the
rheological properties of CPB. However, no studies have investigated the rheological properties
of CPB samples with Slag or FA activated with sodium silicate. This knowledge gap is addressed
in the following section.

In this part of the study, the effect of type of binder on the rheological properties of CPB
samples with sodium silicate is assessed by determining the yield stress and viscosity of samples
cured for 2 h with different types of binders: (i) 100% PCI, (ii) 50% PCI and 50% Slag, and (iii)

68
50% PCI and 50% FA. All of the samples contain 0.3% of sodium silicate. The effects of the type
of binder on the yield stress and viscosity with respect to the mixing and curing temperatures are
shown in Figures 3.8 and 3.9, respectively.

600

500

400
Yield Stress (Pa)

300

200

100

0
2 20 35
Temp (°C)
PCI PCI/Slag PCI/FA
Figure 3.8 Effect of temperature and binder type on yield stress of CPB samples with 0.3% sodium
silicate cured for 2 hrs.

6
Viscocity (Pa.S)

0
2 20 35
Temp. (°C)
PCI PCI/Slag PCI/FA
Figure 3.9 Effect of temperature and binder type on viscosity of CPB samples with 0.3% sodium silicate
cured for 2 hrs.

69
As shown in Figures 3.8 and 3.9, high temperatures significantly increase the yield stress
and viscosity of CPB with sodium silicate, regardless of the type of binder. The key process
responsible for this behaviour is the temperature-induced increase of the binder hydration products,
as discussed earlier. However, it should be noted that increasing the temperature will also enhance
the activation of the clinker, FA and Slag particles, which is obviously associated with the
production of more binder hydration products in comparison to the samples without sodium
silicate. The latter is an additional factor that contributes to explaining the significant increase in
the yield stress and viscosity of the CPB samples with sodium silicate and Slag or FA with
temperature (Figures 3.8 and 3.9), as will be discussed later.

Moreover, it can be observed in Figures 8 and 9 that the partial replacement of PCI with
Slag or FA reduces the yield stress and viscosity of the sodium silicate rich CPB, regardless of the
temperature. It is evident that at all testing temperatures, the yield stress and viscosity of CPB
made with PCI/Slag as the binder (and containing sodium silicate) are lower than those of the PCI-
CPB samples (and containing sodium silicate). Furthermore, CPB samples made with PCI/FA as
the binder (and containing sodium silicate) have the lowest yield stress and viscosity.

The reduction of the yield stress and viscosity when FA and PCI are used as the binder can
be explained by the combined effect of three processes: (i) the spherical shape of the FA particles
reduces the friction forces between the particles (Siddique 2004, Park et al. 2005); (ii) the partial
replacement of PCI with FA increases the packing density of the CPB particles because the FA
particles are much finer than the PCI particles, which will reduce the volume of pore-water
required to fill the voids between particles. This will allow the water to act as a lubricant for the
particles (Lee et al. 2003, Kwan and Chen 2013), and reduce the stickiness and consequently
reduce the yield stress and viscosity; and (iii) the internal heat within the cemented mixture due to
the cement hydration process will be significantly reduced by using FA (Le et al. 2019). Hence,
the hydration process will slow down and thus the yield stress and viscosity will decrease at the
very early ages of hydration. On the other hand, the higher values of yield stress and viscosity of
CPB when Slag and PCI are used as the binder (compared with PCI/FA as the binder) can be
attributed to the fact that the combination of Slag and PCI increases the volume of cement
hydration products produced as a result of the exothermic reaction between Slag and CH that is
facilitated from the early hydration of cement (Fall et al. 2010, Tariq and Yanful 2013). However,
it is evident that the yield stress and viscosity of CPB made with PCI/Slag are lower than those of
the PCI-CPB samples. This phenomenon is due to the initial setting time, which significantly
affects the rheology of cement paste as the initial setting time is when the cement paste gradually
loses its plasticity (Sarda et al. 2001), and is associated with the cement hydration process (Zheng
et al. 2015), which occurs at longer curing times for PCI/Slag systems (Brough and Atkinson
2002). Moreover, using Slag will reduce the heat released during the hydration process (Kolani et
al. 2012), and thus slow down the hydration process and reduce the volume of cement hydration
products at the very early ages.

70
This assertion is experimentally validated by the results of the thermal analyses (TG/DTG
and XRD) performed on the CPB samples with different types of binders and 0.3% of sodium
silicate, and cured for 2 hours at 35oC. The TG/DTG graphs in Figure 3.10 confirm that the sodium-
silicate rich CPB sample prepared with only PCI as the binder has a larger volume of hydration
products, while there are less (e.g., C-S-H, ettringite) within the samples with PCI/Slag compared
to the sample with only PCI, but higher than in the sample with PCI/FA. The graphs in Figure 10
show that from the first weight loss and peak in the TG and DTG curves, respectively, which
occurred at about 100oC, the volume of cement hydration products, such C-S-H and ettringite, is
more in the PCI-CPB samples, followed by PCI/Slag and finally PCI/FA. The lower peak and
weight loss at 450C of the PCI/Slag sample compared to the PCI/FA sample is the result of the
consumption of CH during the activation of Slag, which obviously results in the generation of
additional C-S-H.
100 0.16
PCI PCI/Slag PCI/FA
0.14
95
0.12

Deriv. Weight (%/°C)


90 0.1
Weight (%)

0.08
85
0.06

80 0.04

0.02
75
0

70 -0.02
0 200 400 600 800 1000
Temperature (°C)

Figure 3.10 DT/DTG diagrams for CPB samples with 0.3% of sodium silicate and cured for 2 h with
different binder type at 35oC.
Moreover, the plotted information in the TG/DTG graphs is consistent with the results of
the XRD results (Figure 3.11), in which the type and relative volume of cement hydration products
formed within these CPB samples are illustrated. It is apparent that some of these hydration
products show a higher intensity for the CPB sample with only PCI compared to the samples with
PCI/Slag, while the intensity is the lowest for the CPB samples with PCI/FA. In other words, more
hydration products formed in the PCI-CPB samples than the PCI/Slag samples, with the lowest
volume in the samples with PCI/FA. For example, the intensity of portlandite at 18o-2θ is 621 CPS
and 196 CPS for the CPB samples with PCI and PCI/FA, respectively. Similarly, the intensity of

71
ettringite at 41o-2θ is 117 CPS, 99 CPS, and 96 CPS for the PCI-CPB, PCI/Slag and PCI/FA
samples, respectively.

1200
PCI PCI/Slag PCI/FA

1000

Portlandit
800
Intensity (CPS)

Portlandite
600 Ettringite
Ettringite Ettringite
400
Gypsum

200

0
0 20 40 60 80
Degrees 2-theta-Cu Ka Radiation
Figure 3.11 XRD intensities for CPB samples with 0.3% of sodium silicate cured for 2 h with different
binder type at 35oC.
Furthermore, the EC monitoring results in Figure 3.12 support this result. It can be
observed from Figure 3.12 that the CPB sample with only PCI has the highest EC over a curing
time of 4 h. This indicates that the concentration and mobility of ions are the highest when only
PCI is used, which would also suggest that the hydration reaction within the PCI-CPB samples
consumes more free pore-water, so that the CPB becomes more viscous (Figure 3.9). Moreover,
the EC of the PCI-CPB peaks earlier than that of the PCI/FA or PCI/Slag samples. In other words,
the cement hydration is more rapid in the PCI-CPB samples, which is consistent with the results
presented in Figures 3.8 and 3.9. On the other hand, the PCI/FA samples show a higher EC as
opposed to the PCI/Slag samples for the first 1.6 hours (100 minutes), and then the EC of the
PCI/Slag samples increases and then exceeds the EC of the PCI/FA samples. The EC continues to
be high for the rest of the monitored period during which the PCI/FA samples show a gradual
reduction in EC. This observation shows that the partial use of FA increases the mobility and
concentration of ions at the beginning (compared with the partial use of Slag) and then decreases
after 100 minutes.

The results obtained here suggest that the use of PCI/FA as the binder in CPB with sodium
silicate will impart better flowability in the CPB material even in high temperatures.

72
6.0
PCI PCI/Slag PCI/FA
5.5
5.0
4.5
4.0
EC (mS/cm)

3.5
3.0
2.5
2.0
1.5
1.0
0.5
0.0
0 1 2 3 4
Time (hrs)

Figure 3.12 Development of EC in CPB samples with 0.3% of sodium silicate and different binder type at
35oC.

3.4.3 Combined effect of amount of sodium silicate and curing temperature


The effect of the amount of sodium silicate on the rheological properties of CPB was established
by testing the yield stress and viscosity of the PCI-CPB samples cured for 2 hours with different
amounts of sodium silicate (0.0%, 0.1%, 0.3%, and 0.5%). The effect of the binder content on the
yield stress and viscosity of the PCI-CPB samples with respect to the mixing and curing
temperatures are shown in Figures 3.13 and 3.14, respectively.

It is evident in Figures 3.13 and 3.14 that, for a given temperature, an increase in the sodium
silicate content increases the yield stress and viscosity of the PCI-CPB. The yield stress and
viscosity are found to be directly correlated with the sodium silicate content. Moreover, it can be
also observed that PCI-CPB with 0.3% and 0.5% of sodium silicate show a higher relative increase
in yield stress and viscosity at higher temperatures (20oC and 35oC) compared to a lower
temperature (2oC). This phenomenon is explained as follows. On the one hand, an increase in the
initial and curing temperatures of PCI-CPB samples with sodium silicate enhances the sodium
silicate as an alkali activator (Živica 2007). The enhancement in activation energy of sodium
silicate at high temperature is associated with the increase of absorption ability due to the stress-
induced sodium ion migration (Souquet et al. 2010). An increase (albeit not an excessive one) in
sodium content leads to the activation of more Slag and FA particles as well as the clinker phases
(Taylor 1990, Abdul-Hussain and Fall 2012). Consequently, these two factors lead to the growth
of more cement hydration products, thereby changing the rheological properties of the CPB. On
the other hand, high dosages of an alkali activator also significantly decrease the setting time of

73
CPB (Azadi et al. 2017), and subsequently increase the yield stress and viscosity of CPB (Warner
2004, Azadi et al. 2017). This is also in agreement with the findings in (Kermani et al. 2015b),
who examined the initial setting time of CPB with different amounts of sodium silicate, and found
that the initial setting time of CPB without any sodium silicate is approximately 4.5 hours (270
minutes) but about 3.8 hours (230 minutes) and 3.5 hours (210 minutes) for CPB samples with
0.3% and 0.5% of sodium silicate, respectively. This argument which is related to the formation
of more cement hydration products with higher amounts of sodium silicate is consistent with the
results of the thermal analyses (TG/DTG) carried out on the CPB samples with different amounts
of sodium silicate (0.0%, 0.3% and 0.5%), and cured for 2 hours at 35oC in this study, as shown
in Figure 3.15. For example, it is evident from the first peak in the DTG curves, which occurred at
about 100oC, that the volume of cement hydration products (mainly C-S-H, ettringite and gypsum)
is higher in the samples with 0.5% of sodium silicate, followed by the CPB with 0.3% of sodium
silicate and finally the samples made without sodium silicate. Moreover, the second peak which
occurs at about 450oC shows an almost similar volume of CH produced within the CPB samples
with 0.3% and 0.5% of sodium silicate, and less within the CPB samples with no sodium silicate.

600

500

400
Yield Stress (Pa)

300

200

100

0
2 20 35
Temp (°C)
0.0% 0.1% 0.3% 0.5%

Figure 3.13 Effect of temperature and sodium silicate content on yield stress of PCI-CPB cured for 2 h.

74
8

Viscocity (Pa.S) 5

0
2 20 35
Temp. (°C)
0.0% 0.1% 0.3% 0.5%

Figure 3.14 Effect of temperature and sodium silicate content on viscosity of PCI-CPB cured for 2 h.

100 0.16
0.0% 0.3% 0.5%
0.14
95
0.12

Deriv. Weight (%/°C)


90 0.1
Weight (%)

0.08
85
0.06

80 0.04

0.02
75
0

70 -0.02
0 200 400 600 800 1000
Temperature (°C)

Figure 3.15 DT/DTG of PCI-CPB-SS with different sodium silicate contents prepared and cured for 4 h at
35oC.
Furthermore, the EC monitoring results in Figure 16 support this finding. In the figure, the
initial EC values of CPB with 0.5% sodium silicate are higher than those of CPB with 0.3% sodium
silicate and no sodium silicate mainly due to their initial differences in the sodium silicate,
particularly in the volume of free Na+ ions. Moreover, it is also obvious from these curves that the
75
plotted EC values of the CPB with 0.5% of sodium silicate peak first (at 1.75 hrs), followed by the
CPB with 0.3% of sodium silicate (2 hrs), and finally, the CPB without sodium silicate (2.3 hrs).
This suggests that the rate of binder hydration increases as the sodium silicate content is increased.
It can also be observed from Figure 16 that the CPB with 0.5% of sodium silicate shows the highest
rate of increase in EC and a significantly higher EC persists over the 4 hours of curing, which
indicates that the concentration and mobility of ions are higher with 0.5% of sodium silicate. This
also means that the PCI-CPB consumes more free pore-water, so that the CPB become more
viscous (Figure 3.14). On the other hand, the samples with 0.3% of sodium silicate show a lower
rate of increase in the EC compared to those with 0.5% sodium silicate, but higher rate of increase
than the CPB samples with no sodium silicate. This observation further validates that the addition
of sodium silicate to CPB mixtures accelerates the mobility and increases the concentration of ions
within the CPB, thus resulting in more rapid binder hydration and consumption of more capillary
water, and thereby increasing the yield stress (Figure 3.13) and viscosity (Figure 3.14) of the
backfill material.
7.5
0.0% 0.3% 0.5%
7.0
6.5
6.0
5.5
EC (mS/cm)

5.0
4.5
4.0
3.5
3.0
2.5
2.0
0 1 2 3 4
Time (hrs)

Figure 3.16 Time-dependent changes of electric conductivity of PCI-CPB-SS with different sodium
silicate contents cured at 35ºC.

3.5 Summary and Conclusion


This study mainly examines the combined effects of temperature, type of binder and sodium
silicate content on the time-dependent changes in the rheological properties (yield stress and
viscosity) of CPB material. The primary conclusions based on the results of this research are
summarized as follows:

76
1. An increase in the initial and curing temperatures of CPB with sodium silicate increases
cement hydration reactions, which consequently increases cement hydration and reduce the
water content in the CPB, and results in increased yield stress and viscosity of the CPB
material. The temperature-dependent increase holds, regardless whether the CPB sample is
made with only PCI, PCI/Slag, or PCI/FA.
2. Adding sodium silicate to the CPB significantly reduces its flowability by increasing its yield
stress and viscosity, regardless of the exposed temperature. This reduced flowability is
attributed to the activation to the pozzolanic compounds (Slag particles, FA) and acceleration
of the hydration of the clinker and binder, which result in stronger and more viscous CPB.
However, the magnitude of this effect of the sodium silicate on the rheological properties of
CPB is a function of the sodium silicate content and the temperature.
3. The yield stress and viscosity increments are directly proportional to the increase in the dosage
of sodium silicate regardless of the change in temperature due to more complete activation of
the studied pozzolans and more rapid cement hydration. However, the CPB samples with
sodium silicate show a higher relative increase in yield stress and viscosity at higher
temperatures (20oC and 35oC) compared to a lower temperature (2oC). The behaviour
described above can be explained by the: (i) temperature-induced enhancement of sodium
silicate as an alkali activator, and (ii) acceleration of the setting time of CPB due to high
dosages of sodium silicate.
4. The partial use of Slag or FA as a binder with PCI reduces the yield stress and viscosity
(enhances the flowability) of the CPB samples with sodium silicate. However, the partial use
of FA enhances the flowability more compared with the partial use of Slag, as CPB samples
with PCI/FA show a lower yield stress and viscosity than those with PCI/Slag.

The findings of this investigation demonstrate the significance of temperature and its
effects on the time-dependent rheological behavior or flow ability of CPB with various contents
of sodium silicate and provide significant technical information for the designing of cost-effective
transport systems of CPB with sodium silicate as well as for the optimal assessment of the
flowability of CPB under various mine thermal conditions. Further studies are required to gain
insight into the flow ability of CPB with sodium silicate in sub-zero environments as increasingly
number of mining activities are being conducted in permafrost or cold regions

3.6 Data Availability Statement


All data, models, and code generated or used during the study appear in the submitted article.

3.7 Acknowledgements
The authors of this study would like to thank Sada Haruna and Jean-Claude Celestin for their help
during the experimental program. They also thank the Natural Sciences and Engineering Research
Council (NSERC) of Canada for their financial support, and Lafarge Canada for providing the
pozzolans (slag and fly ash).

77
3.8 References

Abd Elaty, M. A. and Ghazy, M. F. (2012) 'Flow properties of fresh concrete by using modified
geotechnical Vane shear test', HBRC Journal, 8(3), 159-169.

Abdul-Hussain, N. and Fall, M. (2012) 'Thermo-hydro-mechanical behaviour of sodium silicate-


cemented paste tailings in column experiments', Tunnelling and Underground Space
Technology, 29, 85-93.

Ahmari, S., Parameswaran, K. and Zhang, L. (2015) 'Alkali Activation of Copper Mine Tailings
and Low-Calcium Flash-Furnace Copper Smelter Slag', Journal of Materials in Civil
Engineering, 27(6).

Alainachi, I. and Fall, M. (2019a) 'Seismic induced liquefaction of cemented paste backfill: effect
of mixing water', in Proceedings of 7th International Conference on Earthquake
Geotechnical Engineering, Rome, Italy, 17-20 June, 2019,

Alainachi, I. and Fall, M. (2019b) 'Shaking table testing of liquefaction potential of cemented paste
backfill: Effects of backfill temperature', in Geo St-John's 2019 – the 72th Canadian
Geotechnical Conference (CGC), Sept. 29- Oct 2, 2019, St. John's, NL, Canada, September
29 - October 2, 2019,

Alakangas, L., Dagli, D. and Knutsson, S. (2013) Literature review on potential geochemical and
geotechnical effects of adopting paste technology under cold climate conditions, Luleå
tekniska universitet.

Aldhafeeri, Z. and Fall, M. (2016) 'Time and damage induced changes in the chemical reactivity
of cemented paste backfill', Journal of Environmental Chemical Engineering, 4(4), 4038-
4049.

Aldhafeeri, Z., Fall, M., Pokharel, M. and Pouramini, Z. (2016) 'Temperature dependence of the
reactivity of cemented paste backfill', Applied Geochemistry, 72, 10-19.

ASTM D4648/D4648M-16 (2015) 'Standard test method for laboratory miniature vane shear test
for saturated fine-grained clayey soil',

78
Aydın, S. and Kızıltepe, C. Ç. (2019) 'Valorization of Boron Mine Tailings in Alkali-Activated
Mortars', Journal of Materials in Civil Engineering, 31(10), 04019224.

Azadi, M. R., Taghichian, A. and Taheri, A. (2017) 'Optimization of cement-based grouts using
chemical additives', Journal of Rock Mechanics and Geotechnical Engineering, 9(4), 623-
637.

Bannister, C. E. (1980) 'Rheological evaluation of cement slurries: Methods and models', in SPE
Annual Technical Conference and Exhibition, Society of Petroleum Engineers.,

Bauer, E., de Sousa, J. G. G., Guimarães, E. A. and Silva, F. G. S. (2007) 'Study of the laboratory
Vane test on mortars', Building and Environment, 42(1), 86-92.

Bian, J., Fall, M. and Haruna, S. (2019) 'Sulphate induced changes in rheological properties of
fibre-reinforced cemented paste backfill', Magazine of Concrete Research, 1-35.

Brookfield (2014) 'Brookfield Dial Reading Viscometer with Electronic Drive: Model DV‐E
Operating Instructions.', [online], available: https://pim-
resources.coleparmer.com/instruction-manual/98945-xx.pdf].

Brough, A. R. and Atkinson, A. (2002) 'Sodium silicate-based, alkali-activated slag mortars: Part
I. Strength, hydration and microstructure', Cement and Concrete Research, 32(6)(6), 865-
879.

Brykov, A., Danilov, V., Korneev, V. and Larichkov, A. (2002) 'Effect of hydrated sodium
silicates on cement paste hardening', Russian journal of applied chemistry, 75(10), 1577-
1579.

Bullard, J. W., Jennings, H. M., Livingston, R. A., Nonat, A., Scherer, G. W., Schweitzer, J. S.,
Scrivener, K. L. and Thomas, J. J. (2011) 'Mechanisms of cement hydration', Cement and
Concrete Research, 41(12), 1208-1223.

Cao, S., Yilmaz, E. and Song, W. (2018) 'Dynamic response of cement-tailings matrix composites
under SHPB compression load', Construction and Building Materials, 186, 892-903.

Chang, J. J. (2003) 'A study on the setting characteristics of sodium silicate-activated slag pastes',
Cement and Concrete Research, 33(7), 1005-1011.

79
Chen, X., Shi, X., Zhou, J., Du, X., Chen, Q. and Qiu, X. (2019) 'Effect of overflow tailings
properties on cemented paste backfill', J Environ Manage, 235, 133-144.

Cheng, H., Wu, S., Li, H. and Zhang, X. (2020) 'Influence of time and temperature on rheology
and flow performance of cemented paste backfill', Construction and Building Materials,
231.

Cihangir, F., Ercikdi, B., Kesimal, A., Ocak, S. and Akyol, Y. (2018) 'Effect of sodium-silicate
activated slag at different silicate modulus on the strength and microstructural properties
of full and coarse sulphidic tailings paste backfill', Construction and Building Materials,
185, 555-566.

Cihangir, F., Ercikdi, B., Kesimal, A., Turan, A. and Deveci, H. (2012) 'Utilisation of alkali-
activated blast furnace slag in paste backfill of high-sulphide mill tailings: Effect of binder
type and dosage', Minerals Engineering, 30, 33-43.

Clogston, J. D. and Patri, A. K. (2011) 'Zeta potential measurement' in Characterization of


nanoparticles intended for drug delivery, Humana Press, 63-70.

Cui, L. (2017) Multiphysics Modeling and Simulation of the Behavior of Cemented Tailings
Backfill, Doctoral dissertation: University of Ottawa.

Cui, L. and Fall, M. (2016) 'Mechanical and thermal properties of cemented tailings materials at
early ages: Influence of initial temperature, curing stress and drainage conditions',
Construction and Building Materials, 125, 553-563.

Dimitrova, R. S. and Yanful, E. K. (2011) 'Undrained Strength of Deposited Mine Tailings Beds:
Effect of Water Content, Effective Stress and Time of Consolidation', Geotechnical and
Geological Engineering, 29(5), 935-951.

Dungan, P. and Murphy, S. (2012) Mining: Dynamic and Dependable for Ontario’s Future. ,
Ontario Mining Association.

Elakneswaran, Y., Nawa, T. and Kurumisawa, K. (2009) 'Zeta potential study of paste blends with
slag', Cement and Concrete Composites, 31(1), 72-76.

Ercikdi, B., Cihangir, F., Kesimal, A., Deveci, H. and Alp, I. (2009a) 'Utilization of industrial
waste products as pozzolanic material in cemented paste backfill of high sulphide mill
tailings', J Hazard Mater, 168(2-3), 848-56.

80
Ercikdi, B., Kesimal, A., Cihangir, F., Deveci, H. and Alp, İ. (2009b) 'Cemented paste backfill of
sulphide-rich tailings: Importance of binder type and dosage', Cement and Concrete
Composites, 31(4), 268-274.

Esmaeili, J., Aslani, H. and Onuaguluchi, O. (2020) 'Reuse Potentials of Copper Mine Tailings in
Mortar and Concrete Composites', Journal of Materials in Civil Engineering, 32(5),
04020084.

Fall, M. and Benzaazoua, M. (2005) 'Modeling the effect of sulphate on strength development of
paste backfill and binder mixture optimization', Cement and Concrete Research, 35(2),
301-314.

Fall, M., Benzaazoua, M. and Ouellet, S. (2005) 'Experimental characterization of the influence of
tailings fineness and density on the quality of cemented paste backfill', Minerals
Engineering, 18(1), 41-44.

Fall, M., Célestin, J. C., Pokharel, M. and Touré, M. (2010) 'A contribution to understanding the
effects of curing temperature on the mechanical properties of mine cemented tailings
backfill', Engineering Geology, 114(3-4), 397-413.

Farkish, A. and Fall, M. (2013) 'Rapid dewatering of oil sand mature fine tailings using super
absorbent polymer (SAP)', Minerals Engineering, 50-51, 38-47.

Ghirian, A. and Fall, M. (2014) 'Coupled thermo-hydro-mechanical–chemical behaviour of


cemented paste backfill in column experiments', Engineering Geology, 170, 11-23.

Haiqiang, J., Fall, M. and Cui, L. (2016) 'Yield stress of cemented paste backfill in sub-zero
environments: Experimental results', Minerals Engineering, 92, 141-150.

Haruna, S. and Fall, M. (2019) 'Time- and temperature-dependent rheological properties of


cemented paste backfill that contains superplasticizer', Powder Technology.

Huynh, L., Beattie, D. A., Fornasiero, D. and Ralston, J. (2006) 'Effect of polyphosphate and
naphthalene sulfonate formaldehyde condensate on the rheological properties of dewatered
tailings and cemented paste backfill', Minerals Engineering, 19(1), 28-36.

Jawed, I. and Skalny, J. (1978) 'Alkalies in cement: A review: II. Effects of alkalies on hydration
and performance of Portland cement', Cement and Concrete Research, 8(1)(1), 37-51.

81
Jiang, H. and Fall, M. (2017) 'Yield stress and strength of saline cemented tailings in sub-zero
environments: Portland cement paste backfill', International Journal of Mineral
Processing, 160, 68-75.

Jiang, H., Qi, Z., Yilmaz, E., Han, J., Qiu, J. and Dong, C. (2019) 'Effectiveness of alkali-activated
slag as alternative binder on workability and early age compressive strength of cemented
paste backfills', Construction and Building Materials, 218, 689-700.

Kermani, M., Hassani, F. P., Aflaki, E., Benzaazoua, M. and Nokken, M. (2015a) 'Evaluation of
the effect of sodium silicate addition to mine backfill, Gelfill − Part 1', Journal of Rock
Mechanics and Geotechnical Engineering, 7(3), 266-272.

Kermani, M., Hassani, F. P., Aflaki, E., Benzaazoua, M. and Nokken, M. (2015b) 'Evaluation of
the effect of sodium silicate addition to mine backfill, Gelfill – Part 2: Effects of mixing
time and curing temperature', Journal of Rock Mechanics and Geotechnical Engineering,
7(6), 668-673.

Kesimal, A., Yilmaz, E., Ercikdi, B., Alp, I. and Deveci, H. (2005) 'Effect of properties of tailings
and binder on the short-and long-term strength and stability of cemented paste backfill',
Materials Letters, 59(28), 3703-3709.

Kolani, B., Buffo-Lacarrière, L., Sellier, A., Escadeillas, G., Boutillon, L. and Linger, L. (2012)
'Hydration of slag-blended cements', Cement and Concrete Composites, 34(9), 1009-1018.

Kondraivendhan, B. and Bhattacharjee, B. (2015) 'Flow behavior and strength for fly ash blended
cement paste and mortar', International Journal of Sustainable Built Environment, 4(2),
270-277.

Koohestani, B., Khodadadi Darban, A. and Mokhtari, P. (2018) 'A comparison between the
influence of superplasticizer and organosilanes on different properties of cemented paste
backfill', Construction and Building Materials, 173, 180-188.

Kou, Y., Jiang, H., Ren, L., Yilmaz, E. and Li, Y. (2020) 'Rheological Properties of Cemented
Paste Backfill with Alkali-Activated Slag', Minerals, 10(3).

Kwan, A. K. H. and Chen, J. J. (2013) 'Adding fly ash microsphere to improve packing density,
flowability and strength of cement paste', Powder Technology, 234, 19-25.

82
Le, T. H. M., Park, D.-W., Park, J.-Y. and Phan, T. M. (2019) 'Evaluation of the Effect of Fly Ash
and Slag on the Properties of Cement Asphalt Mortar', Advances in Materials Science and
Engineering, 2019, 1-10.

Lee, S. H., Kim, H. J., Sakai, E. and Daimon, M. (2003) 'Effect of particle size distribution of fly
ash–cement system on the fluidity of cement pastes', Cement and Concrete Research,
33(5), 763-768.

Li, W. and Fall, M. (2016) 'Sulphate effect on the early age strength and self-desiccation of
cemented paste backfill', Construction and Building Materials, 106, 296-304.

Mehta, P. K. and Monteiro, P. J. (2001) Concrete: microstrutture, Properties, and Materials,


University of California, Berkeley, USA.

Mojid, M. A. (2011) 'Diffuse Double Layer (DDL)' in Gliński, J., Horabik, J. and Lipiec, J., eds.,
Encyclopedia of Agrophysics, Dordrecht: Springer Netherlands, 213-214.

Nagele, E. (1986) 'The zeta-potential of cement: Part II: Effect of pH-Value', Cement and Concrete
Research, 16, 853-863.

Nasir, O. and Fall, M. (2010) 'Coupling binder hydration, temperature and compressive strength
development of underground cemented paste backfill at early ages', Tunnelling and
Underground Space Technology, 25(1), 9-20.

Niroshan, N., Sivakugan, N. and Veenstra, R. L. (2017) 'Laboratory Study on Strength


Development in Cemented Paste Backfills', Journal of Materials in Civil Engineering,
29(7).

Ontario Mining Association (2017) Ontario Miniral Production.

Palomo, A., Grutzeck, M. W. and Blanco, M. T. (1999) 'Alkali-activated fly ashes: a cement for
the future', Cement and Concrete Research, 29(8)(8), 1323-1329.

Panchal, S., Deb, D. and Sreenivas, T. (2018) 'Variability in rheology of cemented paste backfill
with hydration age, binder and superplasticizer dosages', Advanced Powder Technology,
29(9), 2211-2220.

83
Park, C. K., Noh, M. H. and Park, T. H. (2005) 'Rheological properties of cementitious materials
containing mineral admixtures', Cement and Concrete Research, 35(5), 842-849.

Potvin, Y., Thomas, E. G. and Fourie, A. B. (2005) Handbook on Mine fill, Australian Centre for
Geomechanics – University of Western Australia.

Qing-Hua, C., Tagnit-Hamou, A. and Sarkar, S. L. (1992) 'Strength and Microstructural Properties
of Water Glass Activated Slag', MRS Proceedings, 245.

Roussel, N. (2007) 'Rheology of fresh concrete: from measurements to predictions of casting


processes', Materials and Structures, 40(10), 1001-1012.

Roy, D. M. and Idorn, G. M. (1982) 'Hydration, structure, and properties of blast furnace slag
cements, mortars, and concrete', ACI Journal Proceedings, 79(6)(6), 444-457.

Sarda, S., Fernández, E., Llorens, J., Martinez, S., Nilsson, M. and Planell, J. A. (2001)
'Rheological properties of an apatitic bone cement during initial setting', Journal of
Materials Science: Materials in Medicine, 12(10-12)(10-12), 905-909.

Siddique, R. (2004) 'Properties of concrete incorporating high volumes of class F fly ash and san
fibers', Cement and Concrete Research, 34(1), 37-42.

Simon, D. and Grabinsky, M. (2013) 'Apparent yield stress measurement in cemented paste
backfill', International Journal of Mining, Reclamation and Environment, 27(4), 231-256.

Souquet, J. L., Nascimento, M. L. and Rodrigues, A. C. (2010) 'Charge carrier concentration and
mobility in alkali silicates', J Chem Phys, 132(3), 034704.

Tariq, A. and Nehdi, M. (2007) 'Developing durable paste backfill from sulphidic tailings',
Proceedings of the Institution of Civil Engineers - Waste and Resource Management,
160(4), 155-166.

Tariq, A. and Yanful, E. K. (2013) 'A review of binders used in cemented paste tailings for
underground and surface disposal practices', J Environ Manage, 131, 138-49.

Taylor, H. F. (1990) Cement Chemistry, 2nd ed., London: Academic press.

84
Thompson, B. D., Grabinsky, M. W., Bawden, W. F. and Counter, D. B. (2009) In-situ
measurements of cemented paste backfill in long-hole stope, translated by Toronto, ON,
Canada.

Wang, X. M., Li, J. X., Xiao, Z. Z. and Xiao, W. G. (2004) 'Rheological properties of tailing paste
slurry', Journal of Central South University of Technology, 11(1), 75-79.

Wang, Y., Fall, M. and Wu, A. (2016) 'Initial temperature-dependence of strength development
and self-desiccation in cemented paste backfill that contains sodium silicate', Cement and
Concrete Composites, 67, 101-110.

Wang, Y., Wu, A., Ruan, Z., Wang, H., Wang, Y. and Jin, F. (2018) 'Temperature Effects on
Rheological Properties of Fresh Thickened Copper Tailings that Contain Cement', Journal
of Chemistry, 2018, 1-8.

Warner, J. (2004) Practical handbook of grouting: soil, rock, and structures, John Wiley & Sons.

Wu, D., Fall, M. and Cai, S. J. (2013) 'Coupling temperature, cement hydration and rheological
behaviour of fresh cemented paste backfill', Minerals Engineering, 42, 76-87.

Xiapeng, P., Fall, M. and Haruna, S. (2019) 'Sulphate induced changes of rheological properties
of cemented paste backfill', Minerals Engineering, 141.

Xue, G., Yilmaz, E., Song, W. and Cao, S. (2018) 'Compressive Strength Characteristics of
Cemented Tailings Backfill with Alkali-Activated Slag', Applied Sciences, 8(9).

Yilmaz, E., Belem, T. and Benzaazoua, M. (2013) 'Study of physico-chemical and mechanical
characteristics of consolidated and unconsolidated cemented paste backfills', Gospodarka
Surowcami Mineralnymi - Mineral Resources Management, 29(1), 81-100.

Yilmaz, E. and Fall, M. (2017) PasteTailings Management, Cham, Switzerland: Springer


International Publishing.

Yin, S., Wu, A., Hu, K., Wang, Y. and Zhang, Y. (2012) 'The effect of solid components on the
rheological and mechanical properties of cemented paste backfill', Minerals Engineering,
35, 61-66.

85
Zheng, K., Zhou, J. and Gbozee, M. (2015) 'Influences of phosphate tailings on hydration and
properties of Portland cement', Construction and Building Materials, 98, 593-601.

Živica, V. (2007) 'Effects of type and dosage of alkaline activator and temperature on the properties
of alkali-activated slag mixtures', Construction and Building Materials, 21(7), 1463-1469.

86
CHAPTER 4
Paper II: Temperature Dependency of the Strength
Development of Slag-Cemented Paste Backfill that
Contains Sodium Silicate

87
Paper II: Temperature Dependency of the Strength Development of Slag-Cemented Paste
Backfill that Contains Sodium Silicate

(to be Submitted)

Ghada Ali, Mamadou Fall and Imad Alainachi

4.1 Abstract
Cemented paste backfill (CPB) is a novel backfill technology that is used in mining industry in
order to minimize the geotechnical and environmental hazards associated with the traditional mine
waste management. Also, CPB is used as a backfill to enhance the stability of the mine stope
(underground opening) and the adjacent area as well as to minimize the possible geotechnical risks,
such as ground subsidence. CPB consists of tailings, water, and hydraulic binder (such as cement
and blast furnace slag). Often, CPB is prepared and mixed in the paste plants located in the mine
area, and the fresh CPB is delivered into the mine stope through pipelines by gravity and/or
pumping. Recently, sodium silicate has been used as chemical activator for the hydraulic binders
used in CPB mixture to improve mechanical behavior of CPB. However, in order to achieve the
desired performance of CPB placed in the mine stope, it is necessary to study the strength of early
aged (up to 28 days) CPB. To do so, this study aimed to assess the strength of sodium silicate
activated slag-CPB under diverse conditions. Unconfined compression strength (UCS) of slag-
CPB was experimentally tested in this research for slag-CPB samples that contain sodium silicate
and were cured (up to 28 days) at different curing temperature (2oC, 20oC, and 35oC). Also, in
order to examine the effect of sodium silicate on the strength of CPB cured at different
temperatures, the slag-CPB mixtures that contained different content of sodium silicate (0.0%,
0.3%, and 0.5%) were studied in this research. The result of this study found that the strength
development of CPB increases with time and temperature. It was also found that the increase in
the sodium silicate dosage will enhance the mechanical strength of CPB at any temperature. The
findings of this research will effectively contribute to the safer design of CPB structures.

Keywords: Cemented paste backfill; Tailings; Cement; Strength, Sodium Silicate, Temperature;
Mine

4.2 Introduction
Mining industry has significantly improved the economy of several countries around the world.
For instance, the total mining production in 2017 was around 19 billion USD in Canada, and
around 600 billion USD around the globe (Alainachi and Fall 2019a, Statista 2019). However,
mining activities are associated with several challenges, such as generation of huge quantities of
mine wastes (e.g., tailings). These wastes may negatively impact the environment if they are
improperly disposed of (Farkish and Fall 2013, Alainachi and Fall 2019b). In the past decades, a
novel method called cemented paste backfill (CPB) was proposed in order to better manage the
disposal of the tailings. This technology is characterized by reusing these tailings and returning
them back to the underground mine stopes as a paste after mixing them with water and binder

88
materials (such as cement) (Belem and Benzaazoua 2004, Fall et al. 2005, Abdul-Hussain and Fall
2012). Using CPB method reduces the time needed to complete the mine stope backfilling, which
decreases the stope cycle (stope backfill process) and consequently improves the mine
productivity. Also, using tailings as main component of CPB reduces the amount of these
environmentally challenging wastes. Moreover, using CPB can also improve the stability of the
mine stope in short and long term, which will minimize the geotechnical risks (such as ground
subsidence and mine stope instability) that might endanger the safety of the workers at the mine
workplace and the surrounding area (Ghirian and Fall 2015, Alainachi and Fall 2019a). Therefore,
CPB has become a common technology in mining industry around the world (Brakebusch 1995,
Belem and Benzaazoua 2004, Aldhafeeri et al. 2016, Alainachi and Fall 2019c). However, the
mechanical strength or performance of CPB is significantly affected by several conditions, such
as exposure to variation in static loading, dynamic (seismic) loading as well as the potential
chemical interaction of the chemical elements within the CPB components (Cao et al. 2018, Xue
et al. 2018, Alainachi and Fall 2019b).

The typical components of CPB mixture are tailings with a content of around 70%-85%,
hydraulic binder of around 2%-9% of content (based on the stope requirements), and fresh or mine
processed water. The CBP is generally mixed and prepared in paste backfill plant usually located
on the mine surface and is then delivered into the mine stope through pipelines by gravity and/or
pumping (Yilmaz et al. 2004, Fall et al. 2005, Jamali 2012, Haiqiang et al. 2016, Koohestani et al.
2018).

In order to have safe and cost effective CPB structures, CPB mixtures must demonstrate
an appropriate stability performance as soon as it is placed in the mine stope. Mechanical strength
properties are significant factors in this regard. These properties can be evaluated by determining
the unconfined compression strength (UCS) of early aged CPB (Ghirian and Fall 2013, Cui and
Fall 2016).

It was learnt from previous studies, such as (Fall et al. 2010, Wu et al. 2013), that the
strength and pore-structure of fresh CPB placed in the mine stope are significantly affected by
temperature variation. There are several sources of the temperature variation (Figure 4.1), which
include (i) internal heat generated during cement hydration process (Fall et al. 2010, Alainachi and
Fall 2019c), (ii) temperature variation based on the geographic location of the mine, especially in
the permafrost regions and/or extremely hot regions (Wang et al. 2016), and (iii) heat generated
from the rock masses surrounding the mine stope that varies with the depth of the mine stope and
the geological condition of the mine area (Fall et al. 2007, Aldhafeeri et al. 2016). In addition, the
temperature of CPB in the mine stope might also be affected by other human-induced temperature
variations, such as the heat coming from the mine machinery, lighting, and ventilation, as well as
the heat generated due to the blasting operations (Fall et al. 2010). Anyways, the temperature
variation of CPB in the mine stope, regardless of its source, was found to be of significant influence
on the mechanical behavior (mechanical strength) of CPB (Fall et al. 2010, Fall and Pokharel
2010).

89
Figure 4.1 Sources of temperature change in CPB placed in mine stope.

Other than the temperature, there are other factors (external and internal) that might also
affect the mechanical strength of CPB, such as the density and concentration of CPB (binder type
and water content), and the availability of water-reducing admixture and/or chemical additives
(Yilmaz et al. 2004, Kesimal et al. 2005, Bentz 2006, Ercikdi et al. 2009, Ercikdi et al. 2010,
Abdul-Hussain and Fall 2012, Cihangir et al. 2012, Ghirian and Fall 2014, Kermani et al. 2015a,
Cihangir et al. 2018). However, some of these studies have addressed the mechanical strength of
CPB under the effect of individual factor. For instance, Kesimal et al. (2005), Ercikdi et al. (2009),
and Li and Fall (2018) studied the effect of different types and dosages of binders on the short-
term and long-term strength of CPB. Some other studies have been conducted recently in order to
understand the coupled effect of temperature on the mechanical strength of CPB that is made with
different binder contents. For instance, Fall et al. (2010) studied the impact of different curing
temperatures (2oC, 20oC, 35oC, and 50oC) on the strength of CPB made with different binders,
such as Portland cement (PCI), and blast furnace slag (Slag). This study revealed that the strength
of CPB increases as the curing temperature increases. Moreover, it was also noted that the partial
usage of slag (as a binder) has a significant impact on the temperature induced increase in CPB
strength. Nasir and Fall (2010) developed a numerical model to predict the coupled effect of
temperature and degree of hydration on the unconfined compression strength (UCS) of CPB. This
study found that increasing the initial temperature of CPB from 0oC to 20oC increases the strength
of CPB by 60%. Also, the early strength development of CPB is significantly affected by the binder
content. Furthermore, Pokharel and Fall (2011) compared the effect of curing temperature on the

90
strength development of CPB containing only PCI and CPB containing PCI and slag as binders. It
was revealed that at temperatures of 2oC and 35oC, the UCS of CPB containing PCI and slag is
higher than the UCS of CPB containing PCI only, while at the temperatures of 20oC and 50oC, the
UCS of CPB containing PCI and slag is lower than the UCS of CPB containing PCI only.

On the other hand, some studies (e.g. Brough and Atkinson 2002, Chang 2003, Živica
2007, Kermani et al. 2015a, Kermani et al. 2015b, Cihangir et al. 2018, Jiang et al. 2019) have
used sodium silicate as chemical additive to different mixtures of CPB. These studies addressed
the role of sodium silicate as activator for different binders of CPB (such as PCI and Slag) as well
as the effect of sodium silicate on the mechanical strength of CPB in the long term.

Up to date, there is a paucity of information and technical data on the effect of curing
temperature on the strength development of CPB that is made of PCI and Slag as binders and
contains sodium silicate of different dosages. In other words, no study has evaluated the combined
effect of curing time, temperature, binder content, and sodium silicate content on the strength of
CPB. Accordingly, the goal of this study is to experimentally examine the combined effect of
different curing temperatures (2oC, 20oC, and 35oC), dosages of sodium silicate (0.0%, 0.3% and
0.5%), and progress of cement hydration (curing time) on the strength of slag-cemented paste
backfill (Slag-CPB) mixtures.

4.3 Materials and Equipment Used in the Experiment

4.3.1 Materials

4.3.1.1 Tailings
Ground Silica Tailings (ST), which are synthetic tailings material (manufactured by the U.S. Silica
Co.), were used in this study. The physical properties of ST (Table 4.1) are similar to those of
Canadian mine tailings. For instance, ST are essentially made of quartz, which is the predominant
mineral in Canadian hard rock mine tailings. Also, the grain size distribution of ST is similar to
the average grain size distribution of nine main tailings (9MT) extracted from nine different mines
in eastern Canada (Figure 4.2). Thus, ST was selected in this study to be the foremost component
of CPB mixture. Moreover, ST has high percentage of silica (99.8% SiO2), which makes it a
chemically inter material (Table 4.2). It will control the mineralogical and chemical compositions
of the tailings and minimize the uncertainties related to the use of natural tailings (Aldhafeeri and
Fall 2016). The primary physical properties and main chemical elements of ST are illustrated in
Tables 4.1 and 4.2.

91
100
ST
90 9MT
80

70

60
Passing (%)

52
50

40

30

20

10

0
20
0.1 1 10 100 1000
Grain dimeter (Micrometer)
Figure 4.2 Grain size distribution of silica tailings used in this study and the average grain size
distribution of tailings from nine Canadian mines.
Table 4.1. Primary physical properties of tailings used in this study
Element Gs D10 (μm) D30 (μm) D50 (μm) D60 (μm) Cu Cc
ST 2.7 1.9 9.0 22.5 31.5 16.6 1.3

Table 4.2. Main chemical elements of tailings used in this study


Element K wt.% Pb wt.% S wt.% Ca wt.% Fe wt.% Na wt.% Al wt.% Si wt.%
ST 0.00 0.00 0.00 <0.01 <0.01 <0.01 0.10 99.80

4.3.1.2 Water and Binder


To enhance the mechanical properties of CPB, a sufficient amount of binder is often used in the
mixture. In this study, the binder agent used in the preparation of the CPB mixtures is Portland
cement type I (PCI), Fly Ash (FA), and blast furnace slag (Slag). These binders are the most
frequently used in preparing CPB worldwide (Haiqiang et al. 2016). The properties of these
binders are listed in Table 3.3. PCI is used alone or blended with Slag or FA. The blending ratio
of PCI and Slag or FA is 50/50. This ratio is selected based on: (i) the need of the industrial partner
in this project; (ii) cost savings; (iii) previous studies on performance of this ratio in CPB systems
(e.g., Fall et al. 2010). Tap water is used in this study to mix the binders with the tailings.

92
Table 4.3. Primary physical and chemical properties of binders used in this study

Binder SSA(1) (m2/g) R.D(2) S (wt%) Ca (wt%) Si (wt%) Al (wt%) Mg (wt%) Fe (wt%)

PCI 1.32 3.2 1.5 44.9 8.4 2.4 1.6 1.9


Slag 2.10 2.8 12 26.6 18.9 3.9 6.9 0.5
(1)
Specific surface area and
(2)
Relative density

4.3.1.3 Sodium silicate


Soluble sodium silicates (SS) are silicate polymers, which are inorganic chemical, clear, colorless, and
viscous liquid. SS are also known as water soluble glasses, which are generally produced from varied
proportions of an alkali metal and silicon dioxide (SiO2). There are several applications, wherein SS are
often used. For instance, SS are used in geotechnical engineering field in soil grouting and mine tailing
backfill. They are also used as an admixture for cement products. In this study, the main objective of using
SS was to evaluate its effect on the strength of CPB. In this regard, commercial solution SS (Type N) was
added in liquid form to the CPB mixture to be used as an admixture for the binders. Properties of SS used
in this study are shown in Table 4.4.

Table 4.4. Properties of the sodium silicate used in this study (national Silicate Ltd.).

Properties Values
o
Specific gravity @ 20 C 1.39
Weight ratio, %SiO2/%Na2O 3.22
Na2O% by weight 8.90
SiO2% by weight 28.66
Solids % 37.56

4.3.2 Sample Preparation


For the UCS tests, approximately 72 CPB specimens were prepared by mixing given amount of
tailings, binder (4.5 wt%), and water (w/c ratio =7.8). During the mixing of these ingredients,
sodium silicate was added to each mixture (each sample) with different concentrations of 0, 0.3,
and 0.5% by weight solid components. The sodium silicate gel, which is used in these mixtures,
consists of 62.44% of water. However, this water content was considered while measuring the
amount of mixing water of these samples. Samples were mixed for 7 min in order to obtain
homogenous mixture (Bian et al. 2019). Afterwards, the prepared CPB mixtures were poured into
a 600 ml low-form Griffin beaker and vibrated to remove the entrapped air. Samples were then
poured into plastic cylindrical containers (50 mm diameter × 100 mm height). To prevent water
evaporation, each sample was covered with plastic film. Samples were then securely cured at
different temperatures of 2oC, 20oC, and 35oC and curing times of 1, 3, 7, and 28 days in an
environmentally controlled chamber. Moreover, some cement paste samples of CPB with w/c = 1
(to mimic the high water content of CPB) and additional CPB samples were also prepared in the

93
similar manner as the aforementioned CPB specimens in order to conduct microstructural analyses
or monitoring tests. Table 4.5 below summarizes the testing program and the CPB mixing
composition for each phase of this program.

Table 4.5. Summary of CPB mix proportion and stages of experimental program

Binder %Slag
%PCI in the SS content Curing Curing Time
Sample ID Content in the W/C ratio
binder (%) Temp. (oC) (days)
(%) binder
A. Effect of curing temperature on the time-dependent evolution of strength of Slag-CPB that contain sodium
silicate
SS-CPB-2oC 4.5 50 50 7.8 0.3 2 1, 3, 7, 28

SS-CPB-20oC 4.5 50 50 7.8 0.3 20 1, 3, 7, 28

SS-CPB-35oC 4.5 50 50 7.8 0.3 35 1, 3, 7, 28

B. Effect of Sodium silicate content, temperature and curing time of CPB’s strength
0.0-SS-CPB 4.5 50 50 7.8 0.0 2, 35 1, 3, 7, 28

0.3-SS-CPB 4.5 50 50 7.8 0.3 2, 35 1, 3, 7, 28


0.5-SS-CPB
4.5 50 50 7.8 0.5 2, 35 1, 3, 7, 28

PCI: Portland cement I, SS: Sodium silicate.

4.3.3 Test Methods

4.3.3.1 Unconfined Compression Test (UCS) test


Unconfined Compressive tests (UCS) were conducted on the CPB specimens that were prepared
with different sodium silicate content under different temperatures and were cured under same
temperatures to different lengths of time. In this research, ASTM C39/C39M (2018) was followed
to conduct the UCS tests by using computer-controlled mechanical press of ELE Digital Tritest
50 load frame. The loading capacity and deformation rate of this test was defined as 50 kN and 1
mm/min respectively (Li and Fall 2018). To maintain reputability as well as to ensure accuracy of
gathered results, each test was repeated at least twice.

4.3.3.2 Microstructural analysis


To understand the effect of microstructural evolution of CPB on its strength development under
the influence of test conditions, several series of microstructural analyses were conducted on
selected CPB samples at several testing ages (curing time). Microstructural analyses encompass
thermal analysis (differential thermogravimetry (DTG), thermal gravimetry (TG), and Mercury
Intrusion Porosimetry (MIP) tests. Before conducting microstructural analysis, testing samples

94
were first dried at 45oC in a vacuum oven up to mass stabilization. Thermal analyses, on the other
hand, were undertaken using a TGA Q 5000 IR from TA Instruments. The various (dried) samples
(about 20 mg each) were heated in an inert nitrogen atmosphere at the rate of 10°C per minute up
to a temperature of 1000°C. MIP measurements were performed using Micromeretics AutoPore
III 9420 mercury porosimeter.

4.3.3.3 Monitoring of the specimens


To better understand the coupled effect of curing temperature and sodium silicate content, CPB
samples were prepared and poured into plastic cylinders (diameter of 100 mm and height of 200
mm), and placed in the controlled temperature chambers to be cured at the different curing
temperatures of 20oC and 35oC for 28 days. During curing period, CPB samples undergone
continuous monitoring in terms of the evolution of temperature, electrical conductivity (EC),
volumetric water content (VWC), and suction. Changes in EC reflect the rate of ion movement
due to the chemical reactions between cement and water. Monitoring EC is an effective way to
assess the cement hydration progress and the related structural changes (Li and Fall 2016). On the
other hand, monitoring the VWC enables to assess the self-desiccation of CPB (capillary water
consumed by the cement hydration) as well as the water flow within the CPB mass. The monitoring
of the temperature and suction enables to understand the progress of the cement hydration within
the tested specimens.

In this regard, the ECH2-5TE sensor was used to monitor EC, VWC, and temperature as
this sensor measures EC in the range of 0-23 dS/m with the accuracy of ±0.1, measures VWC in
the range of 0-80% with the accuracy of ±0.01 from 1-40% and the accuracy of ±0.15 from 40-
80%, whereas the temperature measurement accuracy is ±1°C. While dielectric water potential
sensor (ECH2-MPS6 sensor) was used to monitor the suction evolution. This sensor is designed
to measure soil water potential in the range of -9 to -100,000 kPa with a resolution of 0.1 kPa and
accuracy of (±10% of reading + 2 kPa, from -9 to -100 kPa). These sensors were installed within
the CPB specimens and connected to the Em50 series logger in order to record the data.

4.4 Results and Discussion

4.4.1 Effect of temperature on the strength development of Slag-CPB that contains Sodium
Silicate.
Figure 4.3 illustrates the effect of temperature on the strength of slag-CPB samples that contain
sodium silicate. As expected, regardless of the temperature conditions, there is a continuous
improvement in the mechanical strength of slag-CPB samples that contain sodium silicate with 1
day to 28 days curing time. This is mainly attributed to the progress of cement hydration, and the
mechanism of self-desiccation, which is involved in lowering the free water in the CPB material
(Bullard et al. 2011). It is also agreed that the partial usage of slag will increase the generation of
cement hydration products that will contribute to the strength (7 days and more of hydration) of
material cemented with PCI (Mehta and Monteiro 2001). This is because slag will be activated by

95
the CH released by cement clinker hydration and produce additional calcium silicates hydrate
(secondary C-S-H) (Li and Ding 2003, Tariq and Yanful 2013). In addition, it will trigger the role
of the alkali activators (such as sodium silicate) in activating the pozzolans binders (slag) leading
to the acceleration of its hydration process and the formation of more CH (Roy and Idorn 1982,
Kolani et al. 2012).

It can also be observed from figure 4.3 that the increase in curing temperature has a direct
relationship with strength development. Samples cured at the temperature of 35oC showed higher
values of UCS strength, while the samples cured at the temperatures of 2oC showed the lowest
values. This is because curing CPB samples at high temperatures accelerates the cement hydration
process (Aldhafeeri et al. 2016, Cui and Fall 2016) and produces more cement hydration products
(Wang et al. 2016). The accelerated cement hydration process will lead to more consumption of
the CPB pore-water, and the higher production of cement hydration products will result in the
refinement of the pores between CPB particles as well as an enhancement in the cohesion between
the CPB particles. Thus, the strength of CPB material will significantly increase (Alainachi and
Fall 2019c, Haruna and Fall 2019). Moreover, C-S-H was found to be the major binding phase in
hardened cement (Taylor 1990), so the more production of C-S-H, the more development of CPB
strength (Fall et al. 2010).

1800
Cured at 2°C
1600 Cured at 20°C
Cured at 35°C
1400

1200
Strength (kPa)

1000

800

600

400

200

0
0 5 10 15 20 25 30
Time (days)

Figure 4.3 Effect of temperature on the strength development of Slag-CPB that contain 0.3% sodium
silicate.

This argument is experimentally supported by the results of the microstructural analysis,


including thermal analysis (Figure 4.4) and MIP analysis (Figure 4.5). Figure 4.4 illustrates the
TG/DTG diagram for two slag-CPB samples that contain 0.3% of sodium silicate, which were

96
cured (up to 7 days) at different temperatures of 20oC and 35oC. It is evident from these diagrams
that more hydration products are produced within the slag-CPB that was cured at higher
temperature. The first peak (DTG) or change in weight (TG), which is found between 50 oC and
150oC and is the resultant of the dehydration reactions of hydrates (such as C-S-H, ettringite, and
gypsum), and the second peak, which is associated with generation of CH and is observed at 400-
500oC, are much higher for the CPB cured at 35oC as compared to the CPB cured at 20oC. While
the third peak and weight loss, which represents the decomposition of the calcite in the cement and
is found between 600oC and 700oC, was found to be lower within the CPB cured at 35oC as
compared to the CPB cured at 20oC. Figure 4.5, on the other hand, illustrates the differential pore
size distribution of 7-days slag-CPB samples that contain sodium silicate, which were cured at the
temperatures of 20oC and 35oC. It is observed that the threshold pore diameter decreases with the
increase in curing temperature. It indicates a refinement in the pore structure of the cementitious
material cured at high temperature as a result of the large amount of hydration products, and thus
higher mechanical strength (Aldhafeeri et al. 2016, Haruna and Fall 2020).
100 0.2
Cured at 20°C
0.18
Cured at 35°C
95 0.16
0.14

Deriv. Weight (%/°C)


90
0.12
Weight (%)

0.1
85
0.08
0.06
80
0.04

75 0.02
0
70 -0.02
0 200 400 600 800 1000
Temperature (°C)

Figure 4.4 DT/DTG diagrams for 7 days Slag-CPB specimens that contain 0.3 % sodium silicate, cured at
various temperatures (20oC and 35oC).
Moreover, the acceleration of cement hydration rates due to high curing temperature is also
evident in the EC monitoring results shown in Figure 4.6, which demonstrates the changes in EC
of slag-CPB samples that contain sodium silicate and were cured (up to 28 days) at different
temperatures of 20oC and 35oC. It can be observed from this figure that the time required for EC
within the CPB samples cured at 35oC to reach its maximum value is much less than those of CPB
samples cured at 20oC. In addition, the peak EC values of CPB samples cured at higher temperature
were much higher than those for samples cured at lower temperature. The CPB samples cured at

97
35oC reached the peak of around 4.0 mS/cm within around 3 hours, while it needed around 11
hours for CPB samples cured at 20oC to reach the peak EC of 3.2 mS/cm. It provides a clear
indication of the rapid increase in the rate of cement hydration due to a higher curing temperature,
and the higher concentration and higher mobility of ions within the CPB samples cured at higher
temperature (Li and Fall 2016, Fang and Fall 2018).

0.8
Cured at 20°C
0.7 Cured at 35°C

0.6
∆V/∆Log P (Diameter)

0.5

0.4

0.3

0.2

0.1

0
0.001 0.01 0.1 1 10 100 1000
Pore dimeter (Micrometer)

Figure 4.5 Differential pore size distribution curves of 7 days CPB specimens that contain 0.3% sodium
silicate, cured at various temperatures (20oC and 35oC).
Furthermore, there is a common consent that high curing temperature of CPB material
intensifies the cement self-desiccation of the material (Alainachi and Fall 2019c). Self-desiccation
is the shrinkage of pores due to the progress of cement hydration process, which will lead to a net
reduction in the total volume of the solids and water within the cementitious materials (Bentz
2008). This will consequently reduce the volumetric water content within the CPB pores, and
thereby decrease the pore-water pressure (PWP) and could lead to generation of negative pore-
water pressure (suction). Hence, the temperature-induced increase in self-desiccation will enhance
the mechanical strength of CPB (Abdul-Hussain and Fall 2012, Ghirian and Fall 2013). This
argument can also be supported by the results presented in Figure 4.7, which illustrate the evolution
of suction and volumetric water content (VWC) of slag-CPB samples that contain sodium silicate
and were cured (up to 28 days) at the temperatures of 20oC and 35oC. These results showed that
after 1 day of curing, there was a rapid decrease in VWC and rapid increase in suction values
(higher negative PWP) within the CPB sample cured at 35oC as compared to those cured at 20oC.
This will help in understanding the effect of curing temperature in accelerating the cement
hydration process and thereby enhancing the mechanical strength of CPB samples.

98
5.0
Cured at 20°C Cured at 35°C
4.5

4.0

E.C. (mS/cm) 3.5

3.0

2.5

2.0

1.5

1.0

0.5

0.0
0.01 0.10 1.00 10.00 100.00
Time (Days)

Figure 4.6 EC of the 28 days Slag-CPB samples that contain 0.3% sodium silicate, cured at different
temperatures of 20oC and 35oC.
1.0 0
VWC - Cured at 20°C
0.9 VWC - Cured at 35°C -25
Suction - Cured at 20°C
0.8 -50
Suction - Cured at 35°C
0.7 -75

Negative PWP (kPa)


VWC (m3/m3)

0.6 -100
0.5 -125
0.4 -150
0.3 -175
0.2 -200
0.1 -225
0.0 -250
0 5 10 15 20 25 30
Time (Days)

Figure 4.7 Coupled development of VWC and suction (negative PWP) of the 28 days Slag-CPB samples
that contain 0.3% sodium silicate, cured at different temperatures of 20oC and 35oC.

99
4.4.2 Coupled effect of sodium silicate content and curing temperature on CPB strength.
The effect of sodium silicate content on the strength development of slag-CPB samples with
respect to different curing temperatures is shown in Figure 4.8 (a and b). The increase in
temperature has significantly increased the UCS of CPB samples regardless of the sodium silicate
content, as shown in figure 8 (a and b). As discussed earlier, this is because of the high temperature-
induced acceleration of cement hydration processes, which leads to more production of cement
hydration products (Fall et al. 2010) and higher intensity of the self-desiccation within the CPB
samples (Alainachi and Fall 2019c).

Moreover, it is evident from Figure 4.8 (a and b) that the increase in sodium silicate content
increases the strength of CPB regardless of the temperature conditions. It is evident that at both
testing temperatures, the UCS of CPB that does not contain sodium silicate (0.0% of sodium
silicate) is lower than that of those samples that contain sodium silicate. Furthermore, it was found
that the CPB strength increases with the increase in the sodium silicate content. Indeed, the
increase in curing temperature of CPB that contains sodium silicate accelerates the role of sodium
silicate as alkali activator (Živica 2007). The high temperature-induced increase in activation
energy of sodium silicate is associated with the increase in absorption ability as a result of the
stress-induced sodium ion migration (Souquet et al. 2010). Accordingly, it is clearly observed from
Figure 8 (a and b) that the increase in UCS of CPB samples that contain sodium silicate was
relatively higher when cured at high temperature (35oC) as compared to those that were cured at
low temperature (2oC). It means that the sample with higher content of sodium silicate was more
affected (activated) at higher temperature, and thereby showed higher mechanical strength. This
finding can be attributed to two factors: (i) the effect of high temperature on enhancing the role of
sodium silicate as an activator (as discussed above) (Živica 2007); and (ii) the setting time of slag
pastes becomes relatively faster when exposed to high dosage of alkali activator (such as sodium
silicate) (Azadi et al. 2017), which enhances the mechanical strength of CPB (Warner 2004, Azadi
et al. 2017). This is also consistent with the findings of Kermani et al. (2015b), who found that the
final setting time of sodium silicate-free CPB is around 7.0 hours (420 minutes), while it is around
6.3 hours (375 minutes) and 5.8 hours (350 minutes) for CPB samples that contain 0.3% and 0.5%
of sodium silicate respectively.

This argument was experimentally confirmed by conducting microstructural analysis,


including thermal analysis (Figure 4.9) and MIP analysis (Figure 4.10). In Figure 4.9, the TG/DTG
diagrams of two slag-CPB samples, which contain 0.0% and 0.3% of sodium silicate, and were
cured (up to 7 days) at the temperature of 35oC, are presented. These diagrams confirm that the
presence of sodium silicate in CPB samples increases the cement hydration and generates more
amounts of hydration products.

100
1800
SS - 0.0%
1600 SS - 0.3% (a)
SS - 0.5%
1400
Strength (kPa) 1200

1000

800

600

400

200

0
0 5 10 15 20 25 30
Time (days)

1800
SS - 0.0%
SS - 0.3%
(b)
1600
SS - 0.5%
1400

1200
Strength (kPa)

1000

800

600

400

200

0
0 5 10 15 20 25 30
Time (days)

Figure 4.8 Effect of temperature and sodium silicate content on the strength development of Slag-CPB
that contain sodium silicate and cured at: (a) 35oC; (b) 2oC.

101
100 0.2
SS - 0.0%
SS - 0.3% 0.18
95 0.16
0.14

Deriv. Weight (%/°C)


90
0.12
Weight (%)

0.1
85
0.08
0.06
80
0.04

75 0.02
0
70 -0.02
0 200 400 600 800 1000
Temperature (°C)
Figure 4.9 Effect of sodium silicate content on TG/DTG diagrams for 7 days Slag-CPB specimens that
cured at 35oC.
It can be seen from the graphs in figure 4.9 that the sudden change in weight (TG curves)
indicates highest weight loss from the CPB sample that contains 0.3% of sodium silicate as
compared to the sodium silicate-free samples. In contrast, it is evident from the first peak in the
DTG curves, which occurred at about 100oC, that the amount of cement hydration products
(mainly C-S-H, ettringite, and gypsum) was higher in the samples that contain sodium silicate as
compared to the sample that was prepared without sodium silicate. Moreover, the second peak,
which occurred at about 450oC, shows more production of CH within the slag-CPB samples that
contain 0.3% sodium silicate as compared to the slag-CPB samples of 0.0% of sodium silicate.

Figure 4.10, on the other hand, shows the differential pore size distribution of two slag-
CPB samples that were prepared with and without sodium silicate, and were cured (up to 7 days)
at the temperature of 35oC. It is observed that due to the presence of sodium silicate, the threshold
pore diameter slightly decreases. In other words, the presence of sodium silicate will lead to a
slight refinement in the pore structure of the cementitious material as a result of the larger amount
of hydration products, which will improve the mechanical strength of CPB (Aldhafeeri et al. 2016,
Haruna and Fall 2020).

Furthermore, the EC monitoring results in Figure 4.11 support this assertion. It can be
observed from this figure that the EC within the CPB samples prepared with sodium silicate and
cured (up to 28 days) at 35oC reached its maximum value faster than the sodium silicate-free CPB
samples. In addition, the peak EC values of CPB samples that contain sodium silicate were much
higher than the peak EC values of samples prepared without adding sodium silicate. The CPB
samples that contain sodium silicate reached the peak of around 4.0 mS/cm within around 3.0

102
hours, while the one that does not contain sodium silicate reached the peak of 3.6 mS/m in around
4.0 hours. It provides a clear indication of the rapid increase in the rate of cement hydration due to
the presence of sodium silicate and the lower concentration and lower mobility of ions within the
sodium silicate-free CPB samples (Li and Fall 2016, Fang and Fall 2018).

1.4
SS - 0.0%
SS - 0.3%
1.2

1
∆V/∆Log P (Diameter)

0.8

0.6

0.4

0.2

0
0.001 0.01 0.1 1 10 100 1000
Pore dimeter (Micrometer)

Figure 4.10 Effect of sodium silicate content on the differential pore size distribution for 7 days Slag-CPB
specimens that cured at 35oC.
Additionally, Figure 4.12 presents the coupled evolution of suction and volumetric water
content (VWC) of two slag-CPB samples that were prepared with and without sodium silicate and
were cured (up to 28 days) at 35oC. The results presented in Figure 4.12 advocate these findings.
These results showed that there was a rapid decrease in VWC of both samples. However, the
reduction in VWC was higher within the CPB sample that contains sodium silicate. On the other
hand, evolution of the negative PWP (suction) was dramatically higher within the CPB sample
that contains sodium silicate as compared to the one that does not contain sodium silicate. It can
be explained as the role of sodium silicate in activating the slag. Hence, the cement hydration
process will be accelerated and will speed up the cement self-desiccation. It will consequently
reduce the volumetric water content within the CPB pores and decrease the water content and pore-
water pressure (PWP), and will consequently generate the negative pore-water pressure (suction)
and increase the mechanical strength of CPB (Bentz 2008, Abdul-Hussain and Fall 2012).
Accordingly, this fact provides an additional support to the idea of the effect of the presence of
sodium silicate in accelerating the cement hydration process, and consequently improving the
mechanical strength of CPB samples.

103
5.0
SS - 0.0% SS - 0.3%
4.5
4.0

E.C. (mS/cm) 3.5


3.0
2.5
2.0
1.5
1.0
0.5
0.0
0.01 0.10 1.00 10.00 100.00
Time (Days)
Figure 4.11 Effect of sodium silicate content on the EC of 7 days Slag-CPB specimens that cured at 35oC.

1.0 VWC - SS - 0.0%


0
VWC - SS - 0.3%
0.9 Suction - SS - 0.0%
-25
Suction - SS - 0.3%
0.8 -50

0.7 -75
VWC (m3/m3)

Suction (kPa)
0.6 -100

0.5 -125

0.4 -150

0.3 -175

0.2 -200

0.1 -225

0.0 -250
0.00 5.00 10.00 15.00 20.00 25.00 30.00
Time (Days)
Figure 4.12 Effect of sodium silicate content on the VWC and suction of 7 days Slag-CPB specimens that
cured at 35oC.

4.5 Summary and Conclusion


This manuscript mainly studied the combined effect of temperature and sodium silicate content on
the time-dependent development of mechanical strength (UCS) of CPB material at early ages (up
to 28 days). Thus, the study revolves around two main axes:

104
a. The effect of temperature on the strength of Slag-CPB material that contains 0.3% of
sodium silicate.

b. The effect of temperature on the strength of Slag-CPB material that contains different
dosages of sodium silicate (0.0%, 0.3%, and 0.5% of the total dry weight).

A total of 72 CPB specimens were prepared at different temperature (2C, 20C, 35C), and
were subjected to UCS tests. To better understand the reason behind the nature of the results
observed, these samples were also subjected to microstructural analysis (including thermal
analysis, and MIP), monitoring of electrical conductivity, volumetric water content, and suction.
The major conclusions, based on the results of this research, can be summarized as:

1. Generally, the strength of slag-CPB, regardless of temperature and/or content of sodium


silicate, increases with the progress of curing time. However, the rate of increase differs with
the change in any of the other conditions.
2. Increase in curing temperature accelerates the cement hydration progress and consequently
increases the strength of slag-CPB material. The temperature-dependent increase is evident
regardless of the content of sodium silicate. However, the amount of increase in the
mechanical strength with temperature is affected by the change in sodium silicate content.
3. Adding sodium silicate to the CPB significantly increases the strength of CPB at any
temperature.
4. The UCS increment is directly proportional to the increase in the dosage of sodium silicate
regardless of the change in temperature. However, at high temperature (35C), there is a
significant increase in strength of slag-CPB samples that contain high dosage of sodium
silicate (0.3% and 0.5%) as compared to those that do not contain sodium silicate.
5. The results of this research will contribute to more efficient and safer design of CPB structures
in mining areas.

4.6 Data Availability Statement


All data, models, and code generated or used during the study appear in the submitted article.

4.7 Acknowledgements
Authors of this research would like to thank Sada Haruna and Jean-Claude Célestin for their help
during the experimental program. They also thank the Natural Sciences and Engineering Research
Council (NSERC) of Canada for their financial support, and Lafarge Canada for providing the
pozzolans (slag).

4.8 References

Abdul-Hussain, N. and Fall, M. (2012) 'Thermo-hydro-mechanical behaviour of sodium silicate-


cemented paste tailings in column experiments', Tunnelling and Underground Space
Technology, 29, 85-93.

105
Alainachi, I. and Fall, M. (2019a) 'Liquefaction potential of cementing tailing backfill: shaking
table test results', in Proceedings of 16th Panamerican conference on soil mechanics and
geotechnical engineering, Cancun, Mexico, IOS Press, 1949-1956.

Alainachi, I. and Fall, M. (2019b) 'Seismic induced liquefaction of cemented paste backfill: effect
of mixing water', in Proceedings of 7th International Conference on Earthquake
Geotechnical Engineering, Rome, Italy, 17-20 June, 2019,

Alainachi, I. and Fall, M. (2019c) 'Shaking table testing of liquefaction potential of cemented paste
backfill: Effects of backfill temperature', in Geo St-John's 2019 – the 72th Canadian
Geotechnical Conference (CGC), Sept. 29- Oct 2, 2019, St. John's, NL, Canada, September
29 - October 2, 2019,

Aldhafeeri, Z. and Fall, M. (2016) 'Time and damage induced changes in the chemical reactivity
of cemented paste backfill', Journal of Environmental Chemical Engineering, 4(4), 4038-
4049.

Aldhafeeri, Z., Fall, M., Pokharel, M. and Pouramini, Z. (2016) 'Temperature dependence of the
reactivity of cemented paste backfill', Applied Geochemistry, 72, 10-19.

ASTM C39/C39M (2018) 'Standard test method for compressive strength of cylindrical concrete
specimens',

Azadi, M. R., Taghichian, A. and Taheri, A. (2017) 'Optimization of cement-based grouts using
chemical additives', Journal of Rock Mechanics and Geotechnical Engineering, 9(4), 623-
637.

Belem, T. and Benzaazoua, M. (2004) An overview on the use of paste backfill technology as a
ground support method in cut and fill mines, translated by Perth, Australia: CRC Press,
637–650.

Bentz, D. P. (2006) 'Influence of water-to-cement ratio on hydration kinetics: Simple models based
on spatial considerations', Cement and Concrete Research, 36(2), 238-244.

Bentz, D. P. (2008) 'A review of early-age properties of cement-based materials', Cement and
Concrete Research, 38(2), 196-204.

106
Bian, J., Fall, M. and Haruna, S. (2019) 'Sulphate induced changes in rheological properties of
fibre-reinforced cemented paste backfill', Magazine of Concrete Research, 1-35.

Brakebusch, F. W. (1995) 'Basics of paste backfill systems', International Journal of Rock


Mechanics and Mining Sciences and Geomechanics Abstracts, 3(32), 122A.

Brough, A. R. and Atkinson, A. (2002) 'Sodium silicate-based, alkali-activated slag mortars: Part
I. Strength, hydration and microstructure', Cement and Concrete Research, 32(6)(6), 865-
879.

Bullard, J. W., Jennings, H. M., Livingston, R. A., Nonat, A., Scherer, G. W., Schweitzer, J. S.,
Scrivener, K. L. and Thomas, J. J. (2011) 'Mechanisms of cement hydration', Cement and
Concrete Research, 41(12), 1208-1223.

Cao, S., Yilmaz, E. and Song, W. (2018) 'Dynamic response of cement-tailings matrix composites
under SHPB compression load', Construction and Building Materials, 186, 892-903.

Chang, J. J. (2003) 'A study on the setting characteristics of sodium silicate-activated slag pastes',
Cement and Concrete Research, 33(7), 1005-1011.

Cihangir, F., Ercikdi, B., Kesimal, A., Ocak, S. and Akyol, Y. (2018) 'Effect of sodium-silicate
activated slag at different silicate modulus on the strength and microstructural properties
of full and coarse sulphidic tailings paste backfill', Construction and Building Materials,
185, 555-566.

Cihangir, F., Ercikdi, B., Kesimal, A., Turan, A. and Deveci, H. (2012) 'Utilisation of alkali-
activated blast furnace slag in paste backfill of high-sulphide mill tailings: Effect of binder
type and dosage', Minerals Engineering, 30, 33-43.

Cui, L. and Fall, M. (2016) 'Mechanical and thermal properties of cemented tailings materials at
early ages: Influence of initial temperature, curing stress and drainage conditions',
Construction and Building Materials, 125, 553-563.

Ercikdi, B., Cihangir, F., Kesimal, A., Deveci, H. and Alp, I. (2010) 'Utilization of water-reducing
admixtures in cemented paste backfill of sulphide-rich mill tailings', J Hazard Mater,
179(1-3), 940-6.

107
Ercikdi, B., Kesimal, A., Cihangir, F., Deveci, H. and Alp, İ. (2009) 'Cemented paste backfill of
sulphide-rich tailings: Importance of binder type and dosage', Cement and Concrete
Composites, 31(4), 268-274.

Fall, M., Benzaazoua, M. and Ouellet, S. (2005) 'Experimental characterization of the influence of
tailings fineness and density on the quality of cemented paste backfill', Minerals
Engineering, 18(1), 41-44.

Fall, M., Célestin, J. C., Pokharel, M. and Touré, M. (2010) 'A contribution to understanding the
effects of curing temperature on the mechanical properties of mine cemented tailings
backfill', Engineering Geology, 114(3-4), 397-413.

Fall, M., Nasir, O. and Celestine, J. (2007) 'Paste backfill responses in deep mine temperature
conditions', in Symposium minefill, Montreal, Canada, CD-Room,

Fall, M. and Pokharel, M. (2010) 'Coupled effects of sulphate and temperature on the strength
development of cemented tailings backfills: Portland cement-paste backfill', Cement and
Concrete Composites, 32(10), 819-828.

Fang, K. and Fall, M. (2018) 'Effects of curing temperature on shear behaviour of cemented paste
backfill-rock interface', International Journal of Rock Mechanics and Mining Sciences,
112, 184-192.

Farkish, A. and Fall, M. (2013) 'Rapid dewatering of oil sand mature fine tailings using super
absorbent polymer (SAP)', Minerals Engineering, 50-51, 38-47.

Ghirian, A. and Fall, M. (2013) 'Coupled thermo-hydro-mechanical–chemical behaviour of


cemented paste backfill in column experiments. Part I: Physical, hydraulic and thermal
processes and characteristics', Engineering Geology, 164, 195-207.

Ghirian, A. and Fall, M. (2014) 'Coupled thermo-hydro-mechanical–chemical behaviour of


cemented paste backfill in column experiments', Engineering Geology, 170, 11-23.

Ghirian, A. and Fall, M. (2015) 'Coupled Behavior of Cemented Paste Backfill at Early Ages',
Geotechnical and Geological Engineering, 33(5), 1141-1166.

Haiqiang, J., Fall, M. and Cui, L. (2016) 'Yield stress of cemented paste backfill in sub-zero
environments: Experimental results', Minerals Engineering, 92, 141-150.

108
Haruna, S. and Fall, M. (2019) 'Time- and temperature-dependent rheological properties of
cemented paste backfill that contains superplasticizer', Powder Technology.

Haruna, S. and Fall, M. (2020) 'Strength Development of Cemented Tailings Materials Containing
Polycarboxylate ether-based Superplasticizer: Experimental Results on the Effect of Time
and Temperature', Canadian Journal of Civil Engineering.

Jamali, M. (2012) Effect of Binder Content and Load History on the One dimensional Compression
of Williams Mine Cemented Paste Backfill, Master of Applied Science dissertation:
University of Toronto.

Jiang, H., Qi, Z., Yilmaz, E., Han, J., Qiu, J. and Dong, C. (2019) 'Effectiveness of alkali-activated
slag as alternative binder on workability and early age compressive strength of cemented
paste backfills', Construction and Building Materials, 218, 689-700.

Kermani, M., Hassani, F. P., Aflaki, E., Benzaazoua, M. and Nokken, M. (2015a) 'Evaluation of
the effect of sodium silicate addition to mine backfill, Gelfill − Part 1', Journal of Rock
Mechanics and Geotechnical Engineering, 7(3), 266-272.

Kermani, M., Hassani, F. P., Aflaki, E., Benzaazoua, M. and Nokken, M. (2015b) 'Evaluation of
the effect of sodium silicate addition to mine backfill, Gelfill – Part 2: Effects of mixing
time and curing temperature', Journal of Rock Mechanics and Geotechnical Engineering,
7(6), 668-673.

Kesimal, A., Yilmaz, E., Ercikdi, B., Alp, I. and Deveci, H. (2005) 'Effect of properties of tailings
and binder on the short-and long-term strength and stability of cemented paste backfill',
Materials Letters, 59(28), 3703-3709.

Kolani, B., Buffo-Lacarrière, L., Sellier, A., Escadeillas, G., Boutillon, L. and Linger, L. (2012)
'Hydration of slag-blended cements', Cement and Concrete Composites, 34(9), 1009-1018.

Koohestani, B., Khodadadi Darban, A. and Mokhtari, P. (2018) 'A comparison between the
influence of superplasticizer and organosilanes on different properties of cemented paste
backfill', Construction and Building Materials, 173, 180-188.

Li, W. and Fall, M. (2016) 'Sulphate effect on the early age strength and self-desiccation of
cemented paste backfill', Construction and Building Materials, 106, 296-304.

109
Li, W. and Fall, M. (2018) 'Strength and self-desiccation of slag-cemented paste backfill at early
ages: Link to initial sulphate concentration', Cement and Concrete Composites, 89, 160-
168.

Li, Z. and Ding, Z. (2003) 'Property improvement of Portland cement by incorporating with
metakaolin and slag', Cement and Concrete Research, 33(4), 579-584.

Mehta, P. K. and Monteiro, P. J. (2001) Concrete: microstrutture, Properties, and Materials,


University of California, Berkeley, USA.

Nasir, O. and Fall, M. (2010) 'Coupling binder hydration, temperature and compressive strength
development of underground cemented paste backfill at early ages', Tunnelling and
Underground Space Technology, 25(1), 9-20.

Pokharel, M. and Fall, M. (2011) 'Coupled Thermochemical Effects on the Strength Development
of Slag-Paste Backfill Materials', Journal of Materials in Civil Engineering, 23(5), 511-
525.

Roy, D. M. and Idorn, G. M. (1982) 'Hydration, structure, and properties of blast furnace slag
cements, mortars, and concrete', ACI Journal Proceedings, 79(6)(6), 444-457.

Souquet, J. L., Nascimento, M. L. and Rodrigues, A. C. (2010) 'Charge carrier concentration and
mobility in alkali silicates', J Chem Phys, 132(3), 034704.

Statista (2019) 'Total revenue of the top mining companies worldwide from 2002 to 2017', [online],
available: https://www.statista.com/statistics/208715/total-revenue-of-the-top-mining-
companies/ 2019].

Tariq, A. and Yanful, E. K. (2013) 'A review of binders used in cemented paste tailings for
underground and surface disposal practices', J Environ Manage, 131, 138-49.

Taylor, H. F. (1990) Cement Chemistry, 2nd ed., London: Academic press.

Wang, Y., Fall, M. and Wu, A. (2016) 'Initial temperature-dependence of strength development
and self-desiccation in cemented paste backfill that contains sodium silicate', Cement and
Concrete Composites, 67, 101-110.

Warner, J. (2004) Practical handbook of grouting: soil, rock, and structures, John Wiley & Sons.

110
Wu, D., Fall, M. and Cai, S. J. (2013) 'Coupling temperature, cement hydration and rheological
behaviour of fresh cemented paste backfill', Minerals Engineering, 42, 76-87.

Xue, G., Yilmaz, E., Song, W. and Cao, S. (2018) 'Compressive Strength Characteristics of
Cemented Tailings Backfill with Alkali-Activated Slag', Applied Sciences, 8(9).

Yilmaz, E., Kesimal, A. and Ercidi, B. (2004) Strength development of paste backfill simples at
Long term using different binders, translated by China: 281-285.

Živica, V. (2007) 'Effects of type and dosage of alkaline activator and temperature on the properties
of alkali-activated slag mixtures', Construction and Building Materials, 21(7), 1463-1469.

111
CHAPTER 5
Synthesis and Integration of Results

112
5.1 Introduction
The results obtained from the two technical papers, presented in Chapters 3 and 4 of this thesis,
have been synthesized in this Chapter. Each of these papers has investigated the combined effect
of several factors on the rheological properties and mechanical strength of fresh CPB material as
shown in Table 5.1. Table 5.2 summarizes the experimental tests that were conducted in each
paper.

Table 5.1 Summary of effective factors that were investigated in this thesis
Temperature

Sodium
Binder
Initial
Technical Paper

Curing Time Silicate


Blend
Chapter

Content

PCI/Slag
0.25 hrs.

PCI/FA
28 days
0.0 hrs.

1.0 hrs.
2.0 hrs.
4.0 hrs.
1 days
3 days
7 days

0.1%
0.3%
0.5%
20oC
35oC

PCI
2oC

0%
3 1 × × × × × × × × × × × × × × ×
4 × × ×2 × × × × × × × ×
PCI: Portland Cement Type I; Slag: Blast Furnace Slag FA: Fly Ash

Table 5.2 Summary of experimental tests conducted in this thesis

Rheological Microstructural Monitoring


Technical Paper

Zeta Potential

Properties Analysis Tests


Chapter

UCS

pH
Yield Stress

Viscosity

TG/DTG

Suction

VWC
XRD

MIP

EC

3 1 × × × × × × ×
4 2 × × × × × ×
UCS: Unconfined Compression Strength; TG/DTG: Thermal Analysis; XRD: X-ray Diffraction;
MIP: Mercury Intrusion Porosimetry; EC: Electrical Conductivity; VWC: Volumetric Water Content.

113
Section 5.2 of this chapter presents the representative results from Chapter 3 to show the
combined effect of the curing time, initial temperature, binder type, and sodium silicate dosage on
the rheological properties (yield stress and viscosity) of CPB samples that were: (i) cured for
different curing time (0 hrs, 0.25 hrs, 1 hr, 2 hrs, and 4 hrs), (ii) mixed and cured at different
temperatures (2oC, 20oC, and 35oC), (iii) prepared with binder blends of 100% PCI, 50%PCI-
50%Slag, and 50%PCI-50%FA, and (iv) contained different dosages of sodium silicate (0%, 0.1%,
0.3% and 0.5%). Afterwards, Section 5.2 presents the main results obtained from Chapter 4 which
investigated the combined effect of curing time, curing temperature, and content of sodium silicate
on the mechanical strength of fresh CPB by conducting the UCS tests on Slag-CPB samples that
were cured for 1 day, 3 days, 7 days, and 28 days under different curing temperatures of 2oC, 20oC,
and 35oC), and contained 0%, 0.3% and 0.5% of sodium silicate.

5.2 Effect of Time and Temperature on the Rheological Properties of Cemented Paste
Backfill that Contrains Sodium Silicate
In Chapter 3, the effect of cement hydration progress (curing time), initial (mixing and curing)
temperature, binder type, and dosage of sodium silicate on the rheological properties (yield stress
and viscosity) of fresh CPB samples were assessed. Yield stress and viscosity tests, pH and zeta
potential tests, microstructural analysis (TG/DTG and XRD), and electrical conductivity
monitoring tests were conducted on 125 CPB samples that were prepared with different binder
blends (PCI, PCI/Slag and PCI/FA) and contained different dosages of sodium silicate. These
samples were prepared and cured under different temperatures of 2oC, 20oC, and 35oC and were
cured for 0 hrs, 0.25 hrs, 1 hr, 2 hrs, and 4 hrs of curing time.

It has been observed that the rheological properties of CPB increase with the increase in
curing time owing to the progress of cement hydration process with time. They also increase with
the increase in initial temperature because the high mixing and curing temperature accelerates the
cement hydration process and produces more cement hydration products. In addition, they increase
with the increase in sodium silicate content due to the increase of sodium silicate rule in activating
the pozzolanic compounds (slag and FA) and accelerating the hydration of the clinker and binder.
It was also found that the partial usage of slag or FA as a binder with PCI reduces the yield stress
and viscosity (enhances the flowability) of the CPB samples with sodium silicate. The
enhancement in the flowability of CPB with the partial usage of FA was more evident than the
case of the partial usage of slag. These understandings of different behavior of the rheology of
CPB can help in designing CPB mixtures with better flowability and workability, and thus increase
the efficiency of these designs.

It is commonly known that 10% - 20% of the total cost of the mining operations in
underground mines are related to the cost of the backfilling processes. However, 70% of the
backfill system failures were found to be attributed to the failure in backfill delivery system caused
by the transportation of backfill material that does not have appropriate flowability, leading to
possible flow delays and/or temporary interruption of the progress of CPB production as well as

114
possible replacement of the pipelines (Grice 1998, Fehrsen and Cooke 2006, Haruna and Fall
2019). So, understanding the flow behavior of backfill can help in saving large amount of the
possible related financial losses. In addition, shortening the time required to deliver the backfill
into the mine stope can reduce the cost of the backfilling process by allowing the mine operations
to be conducted at faster rate. Therefore, many mines have implemented several approaches to
accelerate the backfill transportation, such as increasing the flow rate, increasing the fill slump by
adding water, and using pipelines of high diameters, and/or using high-pressure pumps.
Nevertheless, these approaches might have negative impacts on the backfill process. For instance,
increasing the fill flow rate has a great influence on the pressure gradient and causes higher stresses
to be applied to the walls of pipelines and/or the stope barricade, which might exceed their
capacities and cause damages to the pipelines, and/or failure of the stope barricade. Also, adding
water may result in increasing the water-cement ratio of the fill mix, and thereby compromising
the desired strength. Despite the advantage of using pipes of higher diameter and/or using the high-
pressure pumps, such as accelerating the delivery process and avoiding the laminar flow setting
(which occurs while transporting viscus material), this approach may have significant cost
implication in some situations, especially while mining small size mines (Clark et al. 1995, Fehrsen
and Cooke 2006, Paterson 2012).

Accordingly, the results obtained in this study show some technical implications in the
practice of the backfill technology by understanding the factors that influence the flow behavior
of the backfill, and thus reducing the risk of failure in backfill delivery system in addition to using
cement-slag or cement-FA binder blends that reduce the yield stress and viscosity of the fill
mixture, which will accelerate the backfill process without damaging the pipes and/or risking the
stability of the stope barricade, and without compromising the desired strength by adding sodium
silicate to the mixture. Moreover, these implementations can also help in avoiding laminar flow
setting without having to afford the high cost of using high-diameter pipes and/or high-pressure
pumps.

5.3 Effect of Temperature on the Strength Development of Cemented Paste Backfill that
Contrains Sodium Silicate
In Chapter 4, the combined effect of curing time and temperature on the mechanical behavior of
slag-CPB that contains sodium silicate has been studied by conducting UCS tests, microstructural
analysis (TG/DTG and MIP), and monitoring tests (electrical conductivity, suction and volumetric
water content) on 72 CPB samples that were prepared with 50%PCI-50%Slag blend, cured for 1,
3, 7, and 28 days under different curing temperatures (2oC, 20oC, and 35oC), and contained sodium
silicate of different content (0%, 0.3% and 0.5%).

It has been found that the strength of slag-CPB samples increases with time regardless of
the variation of temperature and/or sodium silicate content because of the progress of cement
hydration process with time. However, the rate of increase in the strength of CPB increases with
the increase in curing temperature and the increase in the dosage of sodium silicate. This is

115
attributed to the acceleration of cement hydration process and increase in the generation of the
cement hydration products with the increase in curing temperature as well as the role of sodium
silicate in activating the slag and thus accelerating the binder hydration, which will then increase
the bonds between the CPB particles. These results may help in designing better performing and
safer CPB structures.

In addition, the obtained results may have some practical implications in the CPB
technology that have been described briefly below:

5.3.1 Mechanical Stability of CPB Structures


One of the most important goals of the CPB technology is to provide an adequate ground support
against the mining-induced geotechnical issues, such as the ground subsidence and the instability
of the mine stope, which might danger safety of the mine workers and the surrounding area. To
achieve the desired mechanical performance of CPB, it is imperative to maintain the mechanical
stability or strength of the CPB structure in high standards. For instance, it is indicated that the
required UCS for the CPB in a common underground mining operation is 0.7–2 MPa (Brackebusch
1995, Fall et al. 2010). However, this UCS widely changes, depending on the sizes of the CPB
mass, application or role of the backfill structure and mine characteristics. Therefore, assessing the
factors that affect the mechanical strength of CPB, potentially the UCS, is of high importance.
These factors, as described in this thesis, include the combined effect of temperature, and the usage
of sodium silicate. Several researchers have studied the UCS as a function of the stope size, which
is also called the critical compression strength (UCSc). The determination of UCSc can help
evaluate the mechanical stability in faster and less expensive way. Mitchell (1983) determined the
critical UCS for a free standing CPB structure as: 𝑈𝐶𝑆𝑐 = 𝛾𝐻⁄(1 + 𝐻 ⁄𝐿), where, 𝐻 is the height
of the stope, 𝐿 is the width of the stope, and 𝛾 is the unit weight of the fill. Figure 5.1 shows the
variation of the critical strength of CPB with respect to the stope size.

As shown in Figure 5.1, the UCSc of a stope size of 60 m (H) × 40 m (L) is around 500
kPa. By analyzing these values with the UCS values of CPB obtained from this study (Figures 4.3
and 4.8 in Chapter 4 of this study), it can be seen that the UCS of 28 days-aged CPB material that
does not contain sodium silicate and was cured at low temperature (2oC) was around 100 kPa,
while by increasing the curing temperature to 35oC, the 28 days-UCS reached a value of around
1200 kPa. Moreover, by adding 0.3% of sodium silicate to the mixture and curing it at 35oC, the
28 days-UCS value was more than 1500 kPa. In other words, high curing temperature or adding
sodium silicate will enhance the strength of CPB, and combining the high curing temperature and
sodium silicate will significantly increase the strength of CPB. Owing to these findings, CPB that
contains sodium silicate and was cured at high temperature will have higher mechanical strength
as compared to the CPB that does not contain sodium silicate and was cured at low temperatures.

116
600
UCSc
500

UCS (kPa) 400

300

200

100

0
0 10 20 30 40 50
Stope Width (m)

Figure 5.1 Critical UCS of CPB with respect to the width variation of a 60 m height stope.

5.3.2 Barricade Stability


Before completing the filling process of CPB, retaining wall structure (named stope barricade)
must be constructed at the bottom of the mine stope (at the entrance of the mine drift) in order to
hold the CPB in place. To ensure the safety of the workers of the mine as well as the equipment,
it is imperative to ensure that the stresses applied by the fresh backfill does not exceed the barricade
capacity (strength). The applied stresses on the barricade was found to be directly associated to the
horizontal total stress and PWP of the backfill material. It was also noted that these stresses are
significantly influenced by temperature increase due to cement hydration reaction, as this increase
in temperature will accelerate the cement hydration process and reduce the PWP acting on the
barricade (Ghirian 2016).

The result obtained from the current study revealed that the high curing temperature will
accelerate the cement hydration process, produce more cement hydration products, and reduce the
water content (and consequently reduce the PWP and total stress) of the CPB material. It was also
found that using sodium silicate as a chemical additive will activate the hydraulic binder,
particularly the Slag, and release more of the hydration heat. This will significantly increase the
hydration temperature, accelerate the hydration process, and consequently reduce the PWP and
total stress applied to the barricade. These findings may be taken into consideration during the
backfill design to help mine operators work beside a stable barricade, and thereby in safe work
environment.

117
5.3.3 Mining Cycles
From financial point of view, the total cost of the mining operation highly depends on the
mining cycle time (stope cycle), which represents the time required to complete the whole process
in a single mine stope. Longer mining cycle of a mine stope causes a delay in mining the adjacent
ore reserves. Shortening the mining cycle time, on the other hand, will allow the adjacent stope to
be mined, so the mining productivity will be enhanced, more cost savings, and consequently more
financial revenues will be profited from the mine. Rather than accelerating the filling rate which
might affect the stability of the stope barricade, continuous pouring of backfill into the mine stope
can help in reducing the operation cost and mining cycle time without risking the stability of the
stope barricade. Thus, the mine operators will be able to develop optimal stope cycle for opening
barricades and proceed with mining sequences. However, adopting continuous pour process
necessitates significant evolution of backfill strength, especially at the early hours (or days) of
curing to avoid the risk of backfill and barricade failure.

The enhancement in strength evolution of CPB at early age by adding sodium silicate and
curing at high curing temperature, shown in the current study, can be correlated with the critical
UCS (UCSc) as mentioned in section 5.3.1. For instance, the UCSc of a stope size of 60 m (H) ×
40 m (L), as shown in Figure 5.1, is around 500 kPa, and by analyzing the UCS values of CPB
obtained from this study (Figures 4.3 and 4.8 in Chapter 4 of this study), it can be seen that the
UCS of CPB material that does not contain sodium silicate and was cured at 2oC did not reach
more than 100 kPa in 28 days of curing time, while by curing the same material at 35oC, the UCS
reached the value of 500 kPa in only 3 days of curing time. This is because the high curing
temperature was known to accelerate the cement hydration process. Moreover, by adding 0.3% of
sodium silicate to the mixture and curing it at 35oC, the UCS value exceeded the critical UCS of
500 kPa and reached a UCS of around 650 kPa in less than one day of curing time. This is related
to the fact that the presence of sodium silicate will shorten the setting time of cemented material.
In other words, high curing temperature or adding sodium silicate will enhance the strength of
CPB, and combining the high curing temperature and sodium silicate will reduce the mine cycle
time by significantly increasing the UCS of CPB in very short time, and thereby the mine
productivity will potentially be improved.

5.3.4 Binder Consumption


Despite the fact that the binder used in CPB preparation might typically cost around 75% of the
backfilling process, it is common practice in many underground mines to use high binder content
within the fill in order to accelerate its strength development, especially while placing the first 2-
3 m thick layer (named Plug fill) (Grice 1998, Yilmaz et al. 2015). Accordingly, an obvious cost
saving can be significantly achieved by optimizing the binder consumption. In this study, it was
revealed that the UCS of CPB cured at high curing temperature with the presence of sodium silicate
is shown to result in significant improvement in the mechanical strength of the backfill. This
approach can help reduce the cost of mining operation by reducing the binder usage, and thus
improve the mine productivity.

118
5.4 References

Brackebusch, F. (1995) Basics of paste backfill systems, translated by 122A.

Clark, C., Vickery, J. and Backer, R. R. (1995) 'Transport of total tailings paste backfill: results of
full-scale pipe test loop pumping tests'.

Fall, M., Célestin, J. C., Pokharel, M. and Touré, M. (2010) 'A contribution to understanding the
effects of curing temperature on the mechanical properties of mine cemented tailings
backfill', Engineering Geology, 114(3-4), 397-413.

Fehrsen, M. and Cooke, R. (2006) 'Paste fill pipeline distribution systems–Current status', Rise of
the Machines–The State of the Art in Mining, Mechanisation, Automation, Hydraulic
Transport and Communications.

Ghirian, A. (2016) Coupled Thermo-Hydro-Mechanical-Chemical (THMC) Processes in


Cemented Tailings Backfill Structures and Implications for Their Engineering Design,
unpublished thesis Université d'Ottawa/University of Ottawa.

Grice, T. (1998) 'Underground mining with backfill', 2nd Annual Summit œ Mine Tailings Disposal
Systems, Brisbane, Nov, 24-25.

Haruna, S. and Fall, M. (2019) 'Time- and temperature-dependent rheological properties of


cemented paste backfill that contains superplasticizer', Powder Technology.

Mitchell, R. J. (1983) Earth structures engineering, Chapter 6., Allen & Unwin, London, United
Kingdom.

Paterson, A. (2012) 'Pipeline transport of high density slurries: a historical review of past mistakes,
lessons learned and current technologies', Mining Technology, 121(1), 37-45.

Yilmaz, E., Belem, T., Bussière, B., Mbonimpa, M. and Benzaazoua, M. (2015) 'Curing time effect
on consolidation behaviour of cemented paste backfill containing different cement types
and contents', Construction and Building Materials, 75, 99-111.

119
CHAPTER 6
Conclusions and Recommendations

120
6.1 Conclusions
Following findings have been drawn from the results obtained in this study: -

• There was a gradual increase in rheological properties (yield stress and viscosity) of
CPB with progress of curing time, regardless of the type of the hydraulic binder, initial
temperature, and/or dosage of sodium silicate.

• The increase in initial temperature will significantly increase the rheological


properties, and thus reduce the flowability of CPB.

• Yield stress and viscosity of CPB increase with the increase in the sodium silicate
content.

• The combination of high initial temperature and high sodium silicate content
significantly accelerates the rate of increase in the rheological properties of CPB.

• The partial usage of Slag or FA enhances the flowability of CPB by reducing the yield
stress and viscosity.

• Using PCI-FA blend as binder in CPB that contains sodium silicate better enhances
the flowability of CPB as compared to CPB using the PCI-Slag blend.

• The mechanical strength of CPB that is prepared with PCI-Slag blend as binder
increases with progress of curing time, regardless of the variation of curing
temperature and/or dosage of sodium silicate.

• Slag-CPB samples that contain sodium silicate showed higher UCS values, and thus
higher strength, when cured at high temperature (e.g. 35oC), and significantly lower
UCS values when cured at low temperature (2oC).

• Adding sodium silicate to the slag-CPB potentially increases the strength of slag-CPB
at any temperature, as the sodium silicate activates the pozzolanic binder (slag).

• High dosage of sodium silicate significantly increases the UCS of slag-CPB, especially
when cured at high temperature.

• The results obtained in this research can contribute to efficient, cost effective, and safer
design of CPB structures in mining area.

• The obtained results can reduce the financial losses associated with unsuitable
flowability of CPB transported in the CPB delivering system.

121
6.2 Recommendations
Taking into consideration the time limitations of the experimental tests conducted in this study,
the following recommendations are proposed for future works.

• Investigating the combined effect of curing time, initial temperature, binder type, and
dosage of sodium silicate on the rheological properties and strength development of
CPB material that contains different chemical elements, such as sulphate, sodium
chloride.

• Investigating the combined effect of curing time, binder type, and dosage of sodium
silicate on the rheological properties and strength development of CPB material that is
prepared and curried at sub-zero temperatures.

• Investigating the combined effect of the tested conditions on the rheological properties
and strength development of CPB prepared with different mix design than those
applied in this study, such as:

✓ Using natural tailings instead of silica tailings

✓ Using mine processing water instead of fresh (tap) water

✓ Water-to-cement ratio that is different than the one used in this study

✓ Different binder blends and/or content than those used in this study

• Studying the combined effect of curing time, temperature, and binder type/content of
CPB that contains different chemical additives or activators, such as superplasticizer,
or sodium hydroxide etc. on the rheological properties and/or strength development of
CPB.

122

You might also like